2023 ESH Guidelines for the management of arterial hypertension The Task Force for the management of arterial hypertension of the European Society of Hypertension: Endorsed by the International Society of Hypertension (ISH) and the European Renal Association (ERA) : Journal of Hypertension

Journal Logo

ESH GUIDELINES

2023 ESH Guidelines for the management of arterial hypertension The Task Force for the management of arterial hypertension of the European Society of Hypertension

Endorsed by the International Society of Hypertension (ISH) and the European Renal Association (ERA)

Mancia, Giuseppea,∗; Kreutz, Reinholdb,∗; Brunström, Mattiasc; Burnier, Micheld; Grassi, Guidoe; Januszewicz, Andrzejf; Muiesan, Maria Lorenzag; Tsioufis, Konstantinosh; Agabiti-Rosei, Enricoi; Algharably, Engi Abd Elhadyb; Azizi, Michelj,k; Benetos, Athanasel; Borghi, Claudiom; Hitij, Jana Brguljann; Cifkova, Renatao,p; Coca, Antonioq; Cornelissen, Veroniquer; Cruickshank, J. Kennedys; Cunha, Pedro G.t,u; Danser, A.H. Janv; Pinho, Rosa Maria dew; Delles, Christianx; Dominiczak, Anna F.y; Dorobantu, Mariaz; Doumas, Michalisaa; Fernández-Alfonso, María S.bb,cc; Halimi, Jean-Micheldd,ee,ff; Járai, Zoltángg; Jelaković, Bojanhh; Jordan, Jensii,jj; Kuznetsova, Tatianakk; Laurent, Stephanell; Lovic, Draganmm; Lurbe, Emparnn,oo,pp; Mahfoud, Felixqq,rr; Manolis, Athanasiosss; Miglinas, Mariustt,uu; Narkiewicz, Krzystofvv; Niiranen, Teemuww,xx; Palatini, Paoloyy; Parati, Gianfrancozz,aaa; Pathak, Atulbbb; Persu, Alexandreccc; Polonia, Jorgeddd; Redon, Josepoo,eee,fff; Sarafidis, Pantelisggg; Schmieder, Rolandhhh; Spronck, Bartiii; Stabouli, Stellajjj; Stergiou, Georgekkk; Taddei, Stefanolll; Thomopoulos, Costasmmm; Tomaszewski, Maciejnnn,ooo; Van de Borne, Philippeppp; Wanner, Christophqqq; Weber, Thomasrrr; Williams, Bryansss; Zhang, Zhen-Yuttt; Kjeldsen, Sverre E.uuu

Author Information
Journal of Hypertension 41(12):p 1874-2071, December 2023. | DOI: 10.1097/HJH.0000000000003480

Abstract

Document Reviewers: 

Luis Alcocer (Mexico), Christina Antza (Greece), Mustafa Arici (Turkey), Eduardo Barbosa (Brazil), Adel Berbari (Lebanon), Luís Bronze (Portugal), John Chalmers (Australia), Tine De Backer (Belgium), Alejandro de la Sierra (Spain), Kyriakos Dimitriadis (Greece), Dorota Drozdz (Poland), Béatrice Duly-Bouhanick (France), Brent M. Egan (USA), Serap Erdine (Turkey), Claudio Ferri (Italy), Slavomira Filipova (Slovak Republic), Anthony Heagerty (UK), Michael Hecht Olsen (Denmark), Dagmara Hering (Poland), Sang Hyun Ihm (South Korea), Uday Jadhav (India), Manolis Kallistratos (Greece), Kazuomi Kario (Japan), Vasilios Kotsis (Greece), Adi Leiba (Israel), Patricio López-Jaramillo (Colombia), Hans-Peter Marti (Norway), Terry McCormack (UK), Paolo Mulatero (Italy), Dike B. Ojji (Nigeria), Sungha Park (South Korea), Priit Pauklin (Estonia), Sabine Perl (Austria), Arman Postadzhian (Bulgaria), Aleksander Prejbisz (Poland), Venkata Ram (India), Ramiro Sanchez (Argentina), Markus Schlaich (Australia), Alta Schutte (Australia), Cristina Sierra (Spain), Sekib Sokolovic (Bosnia and Herzegovina), Jonas Spaak (Sweden), Dimitrios Terentes-Printzios (Greece), Bruno Trimarco (Italy), Thomas Unger (The Netherlands), Bert-Jan van den Born (The Netherlands), Anna Vachulova (Slovak Republic), Agostino Virdis (Italy), Jiguang Wang (China), Ulrich Wenzel (Germany), Paul Whelton (USA), Jiri Widimsky (Czech Republic), Jacek Wolf (Poland), Grégoire Wuerzner (Switzerland), Eugene Yang (USA), Yuqing Zhang (China).

Erratum

In ESH 2023 Guidelines , the color in columns 3, 4 and 5 of the row for Class of Recommendation I in Figure 2 has been corrected from white to dark blue in the online version of the paper to match the color used in all other Tables referring to the Level of Evidence. The color of the text has been changed from black to white to aid reading. The corrected Figure 2 is shown below. Additionally, in section 7.5 on page 1916, ‘... to minimize the problem of postexercise hypotension [411–413].’ should be changed to ‘... due to the immediate BP lowering effects of exercise [411–413].’

5+ images

Journal of Hypertension. 42(1):194, January 2024.

Table of Contents

INTRODUCTION          10

WHAT IS NEW AND WHAT HAS CHANGED IN THE 2023 EUROPEAN SOCIETY OF HYPERTENSION ARTERIAL HYPERTENSION GUIDELINES?          12

1. Methodology and definition of evidence          14

 1.1 Methodology of evidence grading          14

 1.2 Level of evidence          14

2. PRINCIPLES OF HYPERTENSION PATHOPHYSIOLOGY          16

3. DEFINITION OF HYPERTENSION AND BP CLASSIFICATION          18

 3.1 Definition of hypertension          18

 3.2 Classification of hypertension          19

 3.3 Prevalence of hypertension          20

 3.4 BP relationship with risk of cerebral, cardiovascular and kidney events          20

 3.5 Hypertension and total CV risk assessment          21

 3.6 Screening versus case finding in the detection of hypertension          24

 3.7 Confirming the diagnosis of hypertension          25

4. BP MEASUREMENT AND MONITORING          26

 4.1 Devices for blood pressure measurement          26

  4.1.1 Standard cuff-based devices          26

  4.1.2 Cuffless blood pressure measuring devices          27

  4.1.3 Validation of blood pressure measuring devices          27

 4.2 Standard office blood pressure measurement          28

 4.3 Unattended office blood pressure measurement          30

 4.4 Blood pressure during exercise          32

 4.5 Blood pressure measurement in hospital          32

 4.6 Central blood pressure          33

 4.7 Home blood pressure monitoring          34

 4.8 Ambulatory blood pressure monitoring          36

 4.9 Clinical indications for HBPM and ABPM          39

 4.10 Blood pressure variability          41

5. PATIENT WORK-UP          42

 5.1 Personal and medical history          42

 5.2 Physical examination          44

 5.3 Routine clinical chemistry investigations          45

 5.4 Other investigations in hypertension          45

  5.4.1 Electrocardiogram          45

  5.4.2 Kidney ultrasound          46

  5.4.3 Selected biomarkers and genetic markers          46

   5.4.3.1 Lipoprotein (a)          46

   5.4.3.2 Cardiac biomarkers          46

   5.4.3.3 Kidney markers          47

   5.4.3.4 Genetic markers          47

5.5 Assessment of hypertension-mediated organ damage          48

  5.5.1 HMOD in the heart          49

   5.5.1.1 Left ventricular mass and geometry          50

   5.5.1.2 Transthoracic echocardiography          50

   5.5.1.3 Cardiac magnetic resonance          51

   5.5.1.4 Computed tomography of the heart          52

  5.5.2 HMOD in the arteries          52

   5.5.2.1 Carotid artery IMT and plaques          52

   5.5.2.2 Pulse wave velocity          53

   5.5.2.3 Ankle–brachial index          55

  5.5.3 HMOD in the kidney          55

  5.5.4 HMOD in the brain          56

  5.5.5 HMOD in the eye          58

 5.6 Using HMOD to help stratify risk in hypertensive patients          60

 5.7 When to refer a patient to a specialist or for hospital-based care          62

6. SECONDARY HYPERTENSION          63

 6.1 Genetic causes of secondary hypertension          67

7. LIFESTYLE INTERVENTIONS          69

 7.1 Relevance of lifestyle changes          69

 7.2 Weight reduction          70

 7.3 Restriction of sodium intake          70

 7.4 Augmentation of dietary potassium intake          71

 7.5 Increase levels of daily physical activity and regular exercise          72

 7.6 Moderation of alcohol intake          73

 7.7 Smoking cessation          74

 7.8 Other dietary interventions          75

 7.9 Improve stress management          75

 7.10 Exposure to noise and air pollution          76

8. BENEFITS OF ANTIHYPERTENSIVE TREATMENT          77

9. ANTIHYPERTENSIVE DRUG TREATMENT INITIATION          79

 9.1 Should treatment initiation be based on total CV risk?          79

 9.2 Office BP thresholds for initiation of drug treatment          79

 9.3 Should BP-lowering treatment be initiated in patients with office BP <140/90 mmHg?          81

 9.4 Drug treatment initiation in older people          82

10. OFFICE BP TARGETS FOR TREATMENT          85

 10.1 Office BP targets in the general hypertensive population          85

 10.2 Office versus home and ambulatory BP targets          87

 10.3 Timing of BP control and time in therapeutic range          88

 10.4 Residual risk          88

 10.5 Challenges associated with evidence on BP targets          89

11. ANTIHYPERTENSIVE DRUGS AND TREATMENT          92

 11.1 Blockers of the renin–angiotensin system          96

  11.1.1 Angiotensin-converting-enzyme inhibitors          96

  11.1.2 Angiotensin receptor blockers          96

  11.1.3 Renin inhibitors          96

  11.1.4 Combination of RAS inhibitors          97

 11.2 Calcium channel blockers          97

  11.2.1 Dihydropyridine CCBs          97

  11.2.2 Nondihydropyridine CCBs          97

 11.3 Diuretics          97

  11.3.1 Thiazide and Thiazide-like diuretics          97

  11.3.2 Loop diuretics          99

  11.3.3 Potassium-sparing diuretics          99

 11.4 Mineralocorticoid receptor antagonists          99

 11.5 Beta-blockers          100

 11.6 Alpha-1 blockers          102

 11.7 Centrally acting drugs          103

 11.8 Vasodilators          103

 11.9 Angiotensin receptor-neprilysin inhibitor          103

 11.10 Antihypertensive drug combinations          104

  11.10.1 Impact on hypertension drug treatment strategy          104

  11.10.2 Drug combinations          106

  11.10.3 Rationale for initial two-drug combination therapy          110

  11.10.4 Up-titration of treatment to three-drug combination          112

  11.10.5 Rationale for single-pill combination therapy          112

  11.10.6 The quadpill concept          113

  11.10.7 The polypill concept          114

  11.10.8 Choice of drug combinations for initiation of treatment          115

  11.10.9 Tolerability and side effects of drugs          115

  11.10.10 Prescribing of antihypertensive drugs          116

   11.10.10.1 Standard drug administration          116

   11.10.10.2 Partial treatment reduction or complete withdrawal          117

   11.10.10.3 Antihypertensive drugs and cancer risk          118

   11.10.11 Concomitant medications          119

   11.10.11.1 LDL-cholesterol lowering          119

   11.10.11.2 Antiplatelet therapy          121

 12. TRUE RESISTANT HYPERTENSION          122

  12.1 Definition, prevalence, pathophysiology and cardiovascular risk          122

  12.2 Diagnostic work-up          123

  12.3 Optimizing lifestyle interventions and ongoing drug therapy          125

  12.4 Fourth and subsequent lines of antihypertensive therapy          126

 13. DEVICE-BASED TREATMENT OF HYPERTENSION          131

  13.1 Renal denervation          131

   13.1.1 Clinical evidence of the BP-lowering effect of RDN          131

   13.1.2 Off-medications studies          132

   13.1.3 On-medications studies          132

   13.1.4 Safety          133

   13.1.5 Durability          134

   13.1.6 Application          134

  13.2 Carotid baroreceptor stimulation          136

  13.3. Other device-based treatments          136

 14. SPECIFIC HYPERTENSION PHENOTYPES          137

  14.1 Sustained hypertension and true normotension          137

  14.2 White-coat hypertension          138

  14.3 Masked hypertension          139

  14.4 White-coat uncontrolled hypertension and masked uncontrolled hypertension          141

  14.5 Isolated systolic hypertension of the young          142

  14.6 Isolated systolic hypertension in older persons          144

  14.7 Isolated diastolic hypertension          146

  14.8 Night-time hypertension and dipping          147

  14.9 Orthostatic hypertension and hypotension          149

  14.10 Baroreflex failure and efferent autonomic failure          150

 15. HYPERTENSION IN DIFFERENT DEMOGRAPHIC SITUATIONS          152

  15.1 Blood pressure in children, adolescents and transition period          152

   15.1.1 Blood pressure measurements in children and adolescents          152

   15.1.2 Hypertension in children and adolescents          153

   15.1.3 Transition period to adulthood          154

  15.2 Hypertension in young adults          155

  15.3 Hypertension in older persons          156

   15.3.1 Patients 65–79 years old          157

    15.3.1.1 Threshold and target for drug treatment          157

    15.3.1.2 Antihypertensive treatment strategies          158

    15.3.1.3 Antihypertensive drugs          158

    15.3.1.4 Monitoring the effects of treatment          159

   15.3.2 Patients 80 years old or beyond          159

   15.3.3 How to assess the level of frailty/functionality to better personalize therapeutic strategies          160

 15.4 Sex and gender aspects in hypertension          163

  15.4.1 Epidemiology and pathophysiology          163

  15.4.2 Blood pressure and cardiovascular risk          164

  15.4.3 Differences in clinical phenotypes          164

   15.4.3.1 White-coat hypertension and masked hypertension          164

   15.4.3.2 Hypertension-mediated organ damage          164

  15.4.4 Sex differences in hypertension outcomes          165

  15.4.5 Benefits of antihypertensive treatment and target blood pressure          166

  15.4.6 Sex differences in hypertension management          167

  15.4.7 Infertility treatments and hypertension in women          167

  15.4.8. Oral contraceptive pills and hypertension          168

  15.4.9 Hormone-replacement therapy and hypertension          169

  15.4.10 Gender-affirming hormone therapy and hypertension          169

  15.5 Hypertension and ethnicity          169

   15.5.1 Nomenclature and relevance          169

   15.5.2 Management          170

 16. HYPERTENSION IN SPECIFIC SETTINGS          171

  16.1 Hypertension disorders in pregnancy          171

  16.1.1 Definition and classification of hypertension in pregnancy          173

  16.1.2 Blood pressure measurement in pregnancy          174

  16.1.3 Laboratory examinations in pregnancy          175

  16.1.4 Prediction and prevention of preeclampsia          175

  16.1.5 Lifestyle interventions          176

  16.1.6 Clinical management of hypertension in pregnancy          176

   16.1.6.1 Mild preexisting essential hypertension          176

   16.1.6.2 Mild gestational hypertension          177

   16.1.6.3 Preeclampsia          178

   16.1.6.4 Severe hypertension          178

   16.1.6.5 Preexisting secondary hypertension          179

  16.1.7 Blood pressure during puerperium          180

  16.1.8 Postpartum hypertension and breastfeeding          181

  16.1.9 Risk of recurrence of hypertensive disorders in a subsequent pregnancy          181

  16.1.10 Long-term cardiovascular consequences of hypertensive disorders in pregnancy          181

 16.2 Hypertensive urgencies and emergencies          183

   16.2.1 Definitions of hypertensive urgencies and emergencies          183

   16.2.2 Hospital work-up, treatments and follow-up          185

   16.2.3 Blood pressure in the emergency department          189

  16.3 Perioperative hypertension and its management          190

 17. HYPERTENSION IN ESTABLISHED CARDIOVASCULAR DISORDERS          192

  17.1 Coronary artery disease          192

   17.1.2 Treatment of hypertensive patients with coronary artery disease          193

  17.2 Heart failure          195

   17.2.1 Prevention of heart failure in hypertension          195

   17.2.2 Heart failure with reduced ejection fraction          197

   17.2.3 Heart failure with preserved ejection fraction          199

   17.2.4 Overall management of heart failure and classification          200

  17.3 Hypertension and atrial fibrillation          201

   17.3.1 Blood pressure measurement in atrial fibrillation          201

   17.3.2 Detection of atrial fibrillation          202

   17.3.3 Prevention and treatment of atrial fibrillation in hypertension          202

   17.3.4 Oral anticoagulation and BP control          204

  17.4 Valvular heart disease          206

   17.4.1 Aortic stenosis          206

   17.4.2 Aortic regurgitation          207

   17.4.3 Mitral regurgitation          207

  17.5 Cerebrovascular disease and cognition          208

   17.5.1 Management of elevated BP in acute stroke          208

    17.5.1.1 Acute hemorrhagic stroke          208

    17.5.1.2 Acute ischemic stroke          209

   17.5.2 Management of elevated BP in patients with previous stroke or transient ischemic attack          210

   17.5.3 Management of patients with cognitive dysfunction and dementia          211

  17.6 Vascular disease          213

   17.6.1 Lower extremity arterial disease          213

   17.6.2 Aortic dilatation, aneurysm and dissection          214

 18. HYPERTENSION AND DIABETES MELLITUS          215

  18.1. Epidemiology and risk classification          215

  18.2 Benefits of BP-lowering          216

  18.3 Antihypertensive drug treatment          217

 19. HYPERTENSION AND THE KIDNEY          219

  19.1 Treatment of hypertension in CKD          219

   19.1.1 Treatment BP targets          220

   19.1.2 Antihypertensive drug treatment          222

   19.1.3 Special therapeutic challenges          223

   19.1.4 Use of additional drugs for cardiovascular and nephroprotection in CKD          224

  19.2 Renovascular disease          227

  19.3 Hypertension in patients with kidney transplantation          228

 20. HYPERTENSION AND OTHER SELECTED COMORBIDITIES          231

  20.1 Obesity          231

   20.1.1 Antihypertensive pharmacotherapy in obesity          231

   20.1.2 Role of nonpharmacological weight loss intervention          232

   20.1.3 Role of weight loss medications          233

   20.1.4 Role of bariatric surgery          234

  20.2 Obstructive sleep apnea          235

  20.3 Asthma          236

  20.4 Obstructive pulmonary disease          237

  20.5 Gout and uric acid          238

  20.6 Immune-mediated inflammatory diseases          238

   20.6.1 Rheumatoid arthritis          239

   20.6.2 Psoriasic arthritis          239

   20.6.3 Systemic lupus erythematosus          240

  20.7 Glaucoma in hypertension          240

  20.8 Hypertension oncology          242

   20.8.1 Hypertension and its association with cancer          242

   20.8.2 Hypertension induced by cancer treatments          243

    20.8.2.1 Hypertension induced by VEGF inhibitors          243

    20.8.2.2 Hypertension induced by other anticancer drugs          244

    20.8.2.3 Hypertension induced by adjunctive therapies, radiotherapy or surgery          245

   20.8.3 Management of hypertension in cancer patients          245

    20.8.3.1 BP monitoring and general management before start of cancer treatment          245

    20.8.3.2 General BP-lowering therapy and management during cancer therapy          247

    20.8.3.3 Treatment of hypertension induced by VEGF inhbitors          248

   20.8.4 Follow-up and management of hypertension in cancer survivors          250

  20.9 COVID-19 and hypertension          250

   20.9.1 COVID-19 and RAS inhibitors          251

   20.9.2 COVID-19 lockdown and hypertension management          252

   20.9.3 Vaccination against SARS-CoV-2 and hypertension          252

   20.9.4 Long COVID-19 and hypertension          253

 21. FOLLOW-UP          253

  21.1 Importance of follow-up          253

  21.2 Adherence          255

   21.2.1 Definitions          255

   21.2.2 Prevalence of nonadherence and associated burden          255

   21.2.3 Methods to detect nonadherence to antihypertensive treatment          256

   21.2.4 Etiology of nonadherence to antihypertensive treatment          257

   21.2.5 When and how to screen for nonadherence          257

   21.2.6 Management of nonadherence to antihypertensive treatment          258

  21.3 Clinical inertia          259

  21.4 Patient empowerment          261

  21.5 Follow-up of low-risk hypertensive patients and deprescribing          262

  21.6 Use of telemedicine and tele-health technologies          262

  21.7 Challenges of long-term follow-up          263

  21.8 Role of general physician, pharmacies and team-based care          263

  21.9 Hypertension clinics          265

  21.10 Health risks at workplace          265

  21.11 Patient organizations          266

 22. GAPS IN EVIDENCE AND FUTURE OPPORTUNITIES          268

 23. REFERENCES          270

INTRODUCTION

The year 2023 marks the 20th anniversary of the hypertension guidelines of the European Society of Hypertension (ESH), which were published for the first time in 2003, following a proposal by Professor Alberto Zanchetti (Fig. 1). Professor Zanchetti thought that it was time for Europe to express its view on diagnostic and treatment aspects of this crucially important medical condition rather than referring, as in the past, to guidelines issued by the WHO, with or without the ISH or the scientific societies in the USA. He played a fundamental role in these first guidelines [1] as coordinator of the Writing Committee appointed by the ESH, and this was rewarded by an unexpected large success, which made these guidelines the fifth most widely quoted paper in the world across all research areas and the most quoted in the medical area. ESH offered to share these hypertension guidelines with the European Society of Cardiology (ESC), which accepted after the manuscript had already been completed, without sharing its publication in the ESC Society Journal. Subsequently, the ESH and the ESC enjoyed an equal collaboration, resulting in three further successful and also widely quoted guidelines in 2007 [2], 2013 [3] and 2018 [4] that were published in the official journals of the two Societies, except for a 2009 reappraisal of the 2007 guidelines, which was prompted by new evidence in the hypertension area and prepared only by the ESH [5].

F1
FIGURE 1:
Alberto Zanchetti.

These 2023 hypertension guidelines have also been prepared only by the ESH because the ESC did not want to continue the previous understanding with ESH to generate “Joint Guidelines” with the equalparticipation of ESH and ESC. The rules of these guidelines, however, are largely, although not entirely, the same as those that were followed in the previous guidelines. That is, in the 2023 guidelines: (i) the members of the Task Force have been appointed by the ESH, based on recognized scientific and clinical expertise in one or more areas covered by the guidelines as well as on the documented absence of relevant conflicts of interest; (ii) selected members were initially asked to write a section or sections of the guidelines related to her or his main scientific expertise, and a small Steering Committee was appointed to harmonize the material received; (iii) multiple revisions of the text were made by back and forth interactions between the Task Force members, with a final collective critical review of the text and (iv) the final manuscript has been sent to external reviewers and further revised according to their suggestions and criticism. Particular attention has been given to the scoring of the strength of the diagnostic and treatment recommendations, which have been graded according to criteria partly different from those used in previous guidelines, i.e. with consideration for the study design but also for the quality of the collected data (see Section 1). Because of the questionable scientific value of voting, disagreements on treatment recommendations have not been resolved that way but by consensus on a shared text. Conflicting evidence or interpretation of the data have been openly admitted.

The similarity of the present and past guidelines extends to the scientific principles on which the guidelines have been based. The 2023 guidelines have been developed after careful search for new studies in the hypertension and related areas. Furthermore, as in the past, RCTs have been assigned a top value while also mentioning their limits when appropriate. However, all other relevant sources of knowledge (from observational studies down to clinical cases) have been considered, and even mechanistic studies have not been ignored, given their relevance for diagnostic and treatment decisions in individual patients. Particular attention has been given to real-world studies, which play a growing role in hypertension research and provide knowledge in areas that cannot be addressed by RCTs. Like the previous guidelines, the 2023 guidelines (i) regard their value as educational, which explains why the text addresses the data justifying the recommendations and (ii) emphasize that their recommendations are not invariably prescriptive for individual patients because they are based on average data and address conditions or diseases in general. In individual patients, the most appropriate diagnostic and treatment decisions may differ from those expressed by the guidelines.

The 2023 guidelines (i) contain several conceptual elements of novelty originated by research performed after the 2018 guidelines; (ii) deal more in depth with topics that were only briefly considered in the past and (iii) extend to several conditions that were previously unaddressed by guidelines, although frequently coexisting with hypertension and leading to specific needs for medical management. Although mainly referring to hypertension in adults, they include for the first time essential recommendations on hypertension in children, adolescents and the transition to young hypertensive adult individuals; and (iv) include a detailed index of sections and subsections focused on specific issues that has been prepared to facilitate reading of these various and multiple aspects. Furthermore, while the text addresses the sometimes nonunivocal evidence provided by research on a given issue, each section offers, as is now usual for many guidelines, a simple final list of key statements and recommendations that translate research achievements into practical use. We hope that this structure will make the ESH guidelines useful not only to the practicing physicians but also to hypertension experts and investigators.

WHAT IS NEW AND WHAT HAS CHANGED IN THE 2023 EUROPEAN SOCIETY OF HYPERTENSION ARTERIAL HYPERTENSION GUIDELINES?

  • 1. Modified and simplified criteria for evidence grading recommendations
  • 2. Pathophysiological background of primary hypertension
  • 3. Clinical BP measurements by different methods and in different settings and clinical conditions
  • 4. Thorough description of office, ambulatory and home BP measurements and value in different demographic and clinical conditions
  • 5. Upgrading of out-of-office BP measurements in hypertension management
  • 6. New HMOD measurements and their clinical value in hypertension work-up
  • 7. New CV risk factors and update on CV risk assessment
  • 8. Update and comprehensive summary of secondary forms of hypertension
  • 9. Update on lifestyle interventions
  • 10. Update on threshold and targets for antihypertensive drug treatment, including their possible heterogeneity in demographic and clinical subgroups of patients
  • 11. Confirmation of preferred use of RAS blockers, CCBs and Thiazide/Thiazide-like diuretics, and their various combinations for BP-lowering treatment. Inclusion of BBs among the major antihypertensive drugs
  • 12. Update on available combination-based drug treatment strategies, including the quadpill and the polypill
  • 13. Emphasis and update on the diagnosis and management of true resistant hypertension
  • 14. Update on use and position of renal denervation for antihypertensive treatment
  • 15. Impact of hypertension and its treatment on cognitive dysfunction and dementia
  • 16. Management of hypertension in older people according to the frailty and functional level
  • 17. Update on treatment of hypertension in HFrEF and HFpEF
  • 18. New diagnostic approaches to diagnosis and treatment in hypertensive patients with AF
  • 19. Update on treatment in CKD, including kidney transplantation
  • 20. Update and novel treatment approaches to patients with type 2 diabetes
  • 21. Epidemiology, diagnosis and treatment in different BP phenotypes
  • 22. Diagnosis, treatment and follow-up of hypertension in demographic and clinical conditions not or only marginally addressed in previous guidelines:
    • a. Children/adolescents and transition to adulthood
    • b. Young patients
    • c. Sex-related differences
    • d. Pregnancy and puerperium
    • e. Peripheral artery disease
    • f. Aortic aneurism
    • g. Valvular heart disease
    • h. Treatment of hypertension in acute cerebrovascular diseases
    • i. Hypertensive emergencies/urgencies
    • j. Perioperative hypertension
    • k. Obesity
    • l. COVID-19
    • m. Chronic inflammatory diseases
    • n. Hypertension in oncology
    • o. Baroreflex failure and dysautonomia
    • p. Glaucoma
  • 23. Detailed recommendations on patients’ follow-up strategies, including assessment and minimization of nonadher-ence and clinical inertia.
  • 24. Mention of new potential approaches to the treatment of hypertension and containment of hypertension-related workload (tele-health, team-based treatment, role of pharmacists)

1. METHODOLOGY AND DEFINITION OF EVIDENCE

1.1 Methodology of evidence grading

The 2023 ESH guidelines aim to summarize the best available evidence for all aspects of hypertension management. The guidelines were developed by a Task Force of 59 experts form European countries, representing the areas of internal medicine, cardiology, nephrology, endocrinology, general medicine, geriatrics, pharmacology and epidemiology. Each topic was assigned to a small group of Task Force members responsible for reviewing and summarizing the available evidence within that topic. The ‘class of recommendation’ (CoR) and ‘level of evidence’ (LoE) for all recommendations were reviewed by an Evidence Grading Committee to make sure that they complied with the predefined criteria outlined in the following. Draft versions were reviewed by the Steering Committee, Task Force members and external reviewers. The final version was approved by all Task Force members.

In accordance with previous versions of the ESH guidelines, a similar system separating CoR and LoE was applied [3,4]. CoR indicates how strong a recommendation is, considering the assumed benefit versus risks and costs on a scale from I to III. Recommendation classes I and III each convey a clear message, namely a general consensus that a measure is either useful (CoR I) or not useful or even harmful (CoR III). If there is no general consensus or only doubtful evidence, an optional recommendation is conveyed with CoR II. In contrast to previous guidelines [3,4], the Task Force finds that a further subdivision of the CoR II into two subclasses (IIa and IIb) adds little value and, for the sake of simplification is no longer used. The LoE indicates how reliable the evidence underlying each recommendation is on a scale from A to C (Fig. 2). Importantly, the CoR and LoE are independent of each other, e.g. strong recommendations may build on weak evidence if the assumed benefit of an intervention or a diagnostic procedure greatly outweighs the potential risks.

F2
FIGURE 2:
Class of recommendation (CoR) and level of evidence (LoE). BP, blood pressure, CVD, cardiovascular disease, HMOD, hypertension mediated organ damage, RCT, randomized controlled trial. aLimitations affecting the level of evidence include (but may not be limited to) high risk of bias, inability to account for important confounding factors in observational studies, questionable external validity and uncertain effect estimates (confidence intervals including negligible effect).

1.2 Level of evidence

The 2023 ESH guidelines employ the same terminology as in the 2018 ESC/ESH guidelines but with updated criteria for assessing the LoE. This revision was influenced primarily by the recommendations from the GRADE working group [6,7], but also by the most recent evidence definition used by the AHA/ACC [7].

The most important difference compared with the previous guidelines is the priority given to patient-important CV outcomes, such as stroke, MI, HF, ESKD and CV or total mortality, acknowledging that the primary aim of antihypertensive treatment is to reduce the risk of clinical outcomes and not only BP. Although BP reduction is strongly associated with a reduction in clinical events, interventions affecting BP may also affect other physiological systems with beneficial or harmful effects, and the benefit/harm ratio cannot be firmly established without clinical outcome trials.

Furthermore, risk of bias and statistical precision were considered when assigning the LoE. This means that recom-mendations supported by well conducted RCTs with CV outcomes were assigned LoE A, whereas recommendations supported by trials with a similar design and with similar outcomes, were downgraded to LoE B or C if the risk of bias was judged as high, or if effect estimates were imprecise. Meta-analyses may contribute to any level of evidence depending on the type of studies included and the quality of the meta-analysis itself [8].

For diagnostic tests and procedures, we have adopted the strategy recommended by the GRADE working group, assessing the evidence for benefit on patient-important outcomes [9]. Many diagnostic procedures rely on studies of accuracy rather than effect on outcomes, and recommendations building on such evidence is generally downgraded for indirectness even if the studies themselves are without important limitations.

2. PRINCIPLES OF HYPERTENSION PATHOPHYSIOLOGY

Hypertension is divided into a primary (formerly and still also currently referred to as ‘essential’) and secondary forms. Secondary hypertension originates from specific causes and can be detected in only a small fraction of hypertensive patients (see Section 6). Primary hypertension covers the remaining large fraction of the hypertensive population, and its origin depends on the complex interaction between a genetic background, a large number of environmental factors [10–12] and the aging process. Both genetic and environmental factors operate through alterations of CV regulatory systems, leading to an increase of systemic vascular resistance, which is the hallmark hemodynamic abnormality responsible for BP elevation in almost all hypertensive patients [13]. In the last few years, important new evidence has been obtained on the genetic background of hypertension, with more than 1000 genetic factors being identified [11,12] together with, in some instances, the biochemical and pathophysiological paths they work through [14]. New environmental factors (e.g. air pollution and noise) have been added to those already documented by older research [15–17]. Furthermore, new experimental and clinical studies have confirmed that alterations of several major CV control systems may contribute to chronic BP elevation. As shown in Fig. 3, primary hypertension may be accompanied by alterations of the RAAS, central and peripheral autonomic cardiac and vascular regulation, the endothelin system and other systems controlling vascular function, including nitric oxide and natriuretic peptides [13,18–22]. More recently, pressogenic effects (increased sodium sensitivity) of gut microbial dysbiosis have also been reported [23,24]. In addition, the immune system is likely to play a pathophysiologic role, with effects that are possibly primarily mediated by inflammation, and involve not only BP regulation (and thus development of hypertension) but also the initiation and progression of HMOD [25,26]. There is extensive experimental and clinical evidence that hypertension is associated with inflammation and immune cell activation, two processes that are driven in large part by oxidative stress. Immune cell activation is characterized by excessive production of reactive oxygen species and an altered oxidation– reduction (redox) state [26], and there is evidence that generation of reactive oxygen species is influenced by factors involved in BP regulation, such as Ang II, endothelin-1 (ET-1), aldosterone and salt (sodium) [26]. Furthermore, evidence is also available that alterations of immunoinflammation is promoted by the above-mentioned hypertension promoters such as genetic susceptibility, neurohumoral activation, salt influences and gut microbiome [10–13,18–22,27]. Although this complex interplay makes it impossible to know whether inflammation is causatively related to hypertension or represents a secondary effect of a chronic BP elevation, it is clear that inflammation and the dysregulated immune system are closely linked to each other and that immunoinflammation is involved in hypertension [25,26]. Indeed, the suggestion has been made that oxidative stress and increased generation of reactive oxygen species represent the common molecular basis linking immunoinflam-mation to hypertension. Alterations in metabolomic pathways, e.g. glucose and lipid metabolism, may also contribute, as exemplified by the sympathostimulating effect of insulin [13,28] and the favoring effect of sympathostimulation on insulin resistance [29]. Regardless of the mechanisms involved, a chronic BP elevation is known to modify the cardiac (e.g. LVH), large artery (increase in collagen and stiffening of the arterial wall) and small artery (increase in wall-to-lumen ratio) structure, which in a later hypertension phase promote the BP increase on a nonspecific anatomical basis [13]. This confirms and expands the former mosaic theory on the pathogenesis of primary hypertension as a multifactorial phenotype, which was already formulated by Page [30] in the pioneer phase of hypertension research more than 70 years ago. To the original theory, modern research has added not only new mechanisms but also, as shown in Fig. 3, strong evidence for the existence of reciprocal influences between different CV control systems, as a result of which alteration of one system may favor or reinforce alterations of the other systems and vice versa [31]. At a practical level, this multimechanistic interactive pathophysiology implies that diagnostic attempts to identify a single responsible mechanism for primary hypertension can often be not only methodologi-cally difficult but also futile. It also explains why an elevated BP can be lowered by drugs with different mechanisms of action as well as why a combination of mechanistically different drugs lowers BP much more effectively than monotherapy.

F3
FIGURE 3:
Mechanisms involved in BP regulation and the pathophysiology of hypertension.

3. DEFINITION OF HYPERTENSION AND BP CLASSIFICATION

3.1 Definition of hypertension

According to the previous 2018 European and current international guidelines [32–34], hypertension is defined based on repeated office SBP values 140 mmHg and/or DBP 90 mmHg. However, there is a continuous relationship between BP and CV or renal morbid or fatal events starting from an office SBP >115 mmHg and a DBP >75 mmHg [35]. Therefore, this definition is arbitrary and has mainly the pragmatic purpose of simplifying the diagnosis and decision on hypertension management. In this context, the above office threshold BP values correspond to the level of BP at which the benefits of intervention (lifestyle interventions or drug treatment) exceed those of inaction, as shown by outcome-based RCTs. Based on available evidence [36] the definition of hypertension remains unchanged from the previous guidelines [4].

3.2 Classification of hypertension

The classification of office BP and definition of hypertension grades also remain the same from previous guidelines (Table 1).

TABLE 1 - Classification of office BP and definitions of hypertension grades
Category Systolic (mmHg) Diastolic (mmHg)
Optimal <120 and <80
Normal 120–129 and 80–84
High-normal 130–139 and/or 85–89
Grade 1 hypertension 140–159 and/or 90–99
Grade 2 hypertension 160–179 and/or 100–109
Grade 3 hypertension ≥180 and/or ≥110
Isolated systolic hypertensiona ≥140 and <90
Isolated diastolic hypertensiona <140 and ≥90
The BP category is defined by the highest level of BP, whether systolic or diastolic.
aIsolated systolic or diastolic hypertension is graded 1, 2 or 3 according to SBP and DBP values in the ranges indicated. The same classification is used for adolescents ≥16 years old (see Section 15.1).

In addition to grades of hypertension, which are based on BP values, we also distinguish stages of hypertension as follows:

Stage 1: Uncomplicated hypertension (i.e. without HMOD or established CVD, but including CKD stage 1 and 2)

Stage 2: Presence of HMOD or CKD stage 3 or diabetes.

Stage 3: Established CVD or CKD stages 4 or 5.

Definition of BP categories, hypertension grades and stages according to office BP

figure4

3.3 Prevalence of hypertension

Hypertension is the most prevalent CV disorder in the world and according to the WHO, it affects 1.28 billion adults aged 30–79 years worldwide, two-thirds living in low-income and middle-income countries. In 2019, the global age-standardized average prevalence of hypertension in adults aged 30– 79 years was reported to be 34% in men and 32% in women [37]. In European countries, the prevalence is similar, with between-country differences and values lower than average in some Western and above average in Eastern European countries [37]. At younger ages (<50 years), hypertension is more prevalent in men, whereas a steeper increase of SBP in women from their third decade (and more so following menopause) makes the prevalence of hypertension greater in women in older age categories (>65 years) [38,39]. SBP increases progressively with age while DBP rises only until the age of 50–60 years, followed by a short period of stagnation and a subsequent mild decrease [40]. This results in an increase of pulse pressure (difference between SBP and DBP) with age [38].

3.4 BP relationship with risk of cerebral, cardiovascular and kidney events

There is a continuous relationship between the increase in BP and the risk of stroke, CAD, HF and development and progression of CKD. This applies to all ages and ethnic groups [41]. In 2002, the Prospective Studies Collaboration Group found that, for each 20 mmHg elevation of office SBP or 10 mmHg elevation of office DBP, the risk for fatal CAD or stroke doubled [35]. SBP is a better predictor of events than DBP after the age of 50 years. In addition to previous studies suggesting that elevated DBP is associated with increased risk in young individuals [42], more recent studies indicate increased risk of CV events for both SBP and DBP elevations in younger adults [43] (see Section 15.2). By reflecting an increase of arterial stiffness, increased pulse pressure was found to be associated with an adverse prognostic impact, additional to that associated with SBP elevation in middle-aged and older people [44,45].

3.5 Hypertension and total CV risk assessment

Hypertension is often associated with other risk factors, including dyslipidemia, impaired glucose tolerance and type 2 diabetes, which further increase CV risk [46,47]. The large number of factors influencing CV risk in patients with hypertension (environmental, lifestyle and clinical CV risk factors plus HMOD and established CVD or CKD categories) are listed in Table 2. Special risk factors apply to women, e.g. hypertension disorders in pregnancy and early-onset menopause [48,49]. HMOD is an important intermediate stage in the CVD continuum between CV risk factors and clinically manifest CVD or advanced CKD stages 4 and 5. HMOD is an important determinant of overall CV risk [1], which is usually high in the presence of HMOD [4]. Diabetes mellitus is listed (Fig. 4) as a separate condition that impacts on CV risk, regardless of the concomitance of HMOD, CVD or CKD. Only diabetic patients with well controlled, short-standing duration of the disease (less than 10 years), no evidence of HMOD and no additional CV risk factors are categorized as at moderate risk [33].

TABLE 2 - Factors that influence CV risk in patients with hypertension
Parameter for risk stratification, which are included in SCORE2 and SCORE2-OP
 Sex (men >women)
 Age
 Level of SBPa
 Smoking – current or past history
 Non-HDL cholesterol
Established and suggested novel factors
 Family or parental history of early onset hypertension
 Personal history of malignant hypertension
 Family history of premature CVD (men aged <55 years; women aged <65 years)
 Heart rate (resting values >80 bpm)
 Low birth weight
 Sedentary lifestyle
 Overweight or Obesity
 Diabetes
 Uric acid
 Lp(a)
 Adverse outcomes of pregnancy (recurrent pregnancy loss, preterm delivery, hypertensive disorders, gestational diabetes)
 Early-onset menopause
 Frailty
 Psychosocial and socioeconomic factors
 Migration
 Environmental exposure to air pollution or noise
Additional clinical conditions or comorbidities
 True resistant hypertension
 Sleep disorders (including OSA)
 COPD
 Gout
 Chronic inflammatory diseases
 Nonalcoholic fatty liver disease (NASH)
 Chronic infections (including long COVID-19)
 Migraine
 Depressive syndromes
 Erectile dysfunction
Hypertension-mediated organ damage (HMOD)
 Increased large artery stiffness:
 Pulse pressure (in older people) ≥60 mmHg
 Carotid–femoral PWV >10 m/s (if available)
 Presence of non-hemodynamically significant atheromatous plaque (stenosis) on imaging
 ECG LVH (Sokolow–Lyon index >35 mm, or R in aVL ≥11 mm; Cornell voltage-duration product (+6 mm in women) >2440 mmms, or Cornell voltage >28 mm in men or >20 mm in women)
 Echocardiographic LVH (LV mass index: men >50 g/m2.7; women >47 g/m2.7 (m = height in meters); indexation for BSA may be used in normal-weight patients: >115 g/m2 in men and >95/m2 in women
 Moderate increase of albuminuria 30–300 mg/24 h or elevated ACR (preferably in morning spot urine) 30–300 mg/g
 CKD stage 3 with eGFR 30–59 ml/min/1.73 m2
 Ankle–brachial index <0.9
 Advanced retinopathy: hemorrhages or exudates, papilledema
Established cardiovascular and kidney disease
 Cerebrovascular disease: ischemic stroke, cerebral hemorrhage, TIA
 Coronary artery disease: myocardial infarction, angina, myocardial revascularization
 Presence of hemodynamically significant atheromatous plaque (stenosis) on imaging
 Heart failure, including heart failure with preserved ejection fraction
 Peripheral artery disease
 Atrial fibrillation
 Severe albuminuria > 300 mg/24 h or ACR (preferably in morning urine) >300 mg/g
 CKD stage 4 and 5, eGFR < 30 mL/min/1.73m2
CKD, chronic kidney disease; COPD, chronic obstructive pulmonary disease; eGFR, estimated glomerular filtration rate; HDL, high-density lipoprotein; LV, left ventricle; OSA, obstructive sleep apnea.
aDBP is not included in the SCORE2/SCORE2-OP tool to estimate CV risk.

F4
FIGURE 4:
Cardiovascular risk according to grade and stage of hypertension.

Estimation of total CV risk is recommended in each hypertensive patient because of its relevance for hypertension management. Computerized methods have been developed for estimating total CV risk, i.e. the likelihood of developing a CV event, usually within the following 10 years. Many risk stratification systems are based on the Framingham study, estimating the 10-year risk for both fatal and nonfatal CAD by SBP and the presence of other risk factors [50]. The Framingham risk stratification is applicable to some European populations [51], but this requires recalibration [52,53], due to geographic differences in the incidence of coronary and stroke events between the European and US populations. Because of these limitations, the SCORE (Systemic Coronary Risk Evaluation) method, based on a large European database, has been developed [54]. SCORE charts estimate the risk of death from CV (not only coronary) disease over 10 years in either high-risk and low-risk European countries [54] and modified charts can, thus, be used for individual countries. The 2021 European Guidelines on CVD prevention made use of SCORE2, which is an updated version of the original SCORE that estimates an individual's 10-year risk of fatal and nonfatal CV events in apparently healthy individuals aged 40– 69 years with risk factors that are untreated or have been stable for several years [55]. For older people (age 70– 89 years), a corresponding SCORE2-OP algorithm for older people is available [56]. SCORE2 and SCORE2-OP are calibrated for four clusters of countries (low, moderate, high, and very high CV risk) that are grouped according to national CV mortality rates published by the WHO.

Estimating a person's 10-year CV risk by SCORE2 and SCORE2-OP is done by first selecting the correct country group and its corresponding risk stratification table [33]. Within the table, the risk is estimated based on sex, age, level of SBP, smoking status, and non-HDL cholesterol level (total cholesterol – HDL cholesterol, Table 2). DBP is not included in the SCORE2/ SCORE2-OP tool, which is a weakness. Overall risk should be stratified in all patients as shown in Fig. 4. Risk stratification is particularly important in individuals with a high-normal BP or grade 1 hypertension, in whom it may influence the decision of whether or how fast to initiate BP-lowering drug treatment. In patients with grade 2 and 3 hypertension, drug treatment should be used regardless of the CV risk level, but risk stratification is nevertheless important for several treatment aspects and FU strategies. The disadvantage of this and other approaches to CV risk quantification is that the estimated risk is usually low in younger adults, particularly in young women who may be stratified as being at low risk even if they have more than one risk factor. Indeed the SCORE2 system does not even provide risk estimations for adult women and men younger than 40 years. By contrast, most older men are considered at high risk, despite being at little increased risk relative to their peers. For young adults, it may be more useful to estimate lifetime risk of CVD and potential CVD-free life-years gained through risk factor optimization [57]. In addition, methods for total CV risk stratification usually underestimate the primary importance of asymptomatic HMOD for the CV risk quantification. In hypertensive patients, HMOD usually indicates a high risk regardless of the organ where the damage is located, and its assessment is, therefore, important for management, particularly in patients who – according to age and general risk stratification – are at apparently low risk [1] (see Section 5).

Risk assessment in hypertension with SCORE2 and SCORE2-OP

figure6

3.6 Screening versus case finding in the detection of hypertension

Because of the high prevalence of hypertension in the general population and its major role as a cause of death and morbidity, its detection is crucially important for public health. Studies performed in different countries have almost universally shown that a considerable fraction of hypertensive individuals is unaware of their condition, with an adverse reflection on the number of patients undergoing treatment and achieving BP control [37]. There is evidence that screening policies can substantially increase the number of individuals in whom hypertension is detected although data on outcome benefit and harm from randomized controlled trials or observational studies at low risk of bias are lacking [58]. However, participation in the screening procedures may be less in some categories (e.g. men, younger individuals, people with lower socioeconomic backgrounds) than in others [59]. In the USA, the US Preventive Services Task Force suggests screening for hypertension in adults aged 18 years or older [60]. Despite limited evidence on the optimal frequency of screening, they recommend yearly screening in adults 40 years old and in those at increased risk for developing hypertension such as blacks, individuals with high-normal BP and people who are overweight or obese [60]. Opportunistic BP measurements are now also encouraged in several countries [59] and have been recently supported in USA by the increased detection, treatment and control of hypertension associated with BP measurements in barber shops or by pharmacists [61]. Based on the available evidence, we recommend opportunistic screening for hypertension in all adults (i.e. 18 years old). Regular BP measurements is particularly important in adults from the age of 40 years and in adults at increased risk for hypertension (special ethnic groups, individuals with high-normal BP, overweight or obesity). Attention should be paid to postmenopausal women and women with a history of gestational hypertension and preeclampsia, i.e. HDP.

Screening for hypertension

figure7

3.7 Confirming the diagnosis of hypertension

Because of the variability of BP, an elevation of office BP (SBP 140 mmHg or DBP 90 mmHg) should be confirmed by at least two to three visits, unless the BP values recorded during the first visit are markedly elevated (grade 3 hypertension) or CV risk is high, including the presence of HMOD. Although available evidence has some diagnostic limitations and, in clinical practice, collecting ABPM or HBPM data on a large scale may be difficult, out-of-office BP measurements are a source of important clinical information. Therefore, ABPM, HBPM or data from both methods should be collected whenever feasible when office BP is elevated, to confirm the diagnosis of hypertension and identify specific BP phenotypes. ABPM and/or HBPM can be especially important when office BP data from different visits provide variable results. These issues are addressed in detail Section in 4.

4. BP MEASUREMENT AND MONITORING

4.1 Devices for blood pressure measurement

4.1.1 Standard cuff-based devices

The accurate determination of BP is the cornerstone for the diagnosis and management of hypertension [62]. Intra-arterial BP measurement is the only method, which provides true arterial BP values, but its invasive nature prevents any wide clinical use. Furthermore, a noninvasive device that provides finger beat-to-beat BP values close to intra-arterial values has been available for many years, but its use has remained largely confined to research [63–65], except for its helpful diagnostic information in people with autonomic failure or orthostatic hypotension. Thus, for clinical purposes, BP measurement almost entirely relies on indirect noninvasive methods, originally described more than a century ago, that are based on a pneumatic cuff for occluding the brachial artery, the radial artery pulse assessment or a stethoscope for detecting the Korotkoff sounds [66,67]. With some theoretical and practical improvements, these methods are still almost universally used today, and their adoption in clinical trials has laid the basis of diagnostic and treatment recommendations. The different types of noninvasive cuff BP measuring devices available on the market are shown in Table 3.

TABLE 3 - Noninvasive arm cuff BP measuring devicesa
Manual auscultatory devices Automated electronic devices
• Mercury sphygmomanometer
• Aneroid sphygmomanometer
• Hybrid device (LED or LCD display, or digital countdown)
• Automated oscillometric (also wrist cuff devices)
• Automated auscultatory
• Semiautomated (manual inflation)
aCuffless BP devices are currently not recommended for clinical use.

The manual auscultatory BP measurement using a mercury or mercury-free sphygmomanometer is still the gold standard method for validation studies testing the accuracy of novel BP measurement technology [62]. However, due to environmental issues related to mercury toxicity, mercury devices have been banned from clinical use [66]. Hybrid manual auscultatory devices (LED or LCD display, or digital countdown), or good-quality (shock resistant) aneroid devices can be used, although they are subject to observer-related errors, such as terminal digit preference, observer prejudice and bias [62]. The automated cuff-based BP measurement method (mostly oscillometric) has been developed to avoid the observer-related errors of the manual auscultatory method [62]. This method is currently the one used for ABPM and self-home monitoring devices used for HBPM, and is the preferred method for OBPM as well [62]. When BP cannot be measured by an upper arm cuff device, a validated electronic wrist-cuff device may be used [62].

4.1.2 Cuffless blood pressure measuring devices

All cuff BP measurement methods have limitations, mainly because they provide snapshot BP values in static conditions and ignore the dynamic nature of BP, i.e. its variability in response to different daily challenges and activities [68,69]. Moreover, errors due to inadequate cuff size, shape and positioning are common, and the limb compression during cuff inflation may cause discomfort, particularly at work and during sleep [68,70]. Novel cuffless BP measuring devices, which use sensors, signal processing, machine learning and other technologies embedded in wearable devices, smartphones, pocket devices or other types of devices are already available on the market [68,71] and have considerable potential to improve awareness, diagnosis and treatment of hypertension. However, there are several issues that need to be adequately addressed before these devices can be recommended for clinical use [68,71] A major concern is that the AAMI/ESH/ISO Universal Standard protocols [68,72], which are used for validating cuff BP devices, are inadequate for cuffless devices which means that their accuracy remains unproven. Furthermore, there is still no agreed standard of performance for these novel technologies [68]. Finally, most cuffless BP devices require periodic calibration using BP values measured by a standard arm cuff device. Some of them also require information on the user such as age, sex and other characteristics [73]. Basically, these devices do not ‘measure’ BP but track BP changes compared to the calibration BP or attempt to predict BP using demographics and machine learning technology [68,74]. For these reasons, at the present time, cuffless BP devices should not be used for the diagnosis or management of hypertension in clinical practice [68].

4.1.3 Validation of blood pressure measuring devices

In the last 30 years, several scientific societies and associations have developed validation protocols for the evaluation of the accuracy of BP measuring devices [75], the most widely used one being the ESH International Protocol validation [76]. In 2018, a Universal Standard for global use was agreed by the US Association for the Advancement of Medical Instrumentation (AAMI), the ESH and the International Organization for Standardization (ISO) [72]. It should be noted that an electronic BP monitor, which is accurate in adults, may not be accurate in some special populations, e.g. in children or pregnant women, which means that under these circumstances, a separate validation process and dedicated devices for these conditions are required for BP monitoring [72].

Unfortunately, validation for accuracy is not mandatory prior to the distribution of BP devices in the market [75,77]. Thus, established protocols have been used and published in fewer than 10% of BP devices [62,78]. Healthcare professionals, patients and the public should check for lists of accurate devices, which are available on the internet [62]. STRIDE BP (www.stridebp.org) an international organization endorsed by the ESH, the International Society of Hypertension (ISH) and the World Hypertension League (WHL), presents updated lists of validated BP monitors for office, home and ambulatory BP measurement in adults, children and pregnant women in English, Spanish and Chinese. Other national online sources for validated BP monitors are those by the British & Irish Hypertension Society (https://bihsoc.org/bp-monitors/), the German League of Hypertension (www.hochdruckliga.de/betroffene/blutdruckmessgeraete-mit-pruefsiegel), the American Medical Association (www.validatebp.org), the Hypertension Canada (www.hypertension.ca/bpdevices) and the Japanese Society of Hypertension (www.jpnsh.jp/com_ac_wg1.html) [62].

Periodic calibration is necessary, particularly for aneroid devices, whereas for electronic devices, the performance of the BP measurement algorithm is not affected by use, and what is mainly needed is maintenance of the device parts (e.g. cuffs, tubes or connections). Checking for maintenance is advisable once a year for professional office and ambulatory BP devices and less frequently for home devices [72,79].

Devices for BP measurement

figure8

4.2 Standard office blood pressure measurement

Conventional or standard OBPM is the most well studied method for assessing BP and the one by which the diagnosis of hypertension, BP classification, the role of BP as a CV risk factor, the protective effect of antihypertensive treatment and the BP thresholds and targets of therapeutic interventions have been established [4,62,80]. Despite several limitations, and the increasing use of out-of-office BP measurements, OBPM remains the most widely used method for hypertension diagnosis and management [4,62,81].

Despite its widespread long time adoption, use of OBPM is often poorly standardized, leading to inaccurate BP estimations, which often consists of an overestimation of BP and overdiagnosis and overtreatment of hypertension [4,62,81]. A systematic review of 328 articles identified 29 potential sources of inaccuracy related to the patient, the device, the procedure and the observer, which can influence the BP levels and lead to unreliable diagnosis [82]. Thus, it is of utmost importance to use a standardized OBPM methodology that allows uniformity of the setting and the conditions of measurement, the patient position, the device, the measurement schedule and the interpretation of the results [62]. Office BP should be measured on a bare arm. An appropriate cuff size is crucial for accurate BP measurement and must be selected according to the arm circumference of each individual. A single cuff cannot fit the range of arm sizes of all adults. A smaller than required cuff overestimates BP whereas a larger cuff underestimates it [83]. For manual auscultatory devices, a cuff with an inflatable bladder length and width of 75– 100% and 37–50% of the individual middle upper arm circumference, respectively, is required [62,72]. Importantly, for automated electronic devices, the cuff size should be selected according to the device instructions. People with large arms (mid-arm circumference >42 cm) require a conic-shaped cuff because rectangular cuffs may overestimate BP [84]. When BP cannot be measured by an upper arm cuff device, a validated electronic wrist-cuff device may be used [62].

Additional important requirements for the proper use of OBPM are the following. One, diagnosis of hypertension should not be based on a single office visit, unless office BP indicates grade 3 hypertension (180/110 mmHg) or the patient is at high or very high risk based on the presence of HMOD or CVD [62] (Fig. 4). In the vast majority of patients, an accurate evaluation of office BP requires at least two to three office visits at 1– 4-week intervals (depending on the BP level and CV risk) using the average of the last two out of three readings per visit [62,85–89]. Two, in older persons (>65 years of age), treated hypertensive patients (especially very old patients), diabetic patients, patients with neurodegenerative disorders, or with symptoms suggesting postural hypotension, BP should also be measured 1 and 3 min after standing for detecting orthostatic hypotension [62]. At the initial office visit, BP should be measured in both arms, ideally with electronic devices that can measure them simultaneously. An interarm SBP difference >10 mmHg must be confirmed with repeated measurements. If confirmed, the arm with the higher BP should be used for all subsequent measurements because its values more accurately reflect the BP level in the major arteries. Moreover, using BP taken on the arm with the higher reading seems to improve the outcome prediction [90]. A consistent interarm SBP difference >15 to 20 mmHg may be due to atherosclerosis and restriction of large intrathoracic or upper arm arteries, requiring investigation for arterial disease [62,91].

Office BP measurements

figure9

4.3 Unattended office blood pressure measurement

Unattended OBPM performed automatically (three or more readings) without the medical staff being present in the examination room (patient alone) favors a standardized office BP evaluation by ensuring a quiet environment, multiple BP readings and no talking [62,92]. This BP measurement method was used by the SPRINT study, although in a retrospective survey, nonattendance by the healthcare personnel appeared to be variable [93]. Available data agree that absence of medical personnel leads to lower BP values than those associated with standard OBPM, because of a reduction of the alerting response to the medical staff's presence or the white-coat effect [92,94,95]. This can make unattended BP values closer to those obtained by out-of-office BP measurements, although quantitative disagreements with home and daytime mean ambulatory BP have been reported [96]. However, a major problem of unattended OBPM is that evidence of its ability to predict reduction of outcomes by treatment is limited to one trial (SPRINT) [97], which is in contrast with the large and consistent volume of trial-based evidence available for standard OBPM. Furthermore, again in contrast with standard OBPM, little evidence exists on the ability of unattended OBPM to sensitively predict CV events including CV mortality and mortality in the general population [92,98]. Finally, unattended OBPM requires facilities (equipment, space and personnel) that can make it difficult or even unfeasible to accommodate the large number of patients frequently dealt with in general practice and even in dedicated outpatient clinics. Another issue is that the variable difference between unattended and standard OBPM reported in various studies does also not allow to develop a correction factor between the two BP measurement approaches, and thus to meaningfully compare different trials for their threshold and target BP values for treatment.

For these reasons, attended OBPM using an automated device and a standardized protocol (triplicate measurements in appropriate conditions and position) appears to be the most reasonable and practical BP measurement recommendation for clinical practice.

4.4 Blood pressure during exercise

BP increases during dynamic and static exercise, and the increase is more pronounced for SBP than for DBP [99], although only exercise SBP can be measured reliably with noninvasive methods. The increase in SBP during exercise is related to preexercise resting BP, age, arterial stiffness and abdominal obesity, and is somewhat greater in men than in women [100]. There is some evidence that an excessive rise in BP during exercise predicts the development of hypertension, independently from BP at rest [100]. There is currently no consensus on the normal BP elevation during exercise. According to a consensus document of the European Association of Preventive Cardiology, a BP above 220 mmHg in male and 200 mmHg in female measured at peak exercise during cycle ergometry warrants further clinical evaluation including ABPM [101]. Two interesting recent findings are that (i) the BP response to submaximal exercise may have a greater prognostic significance than BP measured at peak [101] and (ii) exercise hypotension may also be a sign of an underlying CV disease [100]. Nevertheless, exercise testing is not recommended as part of the routine evaluation of hypertension because of various limitations, including lack of standardized methodology and definitions. The BP rise accompanying exercise should not discourage patients with treated or untreated hypertension from engaging in regular exercise, especially aerobic exercise, except in the presence of very high BP values (grade 3 hypertension). Regular exercise represents an important lifestyle intervention to chronically lower BP (see Section 7.5).

4.5 Blood pressure measurement in hospital

Because the mercury sphygmomanometer has been banned, mercury-free professional automated BP measurement devices such as digital electronic and hybrid devices (which combine some of the features of both electronic and auscultatory devices) should be used in hospital wards. Multiparametric monitors that measure BP, pulse oximetry, temperature and pulse rate are becoming increasingly popular. Some professional monitors offer a high-speed measurement mode that measures BP in less than 30 s, and others may offer the possibility to determine SBP in a very fast mode [102]. These devices may be especially helpful in emergency units. There are several features that are essential for a BP monitor to be used in hospital [102], two of which are the possibility of being programmed to take multiple BP readings at variable intervals and to have a memory capacity that allows to recall previous measurements. All automatic monitors need regular maintenance and calibration and should be provided with at least two sizes of cuffs for adults, standard and large, and a pediatric cuff. Some devices use a wide-range cuff, which can cover a wide range of arm sizes according to manufacturer instructions [103]. Only devices and cuffs validated by accepted standards should be used [104]. In patients having AF, at least three office BP measurements by auscultation are recommended to account for the varying BP values. Automated oscillatory methods can be also used for BP measurement in AF patients, because they satisfactorily measure SBP and only modestly overestimate DBP. Some devices can apply an AF-specific algorithm [105] that allows to automatically detect AF (see Section 17.3).

Automatic oscillometric measurements may also be considered a reasonably good alternative to intra-arterial measurements in ICUs, the resuscitation area or during surgery [106]. Good agreement between the two methods has been documented within the normotensive BP range in critically ill patients [107], while BP underestimations and overestimations have been observed at very low and very high BP values, respectively [108]. The BP underestimation in hypotensive patients is of particular concern because detection and quantification of hypotension in critically ill patients or in other emergency conditions are crucial for detection and prevention of vital organ hypoperfusion. In a meta-analysis of studies with different oscillometric devices that had brachial intraarterial BP for comparison, the mean SBP underestimation at low BP values was 5.7 mmHg [109], with a wide variability of the between-pressure differences in different patients and for different devices. In patients with hypotensive shock, oscillometric mean BP was found to be 13 mmHg higher than the invasively measured BP [110]. Although oscillometric devices are often used for monitoring BP in emergency medicine and perioperatively [106], direct measurement of BP by an arterial catheter should be used in critical conditions, especially in patients suffering from shock, to guide vasopressor and fluid therapy.

4.6 Central blood pressure

Central (aortic) BP can be assessed noninvasively from peripheral BP waveforms, using tonometry or cuff-based devices and dedicated algorithms [111]. Interest in central BP originates from the consideration that (i) central BP is the pressure to which vital organs and vessels developing atherosclerosis are exposed; (ii) peripheral and central BP values differ and (iii) this is also the case for the effects of treatment [112,113]. A meta-analysis of clinical studies showed that central BP is related to LVH, carotid intima– media thickness and albuminuria, independently of peripheral BP [114]. However, studies and meta-analyses on the predictive value of central BP for CV events have led to conflicting results. According to a recent observational study, higher central pulse pressure was associated with increases in CV outcome incidence even after adjustment for several confounders, including peripheral SBP [115]. An association between central BP and CV outcomes was found also in a meta-analysis, but when studies that included both peripheral and central BP in the same model were considered, a similar risk prediction was observed for peripheral and central BP [116]. Elevated brachial and central BP measurements have also been found to be associated with higher risk of CV events in patients with CKD [117,118]. However, even in this setting, measurement of central BP did not improve the ability to predict CV events or mortality on top of brachial BP measurement. At variance with these results, in a recent individual-level meta-analysis, central hypertension was associated with increased CV and cerebrovascular risk irrespective of the brachial BP status [119]. Thus, the incremental prognostic value of central versus conventional office BP measurement is unclear. In addition to the inconsistency of prognostic data, it should be borne in mind that noninvasive central BP measurement is subject to practical limitations. Devices for central BP measurement have to be calibrated with BP values normally obtained by conventional (usually oscillometric) brachial measurements [111]. Furthermore, although some reference data are available [119,120], there is no conclusive information on (i) the cutoff BP values that differentiate normal from high central BP and (ii) the central– brachial BP correspondence in different population and patient strata. Thus, a widespread use of central BP measurement in the management of hypertensive patients cannot be recommended. The main field of application of central BP may be isolated systolic hypertension in the young (ISHY) in which peripheral BP may be disproportionately elevated compared with a normal central BP value. In this setting, central BP assessment can help to distinguish between a ‘spurious’ benign condition and an ISHY due to early arterial stiffening [121] (see Section 14.5). The clinical use of other types of information derived from central waveform characteristics such as the augmentation index or the wave reflection indices needs to be further defined.

4.7 Home blood pressure monitoring

HBPM provides multiple BP readings away from the office, in the usual environment of each individual [62,122]. It is well accepted for long-term use by patients and has a relatively low cost usually covered by the users [62,122]. HBPM data (i) are more reproducible than those provided by office BP [123,124], (ii) predict HMOD, CV outcomes and mortality better than office BP [125–128], (iii) increase the predictive ability for outcomes, albeit to a modest degree, when added to office BP [129,130], (iv) measure day-to-day BP variability[131], which carries an adverse prognostic significance and (v) identify, like ABPM, hypertension phenotypes such as MH or WCH, which are characterized by CV risk levels different from those associated with sustained or established hypertension [62,122,132–134]. Although data are not entirely univocal [135], a further possible advantage of HBPM is that it may improve adherence to treatment, thereby favoring hypertension control, especially when combined with education, counselling, self-titration algorithms or digital interventions (see Section 21.6) [122,135–137]. Combination with telemonitoring and smartphone applications may offer additional advantages, including the ability to store and transfer home BP data in a digital format and facilitate their evaluation by healthcare professionals [138,139].

HBPM should be performed using automated upper arm cuff devices validated according to an established protocol (www.stridebp.org) [62,122]. Devices with automated storage and averaging of multiple readings, mobile phone, PC or internet link connectivity enabling data transfer may be preferred to facilitate the evaluation of BP values by the physician [62,122]. The measurement conditions and posture should be similar to those described for OBPM [62,122]. Because HBPM is devoid or almost devoid of a white-coat effect in most patients [140,141], its values are lower than office BP values, with a difference that becomes progressively less pronounced as office BP decreases. In the absence of out-of-office BP outcome data from RCTs, the threshold for home hypertension is defined as the value corresponding to an office BP of 140 mmHg SBP or 90 mmHg DBP, i.e. 135 or 85 mmHg, respectively. By analogy, the home BP target is regarded as the value corresponding to the recommended office BP target (see Section 10), which is not well defined but probably just few mmHg lower [142–145]. It is important to mention that these correspondence-based threshold and target home BP values should be interpreted with caution, because office, home and 24 h mean BP values obtained by ABPM have a limited correlation to one another both in untreated and treated patients [146]. Importantly, in single individuals, these differences may widely depart from the above reported mean values.

Home BP values should be collected before planned office visits or whenever a clinically significant change in BP is suspected. Ideally, home BP should be monitored for 7 days and never for less than 3 days, with duplicate measurements (1 min apart) in the morning (before drug intake if treated) and the evening [62,122,147,148]. First-day readings (usually higher and unstable) should be discarded and averages of the remaining values should be considered [62,122,147,148] (Fig. 5). HBPM helps to improve persistence of BP control during long-term treatment [149] (see Section 21).

F5
FIGURE 5:
Recommendations for BP measurements in the office and at home. Adapted [62]. aUse an automated electronic (oscillometric) device, which is validated according to an established protocol (www.stridebp.org). A device that takes triplicate readings automatically is preferred. bThe selection of an appropriate cuff size is crucial for accurate BP measurement and depends on the arm circumference of each individual – a smaller than required cuff overestimates BP and a larger underestimates BP. Using automated electronic devices select cuff size according to the device's instructions. At the initial visit, measure BP in both arms. cMeasure in the morning and the evening for 3–7 days. Use the average of all readings excluding the first day. dStrong data linking OBP with CVD. Used in most observational and interventional outcome trials in hypertension.

Limitations of HBPM are that it requires patient training, is often performed by inaccurate devices, can induce anxiety and lead to overly frequent measurements that may be followed by treatment modifications by the patient [122]. Another limitation is a lack of night-time BP evaluation, which is not a marginal disadvantage, because night-time BP has been shown to predict outcomes more effectively than daytime BP[122]. However, novel HBPM devices allow automated BP measurements during sleep, which provide similar asleep BP values as those provided by ABPM, similar correlations with HMOD [70,150] and have an independent prognostic value [151]. Finally, as mentioned above, a most important limitation (shared with ABPM) is that there are no studies on HBPM-guided treatment and outcomes.

Home BP monitoring (HBPM)

figure11

4.8 Ambulatory blood pressure monitoring

Although patients have to remain immobile at the time of the BP measurements, ABPM provides multiple BP readings in conditions that reflect the usual environment, including daily activities and sleep [62]. ABPM has the same advantages over OBPM than those reported for HBPM, i.e. greater reproducibility of 24 h mean BP values, closer association with and prediction of HMOD, better prediction of outcomes and mortality [128,152–154] and the ability to identify WCH and MH. It has as additional advantages the possibility to discriminate between apparent and true resistant hypertension [155] and to quantify BP characteristics such as 24 h BP variability and the morning BP surge, which have been found to have an adverse prognostic value, independently of 24 h mean BP [134,156,157]. A most important and so far almost unique advantage is the quantification of the dipping status, i.e. the magnitude of nocturnal BP change, which is clinically relevant because night BP reduction and absolute night BP values have been found to predict events better than daytime BP, with a markedly elevated risk in patients with no night-time BP reduction or nocturnal hypertension [158,159]. ABPM may facilitate the identification of daily life hypotensive episodes and the persistency of BP control by treatment during the periods between drug intakes. For research on BP-lowering therapies it has the advantage that its use is accompanied by no or a minimal placebo effect [160]. However, ABPM is not suitable for frequent use, it is rather expensive, it is not widely available in primary care settings, and may cause discomfort to some patients, especially during sleep [62]. As for HBPM, the most important limitations are that no outcome-based RCTs have been conducted to explore the effect of ABPM-guided versus OBPM-guided treatment and no BP thresholds and goals for treatment have been directly established [154]. Because ABPM does not elicit a white-coat effect [140,141], ambulatory BP values are lower than office values. The threshold for ambulatory hypertension is defined as a 24 h mean SBP of 130 mmHg or a DBP of 80 mmHg corresponding to office BP values of 140 or 90 mmHg, respectively. As for HBPM, in the absence of trials on the effect of ambulatory BP reduction on outcomes, the ambulatory BP values to reach with treatment are inferred from their correspondence with the target office BP. This carries the same limitations as those reported in Section 4.7.

ABPM is performed using fully automated devices programmed to record BP automatically at preselected intervals for 24 h in a usual workday [62,161]. Patients are instructed to keep a diary of their activities, symptoms, meals, drug intake times, sleep times or any unusual problem. Average daytime, night-time and 24 h BP values are provided by the device software [62], which also provides the hourly BP profile. Several guidelines require a minimum of 20 valid awake and seven valid asleep BP readings for a valid 24 h recording [62,162]. However, because some readings may be eliminated by the device software, this may lead to hours without BP values, particularly during the night. Furthermore, there is evidence that measuring BP at 60 min intervals (i.e. a total of 24 values during the 24 h) may provide an incorrect 24 h mean BP [163,164]. To limit the risk that poor quality of the collected data invalidates the advantages of ABPM, an appropriate procedure can be to measure BP every 20 min throughout the day and night. This will avoid the paradox of making BP information more scarce during the prognostically more important fraction of the 24 h (night-time) (Table 4).

TABLE 4 - Definitions of hypertension according to the correspondence of home and ambulatory BP values with office BP
Method SBP (mmHg) DBP (mmHg)
Office BPa ≥140 and/or ≥90
Ambulatory BP
 Awake mean ≥135 and/or ≥85
 Asleep mean ≥120 and/or ≥70
 24 h mean ≥130 and/or ≥80
Home BP mean ≥135 and/or ≥85
aRefers to standard office BP measurements (not unattended measurements). Data compare the averages from cohorts of untreated and treated individuals. Given the low correlation between office and out-of-office BP values, individuals can have considerable discrepancies from the averages.

Additional indices derived from ABPM recordings have been considered, and some were found to have an independent prognostic value, including indices of BP variability [165–168], the morning BP surge [156] and the ambulatory arterial stiffness index [169–171]. However, their incremental predictive value is still unclear [165]. Thus, at present, these indices should be regarded as research tools with no indication for routine clinical use.

Ambulatory BP monitoring (ABPM)

figure12

4.9 Clinical indications for HBPM and ABPM

HBPM and ABPM share several clinical indications (Table 5). Both HBPM and ABPM can identify WCH and MH and should be considered when, based on OBPM, the odds of one or the other condition is higher, i.e. in patients with grade 1 hypertension for WCH and with high-normal BP for MH [62]. Both HBPM and ABPM are also diagnostically important in patients in whom (i) a normal office BP is accompanied by the presence of HMOD or high CV risk and (ii) elevated office BP (particularly if in the grade 2 or 3 range and long-lasting) is not accompanied by HMOD. A common indication is also the condition in which a large variability of office BP values does not allow to reach a clear decision on the diagnosis of hypertension. Finally, HBPM and ABPM can both identify different treated patient phenotypes, i.e. when there is no control of both office and out-of-office BP, selective control of only one BP or control of both BP values.

TABLE 5 - Clinical indications for home and ambulatory BP monitoring
Conditions in which white-coat hypertension is more common, e.g.:
 • Grade I hypertension on office BP measurement
 • Marked office BP elevation without HMOD
Conditions in which masked hypertension is more common, e.g.:
 • High-normal office BP
 • Normal office BP in individuals with HMOD or at high total CV risk
In treated individuals:
 • Confirmation of uncontrolled and true resistant hypertension
 • Evaluation of 24 h BP control (especially in high-risk patients)
 • Evaluating symptoms indicating hypotension (especially in older patients)
Suspected postural or postprandial hypotension in treated patients
Exaggerated BP response to exercise
Considerable variability in office BP measurements
Specific indications for ABPM rather than HBPM:
 • Assessment of nocturnal BP and dipping status (e.g. sleep apnea, CKD, diabetes, endocrine hypertension, or autonomic dysfunction)
 • Patients incapable or unwilling to perform reliable HBPM, or anxious with self-measurement
 • Pregnancy
Specific indications for HBPM rather than ABPM:
 • Long-term follow-up of treated individuals to improve adherence with treatment and hypertension control
 • Patients unwilling to perform ABPM, or with considerable discomfort during the recording
Indications for repeat out-of-office BP evaluation (same or alternative method – HBPM/ABPM)
 • Confirmation of white-coat hypertension or masked hypertension in untreated or treated individuals
BP, blood pressure; CKD, chronic kidney disease; CV, cardiovascular; HMOD, Hypertension-mediated organ damage.

There are also selective indications for ABPM or HBPM. ABPM should be performed when postural or postprandial hypotension is suspected, which is more frequently observed in older patients, diabetic patients, patients with dysautonomia, and patients developing anxiety with self-BP measurements or patients who are unable to perform and correctly report the results of the procedure. HBPM may be indicated in patients reporting discomfort or sleep problems with ABPM as well for checking long-term BP control during treatment. On the other hand, and despite technological advancement on HBPM, ABPM remains the best approach to characterize nocturnal BP phenotypes, i.e. the dipping status, which is frequently altered in conditions such as diabetes, OSA, obesity and CKD, with a prognostic reflection that extends to the general population. Nevertheless, it is important to know that different dipping patterns and other BP phenotypes based on office and out-of-office BP measurements have a limited reproducibility [172–174]. Thus, these phenotypes should be identified by more than one out-of-office BP monitoring, e.g. at least two monitorings spaced by several days or weeks [172,173].

In conclusion, despite the technological and clinical advances obtained in the last decades, use of out-of-office BP monitoring still faces some unanswered questions. The most important questions are (i) whether the improved prognostic ability associated with addition of out-of-office to office BP is substantial or modest, in particular when OBPM is appropriately measured [129,130], (ii) whether HBPM-guided or ABPM-guided therapy results in greater reductions in morbidity and mortality than conventional office BP-guided treatment and (iii) which are the out-of-office BP thresholds and goals for treatment. The last two questions will need an answer from outcome-based RCTs comparing out-of-office versus office BP-based treatments [4,62]. Yet, the large amount of additional clinical information provided by out-of-office BP measurements cannot be ignored and, therefore, the present guidelines support collection of out-of-office BP data not only for specific indications but more in general as an important source of useful clinical information whenever collection of these data is feasible in the individual patient and compatible with the healthcare organization and resources. Data collection extends to both HBPM and ABPM, because available evidence suggests that their specific indications and clinical value are not redundant, competitive or interchangeable but complementary. HBPM is only performed in a standardized sitting posture at home, whereas ambulatory BP is measured at different postures (sitting, standing and lying), in different environments (work, home, other) and during routine daytime activities and night-time sleep [4,62,122,161]. Data from a general population suggest that the risk of events increases progressively with the progressive elevation of three available BPs (OBPM, HBPM and ABPM) [159]. In the same population, the risk of CV mortality associated with WCH was lower when normalityvalues extended to both HBPM and ABPM compared with when it only involved one of these two out-of-office BP [175].

4.10 Blood pressure variability

Old studies on ambulatory intra-arterial BP monitoring have shown that BP is highly variable during the day and to a lesser extent during the night [176,177] due to the interplay between central factors, humoral influences, local vasoactive mechanisms and the buffering influences of the baroreflex [69]. This short-term BP variability was found to be quantitatively related to the BP levels, and thus greater in hypertension than in normotension [177], and to have an adverse effect on the genesis of HMOD [178]. These observations were confirmed by studies with noninvasive ambulatory monitoring, which also showed that 24 h or short-term BP variability is adversely related to the risk of CV outcomes, independently of the 24 h mean BP value [157,167,179,180]. However, although several studies have shown that treatment lowers 24 h BP variability, no study has ever addressed whether a treatment-related reduction of 24 h BP variability attenuates CV risk [157,167,179,180]. A number of studies have also focused on other types of BP variability. Conflicting results have been reported on the prognostic value of within-visit BP variations [181], whereas some studies have reported an association between day-to-day BP variability as assessed by HBPM and the risk of CV outcomes [166,182]. However, the largest body of available evidence relates to what is known as visit-to-visit or long-term BP variability. Post hoc analyses of antihypertensive treatment trials have shown that long-term BP variability such as that measurable as BP differences between visits spaced by 6 or 12 months apart, is associated with CV risk in treated hypertensive patients. In post-hoc analyses of three trials, an increase in the number of medical visits in which office BP was reduced to the recommended control value was accompanied by a proportional reduction in the risk of CV outcomes and mortality, independently of the mean office BP reached during the treatment period [183–185]. Furthermore, in trials or treated cohorts of patients with different demographic and clinical characteristics, between-visit office BP variations were found to be associated with the risk of CV and kidney outcomes, also independently of the mean BP values reached during the years of treatment [186–188]. In one study, combined use of on-treatment mean BP and visit-to-visit BP variability identified more accurately the CV risk of treated hypertensive patients than either measure alone [189]. This suggests that in treated patients, protection depends also on time spent under BP control, as more recently confirmed by the relationship between CV events and calculated TTR (time on therapeutic BP range) or BP load (ratio between BP values at BP target and all values during the treatment period) in renal denervated patients and treated diabetic patients, respectively [190,191]. From a practical perspective, this justifies the recommendation to pay attention to consistency of BP control in treated patients, because absence of control at a given visit probably does not represent an innocent BP elevation but a prolonged period with high BP in the preceding months. Evidence from the ELSA trial shows that an inconsistent BP control is common in treated hypertensive patients [192].

5. PATIENT WORK-UP

The work-up required to obtain the information that is necessary in patients with suspected or established hypertension has been accurately described by previous guidelines [4]. Except for some changes or additions to the list of CV risk factors and measures of HMOD, they have not changed substantially in the last years and will thus be reported in the present guidelines only in table format, with the exception of HMOD, for which a more detailed description has been made available (Tables 6–8).

TABLE 6 - Medical and family historya
Personal history
 • Time of the first diagnosis of hypertension, including records of any previous medical screening, hospitalization
 • Stable or rapidly increasing BP
 • Recordings of current and past BP values by self BP measurements
 • Current/past antihypertensive medications including their effectiveness and intolerance
 • Adherence to therapy
 • Previous hypertension in pregnancy/preeclampsia
Risk factors a
 • Family history of hypertension, CVD, stroke or kidney disease
 • Smoking history
 • Dietary history, alcohol consumption
 • Lack of physical exercise/sedentary lifestyle
 • Weight gain or loss in the past
 • History of erectile dysfunction
 • Sleep history, snoring, sleep apnea (information also from partner)
 • Distress or eustress with job or at home (subjective stress level)
 • Long-term cancer survivor
History and symptoms of HMOD, CVD, stroke and kidney disease
 • Brain and eyes: headache, vertigo, syncope, impaired vision, TIA, sensory or motor deficit, stroke, carotid revascularization, cognitive impairment, memory loss, dementia (in older people)
 • Heart: chest pain, shortness of breath, edema, myocardial infarction, coronary revascularization, syncope, history of palpitations, arrhythmias (especially AF), heart failure
 • Kidney: thirst, polyuria, nocturia, hematuria, urinary tract infections
 • Peripheral arteries: cold extremities, intermittent claudication, pain-free walking distance, pain at rest, ulcer or necrosis, peripheral revascularization
 • Patient or family history of CKD (e.g. polycystic kidney disease)
History of possible secondary hypertension
 • Young onset of grade 2 or 3 hypertension (<40 years), or sudden development of hypertension or rapidly worsening BP in older patients
 • History of repetitive renal/urinary tract disease
 • Repetitive episodes of sweating, headache, anxiety or palpitations, suggestive of pheochromocytoma
 • History of spontaneous or diuretic-provoked hypokalemia, episodes of muscle weakness and tetany (hyperaldosteronism)
 • Symptoms suggestive of thyroid disease or hyperparathyroidism
 • History of or current pregnancy, postmenopausal status and oral contraceptive use or hormonal substitution
Drug treatments or use (other than antihypertensive drugs)
 • Recreational drug/substance abuse, concurrent therapies including nonprescription drugs, e.g. glucocorticoids, NSAIDs/COX-2 inhibitors, paracetamol (acetaminophen), immunosuppressive drugs, anticancer drugs, nasal vasoconstrictors
aAdditional factors to be considered are listed in Table 2 (see Section 3.5).

TABLE 7 - Comprehensive physical examination for hypertensiona
Body habitus
• Weight and height measured on a calibrated scale, with calculation of BMI
• Waist circumference
Signs of hypertension-mediated organ damage
• Neurological examination and cognitive status
• Fundoscopic examination for hypertensive retinopathy in emergencies
• Auscultation of heart and carotid arteries
• Palpation of carotid and peripheral arteries
• Ankle–brachial index
Signs of secondary hypertension (Section 6)
• Skin inspection: cafe-au-lait patches of neurofibromatosis (pheochromocytoma)
• Kidney palpation for signs of renal enlargement in polycystic kidney disease
• Auscultation of heart and renal arteries for murmurs or bruits indicative of aortic coarctation, or renovascular hypertension
• Signs of Cushing's disease or acromegaly
• Signs of thyroid disease
aCan be adapted according to the clinical circumstance.

TABLE 8 - Selected standard laboratory tests for work-up of hypertensive patientsa
• Hemoglobin and/or hematocrit
• Fasting blood glucose and HbA1c
• Blood lipids: total cholesterol, LDL cholesterol, HDL cholesterol, triglycerides
• Blood potassium and sodium
• Blood uric acid
• Blood creatinine (and/or cystatin C) for estimating GFR with eGFRa formulas
• Blood calcium
• Urine analysis (first voided urine in the morning), multicomponent dipstick test in all patients, urinary albumin/creatinine ratio, microscopic examination in selected patients
eGFR, estimated glomerular filtration rate; HDL, high-density lipoprotein; LDL, low-density lipoprotein.
aCan be adapted according to the clinical circumstance.

5.1 Personal and medical history

5.2 Physical examination

5.3 Routine clinical chemistry investigations

5.4 Other investigations in hypertension

5.4.1 Electrocardiogram

A 12-lead ECG is part of the routine diagnostic evaluation in all hypertensive patients. LVH represents an important and typical HMOD, which increases markedly with long-standing hypertension and its severity [193]. The Sokolow– Lyon voltage criteria and the Cornell voltage-duration product are important criteria for ECG-based LVH (Table 9). The presence of strain (ST-T depression in lateral anterior leads) is the most serious sign of LVH on ECG [194]. Fulfillment of the Sokolow– Lyon criteria for LVH is associated with older age, male gender and high BP, whereas fulfillment of the Cornell voltage-duration criteria is associated with a younger age, female gender, lower BP and obesity. Specificity is high (about 97%) when using the above ECG criteria together, but sensitivity is low, e.g. in the range of about 40– 50% of LVH cases in persons above 50 years of age and even lower in younger patients. New diagnostic criteria based on machine learning show a better diagnostic ability [195] but for LVH detection, echocardiography remains more sensitive and preferable (if available), especially in hypertensive patients below 50 years of age but also in the context of a more advanced general work-up (see the following). Nevertheless, it should be considered that ECG-based LVH strongly predicts arrhythmias, including sudden cardiac death, AF, myocardial infarction, HF, stroke and a variety of other cardiac and vascular conditions in a fashion that does not completely overlap with the morbidity and mortality prediction offered by echocardiography [196]. It also provides information about heart rate, cardiac rhythm, AV conduction and normality or abnormality of the repolarization phase, which may influence the selection of antihypertensive therapy, e.g. use of BBs or non-DHP-CCBs. Thus, ECG has an added value both for risk prediction and more, in general, for the clinical evaluation of the hypertensive patient.

TABLE 9 - Assessment of hypertension-mediated organ damage (HMOD)a
Basic screening tests for HMOD recommended for all hypertensive patients Aim
12 lead ECG Measure HR and AV conduction, detect cardiac arrhythmias, myocardial ischemia and infarction, screen for LVH
Urine albumin : creatinine ratio (UACR) Detect and classify CKD
Serum creatinine and eGFR Detect and classify CKD
Extended screening for HMOD
 Echocardiography Evaluate structure and function of the ventricles and left atrium, detect valvular disease, aortic root diameter and ascending aortic aneurysm
 cfPWV or baPWV Evaluate aortic/large artery stiffness
 Carotid artery ultrasound Determine carotid intima-media thickness, plaque and stenosis
 Coronary artery calcium scan Determine the presence and extent of coronary calcium to predict CAD events
 Abdominal aorta ultrasound Screen for aortic aneurysm
 Kidney ultrasound Evaluate size and structure of kidney, detect renovascular disease, determine RRI (by spectral doppler ultrasonography)
 Spectral doppler ultrasonography Diagnosis of renovascular disease and determination of RRI
 ABI Screen for LEAD
 Retina microvasculature Detect microvascular changes
 Cognitive function testing (MMSE, MoCA) Screen for early stages of dementia
 Brain imaging (CT, MRI) Detect structural brain damage
aCan be adapted according to the clinical circumstance.

5.4.2 Kidney ultrasound

Due to its low cost and widespread availability, ultrasound imaging of the kidney is commonly used for information on renal morphology (kidney size and structure, roughness, adiposity, kidney stones) and, when contrast-enhanced perfusion is added, microstructure, stiffness, inflammation, edema and abscesses [197]. Kidney ultrasound is a valuable method in the presence of CKD and also a noninvasive examination method for detecting renal artery stenosis because, on a gray-scale ultrasound, it can assess the morphology of both the kidney and the renal arteries. Hemodynamic changes in the renal artery and the kidney can be evaluated with color and spectral Doppler ultrasound. Contrast-enhanced ultrasound may directly show a diameter change in the renal artery with intravascular contrast material that is not harmful even to patients with poor renal function [198].

5.4.3 Selected biomarkers and genetic markers

5.4.3.1 Lipoprotein (a)

Epidemiological and genetic studies involving hundreds of thousands of individuals strongly support a causal and continuous association between Lp(a) concentration and CV outcomes in different ethnicities [199–201]. An elevated Lp(a) is a CV risk factor even at very low levels of low-density lipoprotein cholesterol.

5.4.3.2 Cardiac biomarkers

A large number of studies indicate that both high-sensitivity troponins (hs-cTnT and hs-cTnI) and natriuretic peptides (BNP and NT-proBNP) are able to detect individuals that are at higher CV risk.

Studies suggest that on top of their role in the diagnosis and management of several cardiac conditions including asymptomatic LV dysfunction [202], symptomatic HF [203] and acute coronary syndromes [204][205], these markers may also have a role in the detection of early functional and structural cardiac changes associated with hypertension. Their measurement can be extended to primary care patients without the need for high-technology equipment.

5.4.3.3 Kidney markers

Cystatin C, a 13-kDa cysteine proteinase inhibitor protein, is produced by all nucleated cells at a steady rate and is freely filtered by the kidney with near-complete reabsorption and catabolism in the proximal tubule, and thus no significant urinary excretion [206]. Serum cystatin-C levels are much less affected than creatinine levels by patient characteristics such as gender, age, body size and composition, and nutritional status [207]. Cystatin-C has been found to be a more accurate measure of kidney function than serum creatinine and creatinine-based formulas used to calculate eGFR in different patient populations. Cystatin-C captures earlier, more subtle changes in kidney function, which makes its use suitable to identify patients with early impaired function (when serum creatinine is in the upper normal or mildly elevated range) and to more sensitively measure the risk of progression toward CKD. Currently, however, cystatin C has a low availability, and the cost-effectiveness and clinical relevance of its possible wider use remain to be determined [208,209]. Bioptic detection of tubulointerstitial damage and fibrosis is highly prognostic for subsequent kidney failure. Considerable research is currently being devoted to biomarkers that reflect noninvasively kidney tubular damage and provide information on the risk of CKD progression and associated adverse clinical outcomes beyond the use of eGFR and urinary albumin excretion.

Available kidney tubule biomarkers can be grouped into those that reflect tubule cell injury (kidney injury molecule 1, epidermal growth factor, monocyte chemoattractant protein one) and those that reflect tubule cell dysfunction (a1-microglobulin and uromodulin). These biomarkers provide new opportunities to monitor the response to treatments in CKD patients [210], and, in hypertension, they can be used to distinguish kidney injury from hemodynamic causes of a decline of eGFR. Monitoring these biomarkers serially allows clinicians to monitor persistent beneficial effects of treatment in the presence of benign eGFR declines [211,212].

5.4.3.4 Genetic markers

A positive family history is a frequent feature of hypertensive patients, with hypertension heritability estimates that vary between 35 and 50% in most studies. Rare monogenic forms of hypertension belong to secondary forms of hypertension (see Section 6) [12]. There are also inherited forms of phaeochromocytoma and paraganglioma [213], which require genetic testing for early diagnosis and prognosis as certain mutations are more likely to predict malignancy (see Section 6). Polygenic risk scores, or weighted summations of risk conferred by multiple disease-associated single nucleotide variants, are emerging as tools to improve prediction of common complex CV diseases, including hypertension. For multiple CV diseases, polygenic risk scores are independently associated with respective CVD [214]. Additional potential clinical utility of these scores includes earlier identification for the need of lifestyle interventions, earlier screening for subclinical atherosclerosis, time of initiation of pharmacotherapies and use as a risk-enhancing factor for primary prevention in middle-aged patients at low or moderate conventionally measured 10-year CV risk [215]. However, despite these impressive studies, the role of polygenic risk scores in hypertension, and their possible use in clinical practice should await further clarifications.

5.5 Assessment of hypertension-mediated organ damage

HMOD refers to structural or functional changes in large and small arteries or end-organs (brain, heart, kidney and eyes), caused by elevated BP and is a marker of preclinical or asymptomatic CV or kidney disease [4]. HMOD is common in severe or long-standing hypertension but can also be found in less severe hypertension. With wider use of imaging, HMOD is becoming increasingly detected in asymptomatic patients [216][217]. CV risk increases with the presence of HMOD, and more so when damage increases progressively and affects multiple organs and functions [4,218,219]. Some types of HMOD can be reversed by antihypertensive treatment, especially when treatment starts early, but with long-standing hypertension, HMOD may become irreversible despite BP control [220]. Nevertheless, BP-lowering treatment is important as it may delay further progression of HMOD and oppose the trend toward a progressive increase of CV risk [220]. Although poor technical provision and cost may limit the search for HMOD in some countries, it is recommended that basic screening for HMOD be performed in all hypertensive patients and that a more detailed assessment be implemented when the presence of HMOD is important for treatment decisions. The examinations that can be used to identify HMOD are shown in Table 9.

5.5.1 HMOD in the heart

In hypertension, the heart is directly exposed to an increased load with consequent development of several structural and functional alterations, which are asymptomatic at an early stage but represent a potent risk factor for subsequent CV events, such as HFpEF or HFrEF, AF, CAD, sudden death and also stroke. Preclinical or asymptomatic hypertensive heart disease includes LVH, LV geometric changes, impaired diastolic and systolic function, LA enlargement and greater incidence of arrhythmias. In clinical practice, most or all parameters indicating hypertensive heart disease should be evaluated in a comprehensive examination, using the ECG and available imaging techniques.

5.5.1.1 Left ventricular mass and geometry

As mentioned above, LVH, as detected by two-dimensional transthoracic echocardiography (2D-TE) is a more sensitive marker of LVH than ECG. It is also a major predictor of morbidity and mortality in both hypertensive patients and the general population [221–223]. LVH significantly reclassifies CV risk when added to CV risk factors in most [224,225], although not all [226] studies. In a population-based sample [227], adding LVH to a commonly used risk score (SCORE) significantly reclassified risk in hypertensive patients. Among a variety of echocardiographic measurements, LV mass index was the single, probably most important predictor of adverse events in patients with hypertension [227]. Furthermore, antihypertensive treatment can be accompanied by a regression of LVH, although only for a fraction of the patients exhibiting this HMOD [220]. Regression of LVH during antihypertensive treatment predicts a better prognosis [194,228–230]. LVH can be detected by different methods, all with advantages and disadvantages (Table 10).

TABLE 10 - Advantages and disadvantages of methods to assess LVH in clinical practice
ECG ECHO 3D ECHO CMR
Sensitivity ++ +++ ++++ ++++
Specificity +++ +++ ++++ ++++
Reproducibility ++++ +++ ++++ ++++
Prognostic significance ++++ ++++ + ++++
Availability ++++ +++ ++ ++
Cost + ++ ++ ++++

When LVH is detected, it is recommended to follow patients, preferably by echocardiography for monitoring the evolution of LVH and other structural and functional modifications of the heart, the goal being LVH regression. This may take years, cannot always be achieved, and is less achievable in women, and patients with diabetes, obesity and African American ethnicity [231].

5.5.1.2 Transthoracic echocardiography

2D-TE represents the first and most widely used imaging technique to assess LVH. Image quality is the most important factor for reliable measurements. Use of artificial intelligence has been reported to provide a more accurate measurement of cardiac structural and functional alterations [232]. Echocardiographic LVH at baseline has important independent adverse prognostic significance, and adverse or beneficial prognostic changes have also been observed for LVM increases or reductions during FU or antihypertensive treatment [221,229,230]. 2D-TE also permits the evaluation of LV geometry, LA and aortic root dimensions, LV diastolic dysfunction, end diastolic diameter and LV systolic dysfunction [216]. LV concentric geometry is assessed by an increased RWT, and combining RWT and LVM allows to classify the LV geometric adaptation to hypertension as concentric or eccentric, with or without LVH [233]. A significantly higher risk has been observed in patients with concentric nondilated LVH, and the highest risk is exhibited by patients with LV enlargement [230,234]. Regression of LVH during antihypertensive treatment is associated with a reduced risk of CV outcomes but not necessarily with an improvement of LV diastolic dysfunction [235]. Three-dimensional echocardiography (3D-TE) has a better accuracy and reproducibility than 2D-TE, as it does not rely on geometric formulas. However, further research is needed to more reliably establish normality values, prognostic significance and general feasibility [236,237].

5.5.1.3 Cardiac magnetic resonance

Cardiac magnetic resonance is the gold standard for quantification of cardiac structure and function in clinical studies. An important application of cardiac magnetic resonance is tissue characterization, using late gadolinium enhancement [238] and T1 mapping [239] for extracellular volume. This allows to detect interstitial myocardial fibrosis, which may precede the development of LVH. Further studies, together with a wider availability and substantially lower costs, are needed to increase the clinical use of cardiac MRI. It is also important to mention that the better accuracy and reproducibility of cardiac MRI to detect LV changes provides an important advantage for research in the cardiac HMOD area [240].

Left atrial and aortic root dimensions. LA dilatation is usually a response to increased LV filling pressure and is independently associated with adverse CV events and greater incidence of AF [241,242]. Together with LVH and age, hypertension is also associated with ascending aortic dilatation, which has been found to lead to a greater risk of CV events and may predict the development of aortic regurgitation [243].

Left ventricular diastolic dysfunction. LV diastolic dysfunction is usually the first manifestation of cardiac damage, even before the development of LVH. LV diastolic dysfunction can be detected as an alteration of transmitral inflow pattern (E/A ratio) at Echo-doppler examinations. Initial LV abnormal relaxation may progress to increased LV filling pressure, which may be detected using additional parameters, such as tissue doppler of mitral annulus (e’ velocity) and estimated pulmonary pressures (tricuspid regurgitation retrograde velocity) [244]. Echocardiographic LV ejection fraction may not identify early preclinical LV systolic dysfunction and is also characterized by a large variability at repeated measurements [245]. Speckle tracking echocardiography is a valuable tool to detect early subclinical LV systolic dysfunction, particularly with the measurement of global longitudinal strain, which is characterized by high sensitivity and excellent reproducibility, and has been also standardized in absolute values [246].

5.5.1.4 Computed tomography of the heart

Computed tomography may be used in hypertensive patients to obtain noninvasive coronary angiography, mainly to exclude coronary artery disease, when signs and symptoms are atypical, and the results of other cardiological examinations (i.e. exercise ECG, perfusion scintigraphy, stress echocardiography) are ambiguous.

Coronary artery calcium scanning offers the possibility to detect coronary atherosclerosis in its early stages. Coronary artery calcium is identified by noncontrast cardiac-gated multislice computed tomography of the heart, which is a rapid test using low radiation (<1 mSv) that can be performed at low cost by any modern multidetector computed tomography scanner. Coronary artery calcium is increased in individuals with high BP, predicts the risk of new-onset hypertension, and is also a predictor of CV events in patients with and without hypertension [247,248]. Coronary artery calcium scans have been proposed as a tool to refine quantification of CV risk, particularly CAD events, in patients with hypertension, and to better gauge who may benefit from early initiation of BP medications [249,250]. The effect of antihypertensive treatment on this marker of CV risk is not yet known.

5.5.2 HMOD in the arteries

5.5.2.1 Carotid artery IMT and plaques

Carotid intima– media thickness, combines the thickness of the intimal and medial layer of the carotid artery, is quantified by carotid ultrasound, and can be considered a marker for the early stage of atherosclerosis [251]. It is assumed that the IMT value at the carotid bifurcations primarily reflects atherosclerosis and that the IMT value at the level of the common carotid artery primarily reflects hypertension-related hypertrophy [218]. Carotid IMT predicts CV risk [252,253], and a carotid IMT of more than 0.9 mm is considered abnormal [254], although the upper limit of normality varies with age. However, the relative importance of cIMT evaluation in the overall CV risk quantification is still under debate, because in some studies, addition of cIMT did not improve, or only minimally improved, CV risk stratification [255,256]. The debate includes the prognostic value of treatment-induced IMT changes. In a recent meta-analysis of 119 RTCs involving 100 667 patients (mean age 62 years), the progression of cIMT (mean FU 3.7 years) [257] was reduced by a combination of antihypertensive, lipid-lowering and antidiabetic drugs, as well as by dietary and other interventions. Across all interventions, each 10 mm/year reduction of carotid IMT progression resulted in a significant 9% reduction of the risk of CV outcomes. However, separately analyzed cIMT changes by antihypertensive treatment did not show a significant relationship with CV events, and this was also the case for the separately analyzed cIMT changes with lipid lowering and antidiabetic treatment. The presence of a carotid plaque can be defined by an IMT 1.5 mm, or by a focal increase in thickness of 0.5 mm or 50% of the surrounding carotid IMT value [258]. Carotid plaques have a predictive value for both stroke and myocardial infarction, independent of conventional CV risk factors and risk scores [252,253]. Detection of carotid plaques confer superior prognostic accuracy for future myocardial infarction compared with IMT [259]. It also increases CV risk in patients at all CV risk levels [260]. Carotid imaging is recommended in the presence of carotid bruit, previous TIA or cerebrovascular disease, or as part of the diagnostic work-up of patients with evidence of vascular disease to detect more than mild carotid stenoses (>50% of the vessel lumen).

5.5.2.2 Pulse wave velocity

Increase in large artery stiffness is the most important pathophysiological determinant of age-related increase of SBP and reduction of DBP (and thus of ISH or predominant systolic hypertension) and increase in pulse pressure in the old population [261]. The age-related increase of arterial stiffness is accelerated by uncontrolled hypertension [262]. Recent data suggest that increased arterial stiffness may be involved in the early stages of hypertension, with the stiffening preceding its development [263–265]. Arterial stiffness can be measured in superficial arteries (carotid, brachial, radial arteries) by the slope of the relationship between beat-to-beat BP and arterial diameter changes. However, it is now usually measured by carotid– femoral pulse wave velocity (cfPWV) or brachial– ankle PWV (baPWV). cfPWV is the gold standard for measuring large artery stiffness in Europe [266], and its reference values are available for healthy European populations and patients at increased CV risk [267]. Currently, reference values for baPWV are available for Asian populations [268,269], with European studies emerging [270]. Arterial stiffness increases progressively with the BP increase [271], and is thus variable and greater in the systolic than in the diastolic phase [272]. This is because of the intravascular pressure-related distension, first of the more distensible elastic tissue and then of the less distensible collagen. Changes in the anatomical structure of the vessel wall, usually less elastin and more collagen and fibrotic tissue, are responsible for large artery stiffening with age and diseases. However, stiffness is also modulated by functional factors that may cause fast increases of stiffness in the absence of BP changes, presumably via contraction of smooth muscle tissue in the vessel wall, because distensibility may differ between contracted and relaxed vascular smooth muscle. This may account for the acute stiffening influence of increases in sympathetic activity on large and medium-size arteries [273,274]. In stiffer arteries, the traumatic effect of pulsatile intra-arterial pressure is greater and favors atherogenesis.

A large body of evidence exists that assessment of large artery stiffness, using cfPWV or baPWV, can be clinically useful in hypertensive patients. Evidence from the Framingham and European studies has shown that increase of arterial stiffness is highly prevalent in the hypertensive population [254,275,276]. cfPWV is higher in MH than in normotension, which means that in patients with a normal office BP, increased PWV may identify those in whom out-of-office BP monitoring should be performed to detect this higher CV risk condition. In two meta-analyses, cfPWV or baPWV [277,278] have shown the ability to more accurately classify CV risk compared with conventional risk-based scores, an advantage of particular relevance in young and middle-aged patients in whom the risk falls into the low or moderate level. Adding cfPWV to conventional Framingham CV risk factors resulted in an NRI for CV mortality of up to 27%, while addition of baPWV to a model incorporating the Framingham risk score improved the NRI by 24.7%. Increased cfPWV and baPWV values have been found to predict an increased risk of new-onset hypertension in apparently healthy adolescents [263], young [279] and middle-aged [264] people. Finally, due to its relationship with age, PWV is considered a main element in the assessment of vascular aging, a concept of great current interest and research [280]. All antihypertensive drugs reduce arterial stiffness passively, i.e. by reducing BP and, thus, unloading the stiffer component (collagen) of the arterial wall. The reduction can be maintained over the long-term [281]. A few studies claim that some drugs may reduce stiffness more effectively [282], thus possibly improving arterial structure, but evidence is not conclusive because stiffness must be measured at identical BP levels, which is difficult. There are reports, however, of treatment-induced reductions in stiffness in the absence of BP reductions [283,284]. No RCT so far has proven that reducing arterial stiffness by antihypertensive treatment induces a reduction of CV events. In one recent RCT, i.e. the SPARTE trial, a PWV-based treatment strategy was compared with the classical BP target-based strategy. There was no significant difference in the primary endpoint (CV outcomes) between the two groups for which, however, the study was underpowered. Nevertheless, in the group in which treatment was guided by reduction of PWV, the age-related increase of PWV was significantly lowered [33,285]. Less cfPWV increase has also been reported with a more intense BP reduction in a post hoc SPRINT study [286], in which the effect on PWV was related to less severe outcomes. PWV improvements have also been associated with improved prognosis in patients with resistant hypertension or on dialysis [287].

5.5.2.3 Ankle– brachial index

ABI is the ratio between SBP in ankle and brachial SBP. Measurements are made with the patient supine, using a continuous wave Doppler, a BP sphygmomanometer (preferentially) or an automated oscillometric device [275]. A low ABI (i.e.≤0.9) indicates a LEAD condition. A high ABI (≥1.40) is also abnormal and indicates medial calcification. In addition to their role in diagnosing LEAD, both a low (<1.00) and a high ABI are independently associated with incident CV events. In an individual-participant meta-analysis [288], the 10 year risk of CV mortality associated with a low ABI was greater (4.2-fold in men and 3.5-fold in women), compared with a normal ABI. Adding ABI to the Framingham risk score reclassified the risk category in 19% of men and 36% of women. Measurement of ABI is relatively easy and requires only short training. It should be performed in all patients with symptoms or signs of LEAD.

5.5.3 HMOD in the kidney

Hypertension is the second most important cause of CKD after diabetes and can also be the consequence of a primary kidney disease. Deterioration of kidney function can be detected by routine laboratory testing, using widely available equations for estimation of GFR (eGFR) based on serum creatinine [289]. Serum creatinine alone is an insensitive marker of renal impairment, because a major reduction in kidney function can occur before serum creatinine rises. CKD is classified according to eGFR, calculated by the 2009 CKD-Epidemiology Collaboration formula [290,291], and the presence and amount of albuminuria [289]. The albumin: creatinine ratio (ACR) is measured from a spot urine sample (preferably early morning urine) and is the preferred method to quantify urinary albumin excretion. The diagnosis of HMOD in the kidney is based on the finding of reduced kidney function or detection of albuminuria, but in hypertension-induced kidney disease, albuminuria may not appear until after the reduction in GFR [289].

A lower eGFR and a higher amount of albuminuria, indicating loss of kidney function and kidney damage, are both independent and additive predictors of increased CV risk and progression of kidney disease [291,292]. In a study based on more than three million participants, eGFR and ACR improved risk stratification for myocardial infarction, stroke and CV mortality based on SCORE2 and SCORE2-OP [33] with an NRI improvement of approximately 10%.

On the other hand, whether treatment-induced changes in eGFR and ACR are predictive of kidney and CV outcomes is still under discussion. Short-term effects of eGFR and albuminuria by pharmacological intervention, either an increase or a decrease, may be mixed up with long-term reductions of eGFR and albuminuria. Treatment-associated long-term changes of eGFR have shown a relationship with kidney failure and CV events, whereas treatment effects on proteinuria or albuminuria have been inconsistently related with mortality. In the ONTARGET study, change of albuminuria after 2 years was assessed in more than 20 000 patients at high CV risk, and change of albuminuria was related to total mortality, CV events and kidney outcomes in both diabetic and nondiabetic patients [293]. That is, patients with a decrease of albuminuria by 50% had a lower mortality, while an increase of albuminuria by 100% was associated with adverse CV and kidney outcome as well as a higher mortality. BP reduction by antihypertensive treatment often leads to an acute increase (up to 20– 30%) in serum creatinine, especially if treatment makes use of RAS blockers. This is interpreted as having a hemodynamic basis (BP-dependent reduction of GFR) and, thus, not to usually reflect kidney injury. However, the long-term clinical significance of this early treatment-associated reduction of kidney function is still unclear [294,295]. Serum creatinine, eGFR and ACR should be documented in all hypertensive patients, and if CKD is diagnosed, repeated at least annually [289]. One negative urinary dipstick test does not rule out albuminuria, in contrast to a normal ACR [296]. Spectral Doppler ultrasonography enables to evaluate the renal resistive index (RRI), a noninvasive and reproducible measure of renal and systemic arterial compliance or resistance. In healthy patients, RRI has been shown to vary from 0.58 0.05 (mean SD) to 0.64 0.04, and a value slower than 0.7 has been traditionally taken to indicate normal impedance to renal blood flow, although a considerable heterogeneity has been reported [297]. An elevated RRI is associated with subclinical signs of renal organ damage in untreated patients with hypertension and normal renal function but it has also a prognostic role for CV morbidity, mortality and renal outcomes in essential hypertensive patients, in CKD and in various CV diseases, in addition to albuminuria and eGFR, independently of the traditional risk factors [298,299].

5.5.4 HMOD in the brain

Hypertension is a major risk factor, not only for acute cerebrovascular events such as ischemic stroke, intracranial hemorrhage and TIA, but also for chronic and asymptomatic or oligosymptomatic brain damage, which may ultimately lead to dementia. In particular, long-standing arterial hypertension is known to exert a cumulative effect on cerebrovascular damage, including atherosclerosis, white matter lesions, silent brain infarcts, microinfarcts, microbleeds and brain atrophy [300], especially hippocampal atrophy [301].

Hypertension causes pathological alterations in cerebral microvessels that damage microvascular structure, network architecture and function, and contribute to the genesis of cerebral microbleeds and lacunar infarcts. The latter are small infarcts, 2–20 mm in diameter, in the deep cerebral white matter, basal ganglia or pons that are presumed to result from the occlusion of a single, small perforating artery supplying the subcortical areas of the brain. In addition, white matter lesions, i.e. areas of abnormal myelination in the brain, that are best visualized as hyperintensities on T2-weighted and fluid-attenuated inversion recovery (FLAIR) MRI sequences, develop in hypertension [302,303]. White matter hyperintensities and silent infarcts are associated with an increased risk of stroke and cognitive decline, including dementia [303,304].

In hypertension, aortic stiffening leads to increased propagation of high BP pulsatility to the high flow–low impedance cerebral circulation [305], leading to small artery remodeling and damage. This arteriolar remodeling can be assessed by analyzing retinal arterioles by high-end fundoscopic cameras, but this technology is not widely available. In cross-sectional studies in middle-aged and older adults, the association between stiffening of the aorta, transmission of excessive flow pulsatility into the brain, microvascular structural brain damage and lower scores in various cognitive domains has been shown repeatedly [306,307]. The relationship between transmission of higher pulsatile energy to the brain in individuals of 60 years of age and faster cognitive decline 10 years later has been documented in a longitudinal study [308]. In individuals older than 80 years, PWV has been associated with cognitive decline independently of BP levels [309]. A recent meta-analysis, including 29 cross-sectional and 9 longitudinalstudies, confirmed the negativeassociation between largeartery stiffness (measured by PWV) and cognition, specifically executive function, memory and global cognition. This association seemed to be independent of demographic, clinical and assessment characteristics [310]. Increased BP variability (day-to-day measurements) may also play a role [311], as well as orthostatic hypotension in older people [300]. In clinical studies, long-term elevated SBP and pulse pressure incognitivelyhealthyadults aged 50 years or older was clearlyassociatedwith subsequent cognitive decline and dementia[312]. In hypertensive patients, the presence of lacunes, microbleeds or large white matter hyperintensities on MRI is the second most prevalent HMOD [276]. Low availability and high cost do not permit the widespread use of brain MRI for the evaluation of hypertensive patients, but white matter hyperintensity and silent brain infarcts should be sought in all hypertensive patients with neurological disturbances, cognitive decline or memory loss, if possible [300,304]. For screening in clinical routine, short cognitive screening tests such as the Mini Mental State Examination (MMSE) or the newer Montreal Cognitive Assessment (MoCA) are available and might be considered in hypertensive patients >65 years of age [300]. Their simplicity allows the tests to be performed by primary care physician or referral hypertension specialists. A MMSE score below 24, a MoCA score below 26 or subjective complains of memory loss should lead to referral to a neurologist or a geriatrician [300].

5.5.5 HMOD in the eye

The classification of hypertensive retinopathy is based on fundoscopy, which permits the detection of retinal lesions such as hemorrhages, microaneurysms, hard exudates and cotton wool spots (grade 3), papilledema and/or macula edema (grade 4). These alterations are indicative of severe retinopathy and are specific, reproducible and predictive of all-cause mortality [313]. Retinal damage of grades 1 and 2, such as focal or general arteriolar narrowing and/or arteriovenous nicking, are less specific and reproducible and also have much less predictive value [314]. Hypertension is also a major risk factor for other retinal vascular diseases, including occlusion of retinal veins and arteries and ischemic optic neuropathy [315]. Fundoscopy should be performed only in selected patients, particularly in those with hypertensive emergencies, suspected malignant hypertension or patients with associated diabetes. New techniques to visualize the fundus using smartphone technologies may help to assess hypertensive retinopathy in a larger number of patients [316].

Retinal arterioles may represent a useful indicator of the remodeling of the microcirculation in other vascular beds of hypertensive patients. In recent years, Scanning Laser Doppler and Adaptive Optics have been increasingly used to estimate the wall-to-lumen ratio of retinal arterioles [317]. Retinal wall-to-lumen ratio was found to be directly related to pressure load, other markers of HMOD and small arteries structural alterations measured in different vascular beds with micromyography, which is the gold standard method, albeit invasive, for evaluating microvessels [318]. While the prognostic value of subcutaneous small artery alterations in hypertension has been documented [319], the predictive value for CV events of the retinal wall-to lumen ratio and its change during treatment needs to be demonstrated (Table 11).

TABLE 11 - Criteria to define HMOD
Measurement Parameter Abnormality threshold
ECG
 LVH SV1 + RV5 (Sokolow–Lyon) >35 mm
R wave aVL ≥11 mm
SV3 + RaVL (Cornell voltage) >28 mm (M), >20 mm (W)
 LVH Cornell voltage (+6 mm in W) × QRS duration
(Cornell duration product)
>2440 mm s
ECHO
 LVH LVM/BSA (g/m2) >115 (M), >95 (W)
LVM/height (g/m2.7) >50 (M), >47 (W)
 RWT LV conc. Remodeling ≥0.43
 LV chamber size LVDDiam/height >3.4 (M), >3.3 (W) cm/m
 LV diastolic dysfunction e′ velocity septal <7 cm/s
e′ velocity lateral <10 cm/s
 LV filling pressure E/e′ average ratio >14
LAV/BSA >34 ml/m2
LAV/height2 >18.5 (M) or >16.5 (W) ml/m2
 LV systolic dysfunction GLS <20%
Kidney
 Function eGFR <60 ml/min/1.73 m2
 Albuminuria UACR >30 mg/g
 Renal resistive index RRI >0.7
Large artery stiffness
 Pulse pressure Brachial PP (>60 years) ≥60 mmHg
 Pulse wave velocity baPWV (in people 60–70 years) >18 m/s
cfPWV (in people 50–60 years) >10 m/s
Carotid atherosclerosis
Plaque IMT ≥1.5 mm, or focal increase in thickness ≥0.5 mm, or 50% of surrounding IMT
IMT >0.9 mm
Coronary atherosclerosis
CAC Age-specific and sex-specific reference value
LEAD
ABI <0.9
Eye
KWB score Grade III (hemorrhages, microaneurysms, hard exudates and cotton wool spots) and grade IV (papilledema and/or macula edema)
  Microvascular changes Wall-to-lumen ratio no established reference value
ABI, ankle–brachial index; ACR, albumin : creatinine ratio; baPWV, brachial–ankle pulse wave velocity; BSA, body surface area; CAC, coronary artery calcium; cfPWV, carotid–femoral pulse wave velocity; DDim, diastolic dimension; ECG, electrocardiogram; eGFR, estimated glomerular filtration rate; GLS, global longitudinal strain; IMT, intima–media thickness; KWB, Keith–Wagener–Barker; LAV, left atrial volume; LEAD, lower extremity artery disease; LVH, left ventricular hypertrophy; LVM, left ventricular mass; M, men; RWT, relative wall thickness; w, women.

5.6 Using HMOD to help stratify risk in hypertensive patients

Assessment of HMOD should be performed at the time when thediagnosis of hypertension has been confirmed in orderto fine-tune the CV and kidney risk stratification. The data may influence the decision to initiate or intensify drug treatment. However, assessment of HMOD is also relevant during FU, as it can help physicians to evaluate the efficacy of therapy. A reduction in a previous HMOD may indicate the success of the therapy while, by contrast, the persistence/increment of HMOD may be a clue to review the prescribed treatment, with lack of adherence being a predominant reason. Likewise, the absence of HMOD at the time of initial evaluation should be monitored in the future, as new development of HMOD usually indicates a higher risk. As a consequence, repetition of HMOD assessment should be a main aspect of the FU (see Section 21).

As reported above, HMOD assessment may play a role in stratifying the risk of patients with hypertension. In this regard, LVH [216,221], baPWV [277] and cfPWV [320], carotid IMT [253,321], carotid plaque [252], CAC [322] and ABI [322] have been shown to predict CV risk on top of traditional CV risk factors. A higher number of measures of HMOD is associated with higher CV risk [252]. In multivariable-adjusted models, the presence of HMOD was associated with a two-fold to three-fold increase in the risk of CVD compared with the referent group in the Framingham study in any BP category above the optimal [276]. Moreover, LVH [224,225], baPWV [277] and cfPWV [320], carotid IMT plus plaque [253], CAC [322] and ABI [288] have been able to significantly reclassify CV risk, when added to traditional CV risk factors/risk scores.

In post hoc analyses, BP treatment-induced regression of some (but not all) manifestations of HMOD has been associated with a reduction in CV risk, thereby providing additional information on the effectiveness of treatment in individual patients. This has been best illustrated for the treatment-induced regression of LVH measured by either ECG or echocardiography [229,323]. A reduced incidence of CV events and slower progression of kidney disease has been reported with the treatment-induced reduction in urinary protein excretion in both diabetic and nondiabetic patients [324,325], but results are discordant [326–329]. There is also evidence that treatment-induced changes in eGFR predict CV events and progression to ESKD [330,331]. A very large recent meta-analysis, including >100 000 participants [257], showed a reduction of CV risk with reduced cIMT progression, in contrast with the conclusion of an older and smaller meta-analyses [332]. Improvement in PWV over a few years is associated with improved prognosis in patients with resistant hypertension [287] or under dialysis [333]. The predictive power of changes of ABI over time has been mainly investigated in CKD patients [334].

Guidance on use and repetition of HMOD with time needs to consider several factors, such as the ability of the HMOD marker to be modified by treatment, the reproducibility of the changes, the time necessary to detect them, their prognostic value and the costs. The characteristics of the most frequent HMOD are shown in Fig. 6.

F6
FIGURE 6:
Characteristics of the most frequent markers of HMOD in hypertension.aUsing modern adaptive optics technology.

5.7 When to refer a patient to a specialist or for hospital-based care

Although most patients with hypertension are managed in the primary care setting, in some circumstances, a hospital-based evaluation and treatment may be required, keeping in mind that out-of-office or office-based care of hypertensive patients also depends on the healthcare organization of a given country (Table 12).

TABLE 12 - When to refer a hypertensive patient to a specialist or to hospital
• Patients in whom secondary hypertension is suspected
• Young patients (<40 years) with grade 2 or 3 hypertension in whom secondary hypertension should be excluded
• Patients with sudden onset or aggravation of hypertension when BP was previously normal
• Patients with treatment-resistant hypertension
• Need of more detailed assessment of HMOD, which might influence treatment decision
• Requirement of more in-depth specialist evaluation from the referring doctor
• Hypertensive emergencies (inpatient care will usually be needed)

6. SECONDARY HYPERTENSION

Secondary forms of hypertension account for only a small fraction of the overall hypertension prevalence, which is largely due to primary hypertension. However, their true prevalence is not precisely known, because available data may be confounded by the selection bias of the studies reported in the literature, the number of undiagnosed cases and the varying definition of secondary forms of hypertension. Hence, the classification of OSA, a phenotype more frequently observed in obese patients, as a secondary form of hypertension is questioned by many experts. Nevertheless, despite their limited prevalence, detection and management of secondary forms of hypertension is of utmost importance, because these forms often carry a high or very-high risk of morbidity and mortality and can possibly be cured by timely treatment of their cause [335]. Secondary forms of hypertension require specific diagnostic approaches, which allow to detect their specific causes and to select effective drug treatment or appropriate interventional treatment that control or cure the elevated BP. Secondary forms of hypertension are a frequent cause of severe or true resistant hypertension, worsening of previously controlled hypertension or increased severity of HMOD, which may appear as disproportionate to the duration of hypertension (Table 13). Although secondary forms of hypertension are particularly frequent in younger patients (<40 years) with an elevated BP, some forms (such as atherosclerotic renovascular disease) are more common at an older age (Fig. 7).

TABLE 13 - Patient characteristics that should raise the suspicion of secondary hypertension
Younger patients (<40 years) with grade 2 or 3 hypertension or hypertension of any grade in childhood
Sudden onset of hypertension in individuals with previously documented normotension
Acute worsening of BP control in patients with previously well controlled by treatment
True resistant hypertension hypertension
Hypertensive emergency
Severe (grade 3) or malignant hypertension
Severe and/or extensive HMOD, particularly if disproportionate for the duration and severity of the BP elevation
Clinical or biochemical features suggestive of endocrine causes of hypertension
Clinical features suggestive of atherosclerotic renovascular disease or fibromuscular dysplasia
Clinical features suggestive of obstructive sleep apnea
Severe hypertension in pregnancy (>160/110 mmHg) or acute worsening of BP control in pregnant women with preexisting hypertension

F7
FIGURE 7:
Incidence of selected forms of secondary hypertension according to age.

Screening all hypertensive patients for secondary hypertension is not feasible or cost-effective. However, secondary forms of hypertension frequently show clinical findings that suggest their presence and even their specific nature. More common causes of secondary hypertension are primary aldosteronism, renal parenchymal disease and renovascular disease, while Cushing syndrome, phaeochromocytoma and paraganglioma, or coarctation of the aorta are less frequently observed. The different types of secondary hypertension show different age distributions (Fig. 7).

Diagnostic suspicion (Table 13) should prompt immediate referral to specialized hypertension centers where the appropriate diagnostic tests and subsequent treatments can be performed [213,336–341]. It is beyond the scope of these Guidelines to describe in detail the clinical management of specific causes of secondary hypertension. For most of them, the reader may refer to corresponding ESH position papers [213,338–341]. A summary of important features of selected forms of secondary hypertension is provided in Figs. 8a–e.

F8A
FIGURE 8:
A Atherosclerotic renovascular disease (ARVD). (a) The prevalence of ARVD differs considerably between studied populations – in a population-based cohort >65 years of age, ARVD (defined as >60% stenosis) was identified in 6.8%. Among hypertensives, the prevalence of ARVD is probably around 1% in patients with mild hypertension, but may be as high as 14%–24% in patients with severe or resistant hypertension. (b) In view of the frequent association with atherosclerotic lesions in other arterial beds, a cardiovascular work-up should be considered. (c) Medical management of ARVD should aim to reduction of CV risk and protection of kidney function; hypertension control is a prominent goal. With regard to antihypertensive treatment, an ACEi or an ARB are considered as first-line option (contraindicated in bilateral renal artery stenosis or stenosis in a solitary kidney). (d) Observational data showed that renal artery stenting in addition to medical therapy is associated with renal and CV benefits in patients presenting with high-risk ARVD phenotypes, e.g. resistant hypertension, recurrent pulmonary edema, heart failure and deterioration of kidney function.
F8B
FIGURE 8:
B Fibromuscular Dysplasia (FMD). (a) FMD occurs predominantly in young or middle-aged women. However it may be diagnosed at any age, both in women and men. Renal FMD is the second cause of renovascular hypertension after atherosclerotic renal artery stenosis. (b) Two subtypes of FMD have been described: multifocal FMD (80–90% of cases) and focal FMD (10–20% of cases). The characteristic lesion of multifocal FMD is the “string of beads”, characterized by alternating areas of stenosis and dilatation in the mid and distal portions of the artery. Focal FMD is characterized by focal stenosis of variable length, which may occur in any part of the artery and requires exclusion of atherosclerosis, inflammatory or genetic arteriopathies. (c) In a meta-analysis, the rate of cure of hypertension after angioplasty was 36% (range 14–85%) but may be much higher in younger patients with recent onset hypertension. Angioplasty deserves also to be considered in patients with renal FMD and resistant hypertension. (d) Stent kinking and fracture have been reported in the setting of renal FMD. Accordingly, stenting is usually not recommended in renal FMD and reserved for treatment of flow-limiting per-procedural dissection or in case of renal artery aneurysm. (e) In over 50% of cases, patients with renal FMD have lesions in one or more other arterial beds (multivessel FMD). Patients with FMD also often have arterial dissections, aneurysms or marked arterial tortuosity. For these reasons, it is recommended to perform at least once a life-time head to pelvis angio-CT or if contraindicated MR-angiography in all patients with FMD.
F8C
FIGURE 8:
C Primary aldosteronism (PA). (a) Depends on the population screened – ranges from 3.2% to 12.7% in primary practice and from 1% to 30% in referral centers; prevalence increases with the severity of hypertension to 20%. (b) PA prevalence in patients with adrenal incidentaloma ranges from 1.6% to 4.3%. (c) ARR requires at least normalization of plasma potassium and interruption of existing treatment with spironolactone and BBs. (d) Overall, seated SIT appears reliable and less complicated than FST and SLT. CCT may be a good alternative in patients at risk of potential fluid overload (patients with kidney failure or HF). (e) Although the majority of PA cases are sporadic, up to 5% of patients may have a familial form of the disease. Genetic testing should be performed in all patients with early onset PA (i.e. <20 years of age), irrespective of the severity of the clinical phenotype, and in patients with a family history of PA. (f) Steroidal MRAs are the treatment of choice for PA in patients with bilateral adrenal disease or unilateral disease that cannot be surgically treated.
F8D
FIGURE 8:
D Pheochromocytoma and paraganglioma (PPGL). (a) Varying from 0.2% to 0.6% in hypertensive patients to less than 0.05% in the general population. (b) These symptoms are most commonly reported whereas other (pallor, tremor, nausea, panic/anxiety) occur at a much lower frequency. (c) Routine genetic testing is recommended in all PPGL. The most frequently involved genes include SDHB, SDHD, VHL, RET and NF1. PPGL in patients with a known SDHB mutation do carry higher malignant potential and so require close FU with regular imaging and biochemical screening. (d) Presurgical medical preparation using an alpha-1 receptor blocker (doxazosin or phenoxybenzamine) as first choice is always required for preventing life-threatening perioperative cardiovascular complications. (e) All patients operated for a PPGL should be followed up annually for at least 10 years. The first FU should be at 2–6 weeks after surgery to verify completeness of surgical resection.
F8E
FIGURE 8:
E Cushing's syndrome. (a) In specific populations including patients with difficult to control hypertension or type 2 diabetes. The incidence in the general population is 0.7-2.4 per million per year. (b) In case of abnormal initial result of a first screening test, positivity of at least one of the remaining screening tests is required to establish the diagnosis.

6.1 Genetic causes of secondary hypertension

Several rare monogenic forms of hypertension have been described, in which mutations in specific genes, mostly coding for proteins involved in sodium tubular reabsorption or steroid metabolism account for the pathogenesis of hypertension (Table 14). An exception is represented by familial autosomal dominant hypertension with brachydactyly, in which the pathogenetic mechanism resides inside the vascular smooth muscle cells [342]. Hypertension is usually already present in childhood or early adulthood, though age of onset and severity of hypertension may be modulated by lifestyle, environmental factors and genetic susceptibility. Specific drug treatments addressing the corresponding molecular defects are indicated (for example, amiloride in Liddle's syndrome or dexamethasone in glucocorticoid remediable aldosteronism) [12]. Routine genetic testing by approved genetic laboratories should be done in all phaeochromocytomas and paragangliomas (PPGL) with a yield of genetic mutation at approximately 40%. PPGL patients with a known succinate dehydrogenase subunit B mutation carry a higher malignant potential and should be more closely followed with regular imaging analysis and biochemical screening (Fig. 8).

TABLE 14 - Rare genetic causes of secondary hypertension [343]
Condition Phenotype Mechanism and Treatment
Liddle syndrome Hypokalemia, metabolic alkalosis, low PRA or PRC, low PAC Increased renal tubular ENaC activity; responds to treatment with amiloride
Apparent mineralocorticoid excess Hypokalemia, metabolic alkalosis, low PRA or PRC, low PAC Decreased 11β-hydroxysteroid dehydrogenase isoenzyme 2; responds to spironolactone
Gordon syndrome Hyperkaliemia, metabolic acidosis, low PRA or PRC, low/normal PAC Overactivity of the sodium-chloride cotransporter; responds to thiazides
Geller syndrome Pregnancy-exacerbated hypertension, low PRA or PRC, low PAC Agonist effect of progesterone on the mineralocorticoid receptor (which is constitutively active); responds to amiloride, spironolactone activates instead of blocking the receptor
Glucocorticoid-remediable aldosteronism (familial hyperaldosteronism type 1) Hypokalemia, metabolic alkalosis, low PRA or PRC, increased PAC Chimeric CYP11B1/CYP11B2 gene; responds to glucocorticoids
Familial hyperaldosteronism type 2 Hypokalemia, metabolic alkalosis, low PRA or PRC, increased PAC Increased activity of CLCN2 chloride channel; responds to steroidal MRA
Familial hyperaldosteronism type 3 Hypokalemia, metabolic alkalosis, low PRA or PRC, increased PAC Loss of selectivity of KCNJ5 potassium channel; patients who do not respond to steroidal MRA require bilateral adrenalectomy
Familial hyperaldosteronism type 4 Hypokalemia, metabolic alkalosis, low PRA or PRC, increased PAC Increased activity of CACNA1H calcium channel; responds to steroidal MRA
PASNA syndrome (primary aldosteronism, seizures and neurological abnormalities) Hypokalemia, metabolic alkalosis, low PRA or PRC, increased PAC; neurological defects coexists Increased activity of CACNA1D calcium channel; responds to steroidal MRA and CCB
11beta-hydroxylase deficiency Hypokalemia, metabolic alkalosis, low PRA or PRC, low PAC, virilization of female individuals Reduced activity of 11β-hydroxylase with increase of DOC and androgens; responds to glucocorticoids
17alpha-hydroxylase deficiency Hypokalemia, metabolic alkalosis, low PRA or PRC, low PAC, pseudohermaphroditism in male individuals Reduced activity of 17α-hydroxylase with increase of DOC and reduction of androgens; responds to glucocorticoids
Autosomal dominant hypertension with brachydactyly [342] Brachydactyly type E (BDE), short stature, severe hypertension (salt-independent, age-dependent), high risk of death from stroke before age 50 PDE3A mutations upregulated the cAMP-hydrolytic activity that results in lower cAMP levels in vascular smooth muscle cells

7. LIFESTYLE INTERVENTIONS

7.1 Relevance of lifestyle changes

The adoption of a heart-healthy lifestyle is a fundamentally important approach to prevent or delay the onset of hypertension, reduce the elevated BP values and lower the associated increase of CV risk [4,344]. Individuals with a healthy lifestyle score have an approximately 4– 5 mmHg lower BP, irrespective of the underlying BP genetic risk, than those with an unfavorable lifestyle [345]. Further, healthy lifestyle measures can augment the BP-lowering effect of pharmacological interventions and reduce the number of drugs needed to control BP [346–349]. Each of the lifestyle interventions has greater efficacy at higher starting levels of BP. However, lifestyle changes should never delay the initiation of drug therapy in patients in whom the protective effect of antihypertensive drugs is documented and the related benefits require BP reductions that cannot be obtained by lifestyle-dependent changes only. Although the evidence is largely limited to observational studies and their meta-analyses, all lifestyle interventions seem to have heart-healthy benefits that may go beyond the benefits associated with their effect on BP. The most important and well established effective lifestyle interventions that have been shown to reduce premature CV morbidity and mortality are losing weight [349], the DASH diet [350,351], salt reduction [352], augmentation of potassium intake [353–355], engaging in regular physical activity and structured exercise [356–359] and a moderation of alcohol consumption [360–362]. In addition, smoking cessation and other lifestyle measures are also important beyond BP. Various other nonpharmacological interventions (e.g. dietary components like polyphenols, coffee and tea, or stress-reducing therapies) have been reported to lower BP, but the extent and/or quality of the supporting clinical trial experience is less robust and persuasive.

The Achilles’ heel of treatment strategies based on or inclusive of nonpharmacological interventions is the low persistence of the prescribed measures. This is in part the result of the difficulty of permanently adhering to lifestyles that may interfere with working or home habits and needs. Some lifestyle measures also have a cost, which may not be reimbursed by healthcare providers. After prescribing to their hypertensive patients lifestyle changes that can help or achieve BP control, physicians should establish a FU program that allows to check whether there is adherence to the prescribed measures and whether the therapeutic goal is achieved. This will minimize the risk of the patient continuing with an uncontrolled BP for a prolonged time.

7.2 Weight reduction

Being overweight or obese has been directly associated with hypertension [363,364], whereas weight-loss interventions are well established strategies to lower BP [349,365,366]. A network meta-analysis found reductions of 6.5 mmHg for SBP and 4.6 mmHg for DBP following a low-caloric diet in adults with prehypertension [366]. In patients with hypertension, a low-caloric diet was ranked first among all lifestyle interventions in lowering SBP and DBP [366]. Similarly, a meta-analysis of RCTs further concluded that, for each kilogram of body weight loss, both SBP and DBP were reduced by approximately 1 mmHg [349]. Additionally, attenuation of pressogenic factors such as sympathetic activation [367] and a 15% lower all-cause mortality has been found following weight loss interventions, irrespective of age [368]. Modest weight loss, is therefore, a key recommendation and should ideally be achieved through a combination of a low-caloric diet and exercise [366,369]. It should be acknowledged that achievement and maintenance of weight loss through behavioral changes are often challenging, although feasible, over prolonged periods of FU [370]. A rather frequent phenomenon is weight cycling (sequential losses and regains of body weight), which may adversely affect BP, CV risk and the metabolic profile [371]. For those who do not meet their weight loss goals with nonpharmacological interventions, pharmacotherapy could be considered, although evidence on the effectiveness of weight-loss medicines on BP is scant, based on a small number of studies, and these drugs are often associated with unwanted side effects [372]. The GLP-1 RA reduce body weight and concomitantly lower BP by few mmHg, a favorable therapeutic effect in patients with diabetes and obesity (see Section 20.1.3). Alternatively, bariatric surgery is an effective, longer lasting strategy for morbidly obese patients to manage BP and CV risk factors and might be considered in case of failure of all of the above measures [373], particularly in patients with severe obesity. Amongst specific predictors associated with weight loss interventions, greater initial weight loss and higher adherence to lifestyle advice were accompanied by greater weight loss success [370]. The type of weight loss programme should always be individually tailored, taking into account setting of realistic goals, tailor-made dietary and exercise regimes and frequent FU to motivate and address challenges in behavior change [370].

7.3 Restriction of sodium intake

There is strong evidence for an association between high sodium consumption and increased BP in the overall population [374,375] and hypertensive patients [376]. Furthermore, the relation between sodium-restricted diets and improved BP control has been widely recognized by randomized trials and confirmed by meta-analyses [377,378]. Greater BP reductions have been observed in hypertensive patients and other patient categories (nonwhite people, older populations, patients with diabetes, metabolic syndrome or CKD) [378,379], and restriction of sodium intake has also been reported to lower BP in patients with resistant hypertension [380] and to reduce the number of drugs necessary to achieve hypertension control [381]. A recent network meta-analysis provided evidence for lifestyle interventions restricting sodium intake to <100 mmol (5.8 g salt per day) resulting in an average of approximately 5/2 mmHg SBP/DBP reduction in patients with hypertension [382].

Sodium and corresponding salt (NaCl) values are approximately:

Sodium (Na+) 2.0 g = 87 mmol Na+ = NaCl (salt) 5.0 g.

Moreover, meta-analysis of RCTs examining sodium intake reduction to as low as 800 mg/day showed a linear decrease in BP [352,378], In addition, a weighted average reduction in dietary sodium intake from about 3.6 g/day to about 2.7 g/ day has been found to be associated with an approximately 18– 26% reduction in CV disease [383]. However, whether the best therapeutic strategy should be to pursue unlimited sodium restriction is still a matter of debate [384,385], because there are observational studies showing that below a sodium intake of approximately 3.5 g/day further BP reduction is associated with an increased mortality in both hypertensive patients and the general population [376,386,387]. Furthermore, although no side effects have been reported by epidemiological studies in very low-salt diet populations [353], alterations of BP control mechanisms with low-sodium diets have been observed in experimental settings and studies on hypertensive patients [388,389]. Although some intervention studies are available, a lack of proper long-term randomized trials on the effects of various degrees of sodium restriction on outcomes represents the most important limitation for this medical area. In the studies in which the relationship between dietary sodium and CV outcomes exhibited a J-shaped curve, sodium intake was assessed by sodium excretion in spot urine, and this has been criticized as a measure unable to reflect the more accurate 24 h amount of urinary sodium excretion, from which sodium consumption can be more precisely inferred [390,391]. Larger, longer and more precisely controlled intervention studies than those currently available are needed to shed more light on this issue. Sodium is mainly consumed as salt, which in the diet comes from processed foods or is added to the food during cooking or at the table. For people with a long-established habit of high salt intake, it might be difficult to attain and maintain long-term voluntary salt control, and alternative approaches might be needed. A salt substitute with low-sodium content and an acceptable salty flavor would be an ideal alternative. Evidence supports the use of salt substitutes in adults with prehypertension and hypertension [366][392]. In a random-effects model, participants consuming a salt substitute showed significant SBP and DBP reductions (-4.8 and -1.5 mmHg, respectively) compared with participants consuming normal salt [393]. Of the five studies with mortality outcome data, salt substitute also significantly reduced all-cause mortality (hazard ratio 0.88) [393]. Daily diet modification by this nonpharmacological management may, thus, improve BP control.

7.4 Augmentation of dietary potassium intake

Dietary potassium is associated with BP and hypertension [386], with recent data suggesting a U-shape relation, indicating that an adequate intake of potassium is desirable to achieve a lower BP level but that an excessive potassium intake should be avoided [355]. Potassium supplementations (especially with intakes of 75– 125 mmol per day) have been effective in lowering BP [353,354,394], especially in adults with hypertension [395], adults consuming an excess of sodium and Black people. The typical BP-lowering effect of a 60 mmol (1380 mg) administration of potassium chloride has been about 2 and 4– 5 mmHg in adults with normotension and hypertension, respectively, although the BP response can be up to twice as much in persons consuming a high-sodium diet [395]. The large recent randomized controlled Salt Substitution and Stroke Study (SSaSS) reported that increasing potassium intake as a sodium substitute, i.e. replacing 25% sodium chloride with potassium chloride in salt, reduces the risk of stroke, disease and death in patients with increased CV risk plus low-potassium and high-sodium intake at baseline [392]. A recent meta-analysis [355] provides further support for a population goal of potassium intake recently set by international authorities, such as 90 mmol per day (3500 mg/day). In most trials, potassium supplementation was achieved by administration of potassium chloride pills, but the BP response pattern was similar when dietary modifications were used [394]. Because potassium-rich diets tend to be heart-healthy, they are preferred over the use of pills for potassium supplementation. Good sources of dietary potassium include fruits and vegetables, as well as low-fat dairy products, selected fish and meats, nuts and soy products. Four to five servings of fruits and vegetables will usually provide 1500 to >3000 mg of potassium. This can be achieved by a diet, such as the DASH diet, that is high in potassium content.

7.5 Increase levels of daily physical activity and regular exercise

The acute pressor effect of dynamic and isometric exercise does not contraindicate regular exercise on a chronic basis. Large epidemiological studies, which allowed for age and other confounding factors, have yielded consistent evidence of an inverse relation between the incidence of hypertension and habitual levels of physical activity, assessed by means of questionnaires or sometimes an interview [396,397]. For each 10 metabolic equivalent of task (MET)-hours per week increment in leisure-time physical activity (which corresponds to the recommended minimum physical activity level of 150 min per week), the risk of developing hypertension was found to fall by 6% [396]. In addition, the BP-lowering effect of structured exercise has been repeatedly demonstrated in RCTs [346,356,359,398], especially when focused on dynamic aerobic exercise [356,357,399,400] but also, though to a lesser extent, following dynamic resistance training [358,401] and static isometric exercise [402–404]. The BP reductions have been documented irrespective of age [399,405], sex [406,407] or ethnicity [399,408]. The average SBP reductions with aerobic exercise are approximately 2– 4 and 5– 8 mmHg in adult patients with normotension and hypertension, respectively [409]. Regarding the exercise intensity, moderate-intensity aerobic exercise (40– 60% heart rate reserve) is recommended to prevent and treat hypertension [357], although many hypertensive patients present with diverse comorbidities, are older or limited in the level of physical activity they can undertake. In this context, it is of note that a Cochrane meta-analysis including 73 trials found moderate-certainty evidence that walking already results in meaningful BP reductions [410]. In patients with hypertension, a daily bout of exercise is preferred, to minimize the problem of postexercise hypotension [411–413]. Finally, a physically active lifestyle and regular exercise have positive effects on many other adverse health outcomes and CV risk factors at all ages and sexes, and across all BP categories [398,401]. BP reductions and cardiometabolic benefits have also been reported with low-intensity physical activity (6 min hourly) in highly sedentary people [414]. In addition to its role in the prevention of hypertension and reduction of an elevated BP, there is also evidence from observational studies that a physically active lifestyle prevents the development of CVD, irrespective of BP level [415–419]. The reduction of risk is continuous across the full range of physical activity volumes, but the slope of the risk decline is steepest for the least active individuals [415,418].

7.6 Moderation of alcohol intake

Large-scale observational studies report a strong positive linear association between alcohol consumption and BP [420,421]. Data from epidemiological studies on alcohol consumption largely rely on self-reported alcohol intake of participating people as defined by drinks per day. Sex differences in the metabolism of alcohol with a lower first-pass metabolism in female individuals and differences in distribution due to body composition possibly explain differences in the recommended upper limits for daily pure alcohol intake with higher limits for men than for women [4]. This contrasts with the fact that the global attributable impact of alcohol intake to mortality is more than four-fold higher in men than in women [422]. Previous observational data suggested a decrease in CVD, particularly CAD, with light drinking compared with abstainers [423,424]. However, this potential cardioprotective effect of low-to-moderate alcohol intake on CAD seems largely because of a healthier life style in these individuals, and the effect is attenuated after full adjustment for the confounding factors [425]. Indeed, recent epidemiological and genetic Mendelian randomization studies indicated a continuous nonlinear positive relationship between alcohol intake and BP [425,426]. The risk for hypertension increases in both men and women, if daily alcohol intake is at least one to two drinks (at least 10– 20 g alcohol) per day [427]. An important meta-analysis of 36 RCTs with overall 2865 participants (a smaller fraction of 14%, i.e. 401, were women) revealed that alcohol reduction close to abstinence was associated with a 3.3/2.0 mmHg SBP/DBP reduction [360]. The benefit seems to be consistent across trials but confined to people consuming 3 drinks/day (equivalent to about 42 g alcohol intake/day according to the definition in this report). A dose-dependent effect was observed particularly in heavy drinkers, i.e. in people who consume 6 drinks/day at baseline and reduce their alcohol intake by about 50% experiencing an SBP/DBP reduction of approximately 5.5/4.0 mmHg [360]. Additionally, both trial data and observational literature support the hypertensiogenic effect of binge drinking [428]. In this regard, it is important to mention that together with hypertension, excessive alcohol intake is the most important risk factor for intracranial hemorrhage [429]. Consequently, excessive (binge) drinking should be avoided, and patients with high risk for intracranial bleedings should be advised accordingly. Unfortunately, recommendations among different guidelines vary regarding the upper limits and the definition of drinks, and the recommendations of sex-specific upper limits for alcohol intake appear questionable. Nevertheless, moderation of alcohol intake and implementation of alcohol-free days during the week in people who consume drinks that contain alcohol are generally recommended to improve BP control and overall health [430].

7.7 Smoking cessation

Tobacco smoking is the single largest preventable cause of death and is known to significantly increase the risk of CVD [431,432]. Compared with nonsmokers, smokers more frequently present with MH, documented by normal office and higher daytime ambulatory BP values [433]. Because smoking a cigarette is accompanied by a sympathetic nervous system activation and a prolonged BP increase (about 30 min) [434], the ups and downs of BP also increase daytime BP variability [433]. In addition, smoking may impair the BP-lowering effect of some antihypertensive drugs, i.e. BBs [435]. Therefore, the history of tobacco use should be carefully established, and smokers should be encouraged and counselled regarding smoking cessation. Attention should also be given to passive exposure to smoking, which has been associated with the risk of CVD and a 24 h BP elevation [436]. Brief advice from a physician may already be advantageous when time is limited [437], though combining behavioral support with pharmacotherapy increases the chance of success compared with brief advice alone [438]. Increasing the intensity of this behavioral support, as measured through the number of contacts, duration of each contact, and programme length, had point estimates associated with modestly increased chances of quitting [438]. In recent decades, water pipe smoking has developed into a major and rapidly growing alternative to traditional tobacco smoking within the global tobacco epidemic [439]. Likewise, e-cigarettes [440] originally marketed as potential aids in smoking cessation, have attracted a lot of consumers, including both smokers and nonsmokers. Recent meta-analyses now highlight that these ‘so-called safer’ alternatives acutely increase BP [441,442], heart rate and may also be associated with increased risk of CVD [440,441,443]. Similarly, the few available studies showed no clear difference in the CVD incidence between waterpipe smoking and traditional tobacco smokers [444]. Although the available indirect evidence regarding the CV effect of e-cigarette and water pipe smoking is currently based mainly on nonrandomized observational studies of small sample sizes, overall moderate quality and short-term FU, the evidence to date suggests that they should not be regarded as CV safe products [442]. Health professionals should, therefore, be cautious in recommending the use of e-cigarettes to their patients and the general public.

7.8 Other dietary interventions

Diet is an important modifier of vascular health and BP, and it has been shown that targeting the whole diet has synergistic and cumulative effects on BP beyond individual foods and nutrients. The most well established dietary interventions for the reduction in BP are the DASH diet [366] and the Mediterranean diet [445–447], with the DASH eating plan offering the best demonstration of BP-lowering effectiveness [347,366]. The DASH diet promotes the consumption of whole grains, fruits, vegetables and low-fat dairy products. It provides a means to enhance intake of potassium, calcium, magnesium and fiber [351]. High-quality evidence confirms that the DASH diet results in a significant reduction in SBP and DBP, irrespective of the hypertension status [347,366]. Even modest adherence to the DASH diet is associated with a lower risk of all-cause and cause-specific mortality [448]. Greater adherence to a Mediterranean diet has also been found to be associated with a 10% reduction in CV incidence or mortality [449]. A higher adherence to the DASH diet also strengthens the risk-reducing association [448]. Other diets including vegetarian, paleolithic, low-carbohydrate, low glycemic index, high-protein and low-fat diets have also been shown to reduce BP, though results are inconsistent and the quality of evidence low [450,451]. Coffee has been reported to have a modest short-lasting pressor effect but recent data appear to indicate that its moderate regular consumption does not adversely affect BP and the CV system [452], including the absence of an effect of acute coffee consumption on premature atrial contractions [453]. Results from four observational and one quasi-experimental studies have shown that, depending on individual's CYP1A2 genetic profile, a high caffeine intake may actually protect nonsmokers but not smokers from hypertension [454].

7.9 Improve stress management

Stress and anxiety are associated with an increased risk of hypertension and CV events [455,456]. Patients with mental distress may develop a sudden increase in BP, which may normalize when the distress is relieved [456]. Growing evidence also links the exposure to intensely traumatic life events with an increased risk of hypertension [456–459]. Recent meta-analyses report promising results for the ability of mind-body stress reducing interventions to not only reduce stress and mood swings but also SBP and DBP, although the quality of the evidence is low [366,460]. Meditation [461,462] and breathing control through e.g. yoga are considered to be among the better stress-reduction interventions for lowering BP [366], though their effect sizes are smaller compared with the main lifestyle interventions [366].

7.10 Exposure to noise and air pollution

Exposure to environmental noise and air pollution are two major risk factors that exert a negative impact on CV health, particularly in urbanized settings. Both factors are environmental stressors that have been identified as risk factors for increases in BP, incident hypertension and also HMOD, including vascular stiffness [463]. Air pollution is a complex mixture of gaseous and particulate matter components, and noise exposure is largely due to traffic noise. Interestingly, clinical and experimental studies suggest that the two factors may share common mechanistic pathways, leading eventually to vascular inflammation and endothelial dysfunction that mediate the BP increasing effects. Additional studies showed that that cessation of air pollution or noise reduce both BP and the intermediate pathways, supporting a causal link [464]. Thus, reducing traffic noise and air pollution are important general health policy measures in the global and national context and can make an important contribution to improving BP control and CV health. On an individual level, the possibility to escape from detrimental environmental exposures is obviously limited. However, hypertensive patients can reduce exposure to air pollution by modifying the location, timing and type of outdoor activities and may also try to reduce indoor exposure to noise and air pollution.

Lifestyle interventions

figure20

8. BENEFITS OF ANTIHYPERTENSIVE TREATMENT

Although improvement or correction of inappropriate lifestyle patterns can lower BP and reduce total CV risk [465], most patients with hypertension require antihypertensive drug treatment alongside lifestyle interventions. A large number of outcome-based RCTs support preventing CVD by drug treatment [466–468]. Meta-analyses of RCTs have shown that in hypertensive patients, a 7 mmHg average reduction in office SBP substantially reduced major CV events, i.e. stroke, coronary and HF, as well as CV and all-cause mortality. RCTs have also shown a protective effect against asymptomatic cardiac [466,467,469] and kidney damage [468], and accumulating evidence supports a BP-dependent prevention of cognitive decline and dementia [470,471]. These protective effects have been observed irrespective of baseline BP within the hypertensive range, the level of CV risk, the presence of comorbidities (e.g. diabetes, dyslipidemia and CKD), age, sex and ethnicity. Furthermore, the clinical benefits shown by more recent meta-analysis [472] are similar to those provided by meta-analysis of the older RCTs published in 1994 [473], indicating that the benefits of antihypertensive treatment have not been attenuated by the widespread concomitant prescription that are common in contemporary medicine of lipid lowering, antidiabetic and antiplatelet protective therapies to higher risk patients. It should also be mentioned, that the antihypertensive drug-related benefits are likely to be even greater than those described by RCTs, because the analyses of trials are usually done according to the intention-to-treat principle, which means that they include patients irrespective of whether they adhere to their treatment. It is well known that adherence to the assigned drug treatment strategy is poor in clinical practice but far from being optimal also in trials [474–477], which reduces protection because adherence is closely associated with the benefits of antihypertensive drug treatment [478,479]. Finally, several comprehensive cost-effectiveness analyses on the use of pharmacological treatment of hypertension have been reported [480–483], and there is a general agreement that treatment of hypertension is highly cost-effective, because prevention of a large number of fatal and nonfatal events (the latter leading to hospitalization, complex medical interventions and frequent disabilities) is accompanied by a marked reduction of healthcare-related costs. This can be effectively reduced by treatments, which are largely based on inexpensive drug classes that are almost always also available as generics. The recommendations that follow are based on outcome evidence from RCTs. However, whenever appropriate, mention is also made of other types of data because in hypertension, outcome-based RCTs have some important limitations, such as that data are largely restricted to middle-aged, older and higher risk patients and treatment duration covers a relatively short period, usually 3– 5 years. This means that important recommendations such as treatment of hypertension in young patients and continuation of BP-lowering interventions indefinitively require an extrapolation, the appropriateness of which can be supported by other sources of evidence. Big data, now being collected by national health system registries, health insurance companies, health utilization databases and prolonged observational FU of RCTs, are becoming an important source of long-term information in wider population strata. Evidence from these sources suggests that the benefits of antihypertensive treatment reported by RCTs for a limited number of years are maintained for many years beyond the trial duration.

9. ANTIHYPERTENSIVE DRUG TREATMENT INITIATION

9.1 Should treatment initiation be based on total CV risk?

A recent meta-analysis of RCTs reported that BP reduction was beneficial across almost the entire BP range [484], whereas when BP-lowering data were stratified according to CV risk, relative risk reductions were similar for the various risk strata [485,486], while absolute risk reductions were greater with higher baseline CV risks. These data have been taken as evidence that BP-lowering treatment should be initiated according to the CV risk level and that target patients should be those at the greatest risk, irrespective of their BP. The present guidelines do not support this conclusion, because evidence is available that, although compared with the low-risk condition, patients at high or very-high CV risk exhibit a greater treatment-induced absolute reduction of CV outcomes, higher CV risk levels are associated with a disproportionately greater residual risk compared with patients at low– moderate risk. This means that at high-risk, treatment fails to exert an adequate protection, presumably because a considerable proportion of the high risk is not reversible by treatment anymore [80]. Moreover, failure to recommend antihypertensive treatment in the low-risk condition, such as in younger patients, does not take into account that under this circumstance, the benefit cannot be quantified only by reduction of CV events or mortality, but it includes the delay or prevention of asymptomatic or subclinical organ damage and, thus, progression to a high and largely irreversible risk years later.

The above data and considerations support earlier in-life treatment of hypertension as well as treatment implementation also when CV risk is still low-to-moderate. Although total CV risk provides clinically important information and should always be quantified, the data also support the decision to start antihypertensive drug treatment based on office BP level according to the values mentioned in the subsequent sections.

9.2 Office BP thresholds for initiation of drug treatment

All guidelines agree that patients with grade 2 or 3 hypertension should be offered antihypertensive drug treatment alongside lifestyle interventions [4,32,487,488]. Guidelines are also consistent in recommending that patients with grade 1 hypertension and a high CV risk should be treated both via lifestyle modifications and via BP-lowering drugs. Offering BP-lowering drugs to patients with grade 1 hypertension and low-to-moderate CV risk (no CVD, diabetes, CKD or HMOD) has been denied in the past and is still somewhat controversial. The uncertainty originates from the fact that most RCTs have mainly recruited patients with at least grade 2 hypertension, often within an age range in which age “per se” importantly contributes to a high CV-risk level [489]. Furthermore, some RCT-based meta-analyses have not found a significant treatment-related CV benefit in patients with grade 1 hypertension and low CV risk [486]. However, three more recent meta-analyses [472,490,491], have reported significant treatment-induced reductions of CV events and mortality in patients with grade 1 hypertension. The conclusion of two of these meta-analyses [472,491] is weakened by the fact that a substantial number of patients was on antihypertensive treatment, thus having a higher initial BP and thus presumably not always a grade 1 hypertension. Furthermore, several patients had diabetes and were, therefore, at high CV risk. This was not the case for the third meta-analysis [490], in which patients with grade 1 hypertension and a true low-risk condition showed a 31% combined stroke and CAD reduction with an about 7 mmHg SBP reduction (4 trials, 8073 patients). These findings have been further supported by the results of the HOPE-3 trial [492], showing a 27% reduction in major CV outcomes when SBP was lowered by 6 mmHg in a subgroup of patients mostly (80%) untreated, at intermediate CV risk and with baseline SBP values >143.5 mmHg (average 154 mmHg). Based on the above data, we recommend that in all hypertensive patients, lifestyle advice should be accompanied by BP-lowering drug treatment, and that this should include patients with grade 1 hypertension, irrespective of the CV risk (Fig. 9). However, in patients with grade 1 hypertension in its lower BP range, no HMOD and a low CV risk, the possibility may be considered to start treatment with lifestyle changes only (see Section 7). The duration of a lifestyle-based intervention alone can be of several months (e.g. 3 to 6 months) and will depend on the level of BP within the grade 1 range (closer to 140 mmHg SBP), the opportunities for the implementation of lifestyle changes and the perceived adherence to the lifestyle regimen, all related to the likelihood of achieving BP control. If BP control is not achieved within a few months of a lifestyle-based approach, drug treatment will be necessary.

F9
FIGURE 9:
Diagnosis by office BP and initial management of hypertension.

Office BP thresholds for drug treatment initiation

figure22

9.3 Should BP-lowering treatment be initiated in patients with office BP <140/90 mmHg?

Previous guidelines [4] recommended avoiding antihypertensive treatment in people with high-normal BP and low CV risk. This decision was based on the following findings: (i) in the RCTs and meta-analyses that reported a reduction of CV outcomes by lowering an initial high-normal BP, all or many patients were already under antihypertensive treatment and had thus an original BP higher than that measured in the trials [466]. This has been the case, for example, in the SPRINT trial, in which patients had a baseline SBP slightly below 140 mmHg on a background of almost two antihypertensive drugs per patient [97], (ii) the HOPE-3 trial [492] showed that BP-lowering treatment did not reduce the risk of CV events in people with low-moderate CV risk and SBP values in the high-normal range and (iii) a meta-analysis of 13 RCTs or RCT subgroups of patients at low-moderate CV risk and an untreated baseline BP in the high-normal or normal range showed (n =21 128 patients) no effect of BP-lowering treatment on any CV outcomes [493]. The present guidelines reconfirm the recommendation not to initiate antihypertensive drug treatment in low-to-moderate risk patients with a BP in the high-normal range. In these patients, intervention should be limited to lifestyle advice, because this reduces their risk of progression to established hypertension and increased CV risk. It is important to note that the recommendation not to offer drug treatment to people with a high-normal and low-moderate CV risk does not take into account the results of a large, recent, individual-participant meta-analysis of RCTs, which has reported that in both primary and secondary CV prevention, an SBP reduction of 5 mmHg reduced outcome risk when baseline SBP spanned almost the entire normality range, including SBP values <120 mmHg [484]. However, interpretation of the data from this meta-analysis is problematic for a variety of reasons [36]. For example, many patients included in this meta-analysis were previously treated with BP-lowering drugs, which means that their true baseline BP was variably higher than recorded. Furthermore, data exhibited subgroup inconsistencies, i.e. benefits, were seen in patients with a baseline SBP <120 mmHg but not in some subgroups of patients with SBP values above 120, 130 or even 140 mmHg. Finally, in this meta-analysis, the benefits of antihypertensive treatment at normal baseline BP were at least in part inferred from comparisons of patients in whom BP differences were obtained by increasing treatment in the lower on-treatment BP group and discontinuing treatment in the higher on-treatment BP Group. Treatment discontinuation might have caused a rebound increase in outcome and magnified or created the outcome difference with the group at lower BP values [494,495].

The decision regarding treatment may be different in patients with a high-normal BP and a very high CV risk. In a meta-analysis of 10 RCTs or RCT subgroups that included individuals with established CVD (mainly myocardial infarction) and an untreated high-normal BP (26 863 patients), a few mmHg SBP reduction was accompanied by a reduced risk of stroke, although not of any other CV event [493]. In another RCT meta-analysis of trials in patients with previous CAD and a baseline mean SBP of 138 mmHg, treatment was associated with a 10% reduction in the risk of major CV events, although not with prolonged survival [472]. Thus, treating people with high-normal BP and established CVD, especially CAD, can be recommended because this has a protective effect, albeit limited to some BP-dependent outcomes and restricted to patients at very high CV risk. It should be considered, however, that the vast majority of these patients will probably already be under BP-lowering drugs, administered in the context of GDMT (e.g. RAS inhibitors or BBs in patients with CAD) for their direct CV protective properties.

9.4 Drug treatment initiation in older people

The present guidelines rely on the definition of older patients as those aged ≥ 60 or 65 years [489], although further definitions of older persons as those aged ≥ 80 years are now also used (see Section 15.3.2). Evidence that at ≥ 60 or 65 years of age, antihypertensive drug treatment is beneficial, is unequivocal and greater than that available for younger patients. The BP threshold for drug-based interventions in older patients has for years been an SBP ≥ 160 mmHg, because this was the BP recruitment criterion for all RCTs in older hypertensive people [5]. Information on the treatment benefits with a lower SBP was of little help, because data included a large number of patients who were already treated at trial initiation. Recently, however, evidence in older patients mostly or entirely untreated at trial initiation has been made available. Thomopoulos et al.[489] have shown that, in patients from three major trials aged 60– 79 years, BP-lowering treatment was accompanied by a clear reduction of CV death, major CV events and all-cause death [489]. This has been confirmed in two more recent trials. In the HOPE-3 trial [492], the beneficial effects of BP-lowering on CV outcomes were observed in patients with grade 1 mostly untreated hypertension, whose mean age was about 66 years. In the STEP trial [496], conducted in patients 60– 80 years old, more intense BP-lowering reduced the risk of major CV events, and the CV benefits were obtained primarily in people with a baseline SBP >152 mmHg and a Framingham Risk Score ≥ 15%. Thus, the present guidelines recommend antihypertensive drug treatment to be implemented in patients aged ≥ 60 or 65 years, when SBP is ≥ 140 mmHg, regardless of the DBP level. In older patients aged 60 to 79 years treatment is recommended also when a SBP elevation is accompanied by a normal or even low DBP, because outcome-based RCTs have shown that in ISH patients within this age range, antihypertensive treatment is beneficial also if DBP is <90 or 80 mmHg [145,497,498]. This has been recently confirmed by a meta-analysis of five RCTs in 15 636 ISH patients aged 70– 84 years (mean 71.5 years) in whom a SBP reduction from 171.3 to 145.2 mmHg was accompanied by a 30% reduction of major CV outcomes compared with placebo patients [144]. Because in ISH-dedicated trials, entry SBP was 160 mmHg, candidates to antihypertensive treatment are definitively patients with grade 2 and 3 ISH [497,498]. However, given the considerable number of patients with ISH or a prevalent SBP elevation in most RCTs on patients aged 60 to 79 years with a SBP ≥ 140 mmHg [499], treatment of patients with grade 1 ISH may be considered. This notion is supported by a recent meta-analysis of 23 RCTs in which antihypertensive treatment was beneficial in patients aged on average 67 years in whom mean baseline SBP and DBP values were ≥140 mmHg and <90 mmHg, respectively [500,501]. Antihypertensive treatment in ISH is further addressed in Sections 14.6 and 15.3.

Evidence on the BP threshold for treatment is much more scant for patients aged ≥ 80 years in whom the beneficial effect of SBP reduction has been documented in only one specifically designed outcome trial, HYVET [502]. In this trial, recruitment was based on a SBP ≥ 160 mmHg with or without a DBP elevation, which means that at this more advanced age, antihypertensive treatment can be recommended in grade 2 or 3 hypertension. This recommendation applies also to patients aged 80 years or more with ISH, because in this advanced age range, a selective or prevalent SBP elevation involves most patients, as also shown by the HYVET trial. However, although direct information from RCTs is not available, a lower SBP threshold for drug treatment, i.e. ≥150 mmHg SBP may also be considered based on extrapolation from the HYVET data that the treatment-related benefits were seen at SBP values <150/80 mmHg. Furthermore, although BP values were probably underestimated, in a substudy of SPRINT limited to patients aged 75– 84 years (average 80 years) treatment-related benefits were seen with SBP reductions from initial values that were well below 160 mmHg [503].

A final important recommendation is that in patients under well tolerated treatment, who reach 80 years of age, treatment should be continued, because there is evidence that in hypertensive octogenarians, discontinuation of chronically used BP-lowering drugs is associated with a rebound increase of outcomes. In the HYVET trial, about two-thirds of the overall 3845 patients were on antihypertensive drugs at randomization to either active treatment or placebo [495]. Accordingly, a substantial number of patients stopped treatment when randomized to the placebo group, and in these patients, the greater risk of outcomes compared with treated patients was particularly consistent and marked. An exception to the recommendation to continue antihypertensive treatment in octogenarians is represented by very old patients with low SBP values (120 mmHg or less) or with severe orthostatic hypotension, especially in the presence of polypharmacy and a high frailty level. In these cases, progressive reduction of drug treatment should be considered, but deprescribing should be implemented cautiously because data on the effectiveness of this procedure are still missing.

The present guidelines strongly support the concept that age should be no barrier to antihypertensive drug treatment, as further emphasized in the section specifically devoted to hypertension in old people. However, they also recognize that evidence on antihypertensive treatment in the old patients has limitations that go beyond the need for more evidence on the most appropriate BP threshold for treatment. One limitation is that in HYVET, the patient age was close to 80 years (mean 83 years) [502] and that, thus, no RCT-based evidence is available on whether treatment is beneficial in hypertensive patients close to or above 90 years of age, an expanding category in Europe. Furthermore, it is widely recognized that evidence is severely limited in physically and mentally frail older patients, and absent in institutionalized patients, because these patients were usually excluded from RCTs on hypertension. Thus, the above recommendations mainly relate to relatively fit and independent older patients. Data from several observational studies show an inverse relationship between SBP and morbidity/mortality rates in very old frail patients, especially in those under antihypertensive treatment [504–511]. However, in one of these studies, better adherence to antihypertensive drug treatment was associated with a reduced mortality, including in a subgroup of patients in whom a wide number of comorbidities, and a history of multiple previous hospitalizations increased the risk of mortality to up to about 70% in 6 years. This was the case also for use of a statin in people older than 85 years, in whom mortality exceeded 80%. With the limitations of their observational nature, these studies suggest that antihypertensive as well as other CV drug treatments may also be protective in very frail old patients [512,513]. However, RCTs specifically devoted to frail older patients are necessary to obtain solid evidence on antihypertensive-dependent protective effect as well as on the appropriate BP threshold (and target) values for treatment.

10. OFFICE BP TARGETS FOR TREATMENT

10.1 Office BP targets in the general hypertensive population

Based on RCTs, as well as their meta-analyses and post hoc analyses, the 2018 ESC/ESH guidelines recommended that the first objective of antihypertensive treatment should be to lower BP to <140/90 mmHg in all patients. If drug treatment is well tolerated, treated BP values should be targeted to 130/80 mmHg or lower, with the caveat that in some clinical conditions (e.g. CKD), the evidence for this lower BP target was uncertain. It was additionally recommended that treatment should never target BP values to <120/70 mmHg because of the lack of consistent evidence that this has an incremental protective effect, the risk of introducing harm and the increased risk of side effects, leading to treatment discontinuation and a rebound increase in events. In a large population study from northern Italy, discontinuation of antihypertensive drugs for several months was associated with an almost 40% increase of hospitalization for CAD, stroke and HF compared with patients continuing treatment [514]. With the exception of the STEP trial [496], no further trial on target BP has been made available since the 2018 guidelines, but the issue has been further explored by two Cochrane meta-analyses of RCTs and by a large individual patient-based meta-analysis [484]. The first Cochrane meta-analysis [515] assessed whether targeting BP to <135/85 mmHg is associated with a reduction in mortality and morbidity as compared with standard BP targets, i.e. <140/90 mmHg, in a rather general hypertensive population. The analysis included 11 RCTs involving 38 688 patients with a mean FU of 3.7 years. Patients randomized to the lower target achieved a mean SBP/DBP of 122.8/82.0 mmHg versus 135.0/85.0 mmHg in the standard target group. The authors concluded that the benefits of trying to achieve the lower BP target did not outweigh the harms associated with that intervention, because the number of patients needed to treat to the lower target in order to benefit one patient ranged between 167 and 250 while the corresponding number needed to harm was 37 [515]. However, the results of this meta-analysis are open to criticisms because the lower BP target was somewhat arbitrary (no major trial has ever compared targets <140/90 and <135/85 mmHg), and its selection had the purpose of enriching the otherwise small number of trials included in the meta-analysis. Furthermore, the achieved SBP in patients randomized to the lower target was <123 mmHg, which makes the conclusion more against a <120 mmHg than against a <135 mmHg SBP target. Finally, SPRINT [97] was one of the trials included, which means that in a nonmarginal proportion of patients, BP data were obtained by the unattended BP measurement technique, presumably leading to lower BP values than those obtained in all other trials [516]. A similar absence of clinical benefit in the lower SBP target has been reported by the second Cochrane meta-analysis, which was performed in hypertensive patients with a history of CVD [517]. However, the number of trials and patients was smaller than in the first meta-analysis; several major trials on secondary prevention were not included and among the included trials, the risk of performance bias was high and the quality of the studies was graded as low. On the other hand, the results were different in a third meta-analysis, which reported that a 5 mmHg SBP reduction is associated with reduced outcomes in patients with and without previous CVD, whose baseline BP ranged from >170 to <120, which means that protection was found even at a target SBP value <115 mmHg [484]. In addition to the problems of this meta-analysis discussed in Section 9.3, patients with a baseline SBP <120 mmHg represented less than 2% of the overall database and included the SPRINT data, in which BP values were probably underestimated [36].

Thus, information on SBP and DBP targets for drug treatment still substantially rely on the two large meta-analyses of RCTs referred to in the 2018 ESC/ESH guidelines [466,518]. Both meta-analyses concluded that, compared with on-treatment SBP values ≥ 140 mmHg, fatal CV outcomes, nonfatal CV outcomes and all-cause mortality are reduced at on-treatment SBP values within 130– 139 mmHg but that a further incremental benefit can be seen when SBP is reduced to the 120– 129 mmHg range. Stratification of RCTs for achieved DBP also showed an incremental reduction in all types of CV outcomes and mortality for values <80 mmHg compared with 80–89 mmHg and 90 mmHg. This was the case also in patients aged at least 65 years old [489]. The benefits from intensive BP-lowering referred to patients at all levels of risk, including those with and without existing CVD, diabetes and CKD. Two important findings of one of the two meta-analyses were also that, in absolute values, the incremental outcome benefits of BP-lowering progressively decreased as the target BP was lowered. Furthermore, permanent treatment discontinuation (because of treatment-related adverse effects) steeply increased in patients targeted to progressively lower BP values, a finding consistent with the much greater incidence of kidney and other major side effects reported at lower BP targets in several major trials. For these reasons the present guidelines recommend a SBP <140 mmHg as a target in most hypertensive patients. At variance from the 2018 ESC/ESH guidelines [4], a DBP target <80 mmHg is recommended, also because DBP is usually <80 mmHg when SBP is <140 mmHg, with the exception of isolated diastolic hypertension (see Section 14.7).

However, despite the smaller incremental benefit, an effort should be made to reach a SBP range of 120– 129 mmHg, but only if treatment is well tolerated to avoid the risk of treatment discontinuation because of adverse events, which might offset, in part or completely, the incremental reduction in CV outcomes. Attention should also be given to the fact that evidence on the advantages of a lower BP target are not available or unequivocal in a number of clinically important subgroups of patients. These issues will be discussed in the sections on special conditions.

10.2 Office versus home and ambulatory BP targets

A crucial gap in hypertension research is that no outcome-based RCT has ever used HBPM or ABPM to guide treatment of hypertension. One attempt to address this issue was made more than 20 years ago by a small RCT in which the number of CV events registered during a several-year FU did not differ significantly between treatments based on ABPM versus office BP, although their overall number (n= 25) did not make the data conclusive [519]. Another almost simultaneous attempt did not focus on CV events but showed that over 1 year of antihypertensive treatment, 24 h BP reduction was associated with a more effective regression of LVH than office BP reduction, home BP reduction having an intermediate favorable effect [520]. Thus, although HBPM and ABPM have been shown to be prognostically more sensitive and additive to the prognostic value of office BP [128,521] no information exists on (i) whether guiding antihypertensive treatment by HBPM or ABPM has a greater protective effect than guiding by office BP and (ii) which are the optimal HBPM and ABPM targets for treatment. As mentioned in Sections 4.7 and 4.8, currently these targets are indirectly inferred from 24 h mean BP corresponding to the target office BP of <130/80 mmHg, which unfortunately has important limitations. In this context, however, an accepted notion is that the difference between office and HBPM or 24 h mean BP decreases as office BP decreases. This means that the difference between office and out-of-office BP is considerably less pronounced during than in absence of treatment, especially when treatment provides office BP values close to target. At a certain value, office and out-of-office BPs coincide, but this value varies between studies, although for office and 24 h SBP (for which more data are available) is probably around 120 mmHg [522].

10.3 Timing of BP control and time in therapeutic range

RCTs have usually defined the achieved BP target as the mean BP value during the treatment period. However, it is now clear that this oversimplifies the relationship between BP reduction and outcomes. Post hoc or secondary analyses of several RCTs such as VALUE have shown that in high-risk patients achieving an earlier BP control (within 6 months [523] and or even within 1 month [524]), the risk of outcomes was lower than in those in whom BP control was achieved later. Furthermore, post hoc analyses of large RCTs in high-risk patients or large observational studies have consistently documented that the greater the percentage of visits in which BP is controlled, the lower the risk of outcomes, independently on the mean BP during the overall treatment period [183,185,525]. Similar observations have recently been made in resistant hypertensive patients treated with RDN [191] and in type 2 diabetic patients in whom the measurement was, respectively, the number of visits in which BP was within the target BP range and the ratio between the number of visits with BP control and the overall number of available visits [190]. These data emphasize the importance of avoiding a long titration period during which patients remain with an uncontrolled BP, particularly if their CV risk is high, which supports the recommendation to start treatment with two drugs (see Section 11). Assuming that an uncontrolled BP at a given visit reflects an uncontrolled BP during the previous between-visit interval, they also strongly emphasize the importance for a physician to pursue a consistent BP control across visits, without considering a high BP at a single visit just an occasional and fortuitous finding. The limitations of outcome data based on average on-treatment BP values suggest that future trials in hypertension should consider a time-based analysis of BP control.

10.4 Residual risk

Although substantially reducing hypertension-related outcomes, effective antihypertensive treatment does not return the CV risk of treated hypertensive patients to the level of normotensive people when matched for age, sex and ethnicity. In other words, even when treatment achieves the target BP value that is believed to maximize CV protection, CV risk is not normalized, the excess risk being referred to as ‘residual risk’ [526–528]. The magnitude of the residual risk varies between studies, but it has been shown to be consistent up to FU of about 20 years [528,529]. Residual risk can originate from a considerable number of factors, some of which are still hypothetical, while others are supported by evidence and can be considered for therapeutic interventions. The possibility exists that the genetic component of hypertension includes a portion of irreversible risk. It is also possible that the partial risk irreversibility is generated by late treatment, i.e. by interventions made when risk is too high [485] and alterations of CV structure and function are at least in part irreversible. On the other hand, studies performed in the last few decades have provided evidence that in treated hypertensive patients, control of concomitant CV risk factors is poor [530–532], with an obvious adverse impact on total CV risk. Furthermore, in hypertension, CV risk depends on BP aspects that are much more complex than office BP as averaged across a number of treatment years. In individual patients, the optimal office BP target may depart from the average values reported in trials, and residual risk may be generated by failure of apparently effective antihypertensive treatment to reduce the elevated 24 h BP variability that characterizes hypertension and adversely affects the CV risk profile [180]. Coexistence of office BP control with imperfect control of the prognostically important real-life mean BP values and peaks may also be involved (see Sections 4.7 and 4.8). In this context, interesting progress has been made in recent years by the observation that in treated hypertensive patients, the risk of CV events depends on the time in which BP is controlled, independently of the mean BP value across the entire treatment period (see Section 10.3). In a post hoc analysis of the OnTarget-Transcend study, a greater visit-to-visit BP variability (which reflects the degree of BP inconsistency between different treatment periods [188]) was shown to increase the risk of events when added to mean on-treatment BP [189]. Thus, in addition to pursuing a better control of concomitant risk factors, another important practical intervention against residual risk might be to pursue assessment of consistency of BP control during FU, possibly by a widespread use of HBPM.

10.5 Challenges associated with evidence on BP targets

For guidelines, recommending the BP targets to reach with treatment is a difficult task. First, recommendations on optimal (i.e. most protective) target BP values may vary between different guidelines, and often also within guidelines at a few-year intervals, despite little concomitant increase in available evidence. This is due to the limited consistency of the available data, an example of which is that the incremental benefit of lowering SBP to <130 mmHg shown by some RCTs and meta-analyses does not clearly emerge from other major trials. Second, as mentioned by the present guidelines (Sections on special conditions), evidence on optimal BP targets is not always univocal or equally strong for all clinical hypertensive subgroups. An example can be offered by people with LVH, quite a common condition in hypertensive patients. A real-life study from a large Korean population has reported a lower risk of CV outcomes in hypertensive patients with ECG-LVH in whom SBP was reduced to <130 or even <120 mmHg compared with higher on-treatment values [533]. In contrast, in the LIFE Study on ECG-LVH patients, all-cause mortality increased with an on-treatment SBP <130 mmHg [534] compared with patients who had an average on-treatment SBP >130 mmHg. Likewise, in a post hoc analysis of the high-risk hypertensive population of the VALUE Trial [524], cardiac and all-cause mortality was found to be greater in the group that had ECG-LVH and achieved on average a SBP <130 mmHg, at variance from hypertensive patients with no LVH in whom an SBP reduction <130 mmHg was protective [535,536]. Pathophysiologically, this may be explained by the higher oxygen consumption associated with the increased LVM as well as by the degenerative changes in the microcirculation that accompany a hypertrophic myocardium, both factors making it more susceptible to a reduction of perfusion pressure. In a small old mechanistic study [537], a progressive DBP reduction to about 70 mmHg (nitroprusside infusion, intra-arterial measurement) did not affect coronary blood flow (coronary sinus thermodilution measurement) in hypertensive patients without LVH, whereas in hypertensive patients with LVH, coronary blood flow showed a steep reduction as DBP decreased below 90 mmHg. Finally, post hoc analyses of RCTs, have often shown an increased risk of outcomes in the general trial population or in a usually limited number of patients exposed to intense BP-lowering treatment, i.e. a J-shaped relationship between BP values and outcomes. Although the observational nature of a post hoc approach does not allow to establish whether the increased risk seen at lower BP values is caused by vital organ hypoperfusion or by an originally high risk or frailty status [538], this shows that hypertensive patients may not all uniformly respond to the same BP target. Future studies should expand knowledge on the optimal BP target for treatment in different clinical subgroups and try to clarify the factors and mechanisms behind its possible heterogeneity (Fig. 10).

F10
FIGURE 10:
Office BP targets in the general adult hypertensive population.aThe first objective of antihypertensive treatment should be to lower BP to <140/80 mmHg in most patients, because this accounts for the major portion of the protective effect of BP-lowering.aIf drug treatment is well tolerated, treated SBP values should be targeted to 130 mmHg or lower in most patients up to 79 years old.aDespite the smaller incremental benefit, an effort should be made to reach a BP range of 120–129/70–79 mmHg in patients up to 79 years old, but only if treatment is well tolerated. Evidence on the advantages of this lower BP target range is not available or unequivocal in a number of clinically important subgroups of patients (e.g. patients with LVH, CKD, or ISH). These issues are discussed in the sections on special conditions (see Sections 17 to 20).aIn patients at least 80 years old who are not frail, the first objective of antihypertensive treatment is to lower BP below 150 mmHg. However, a SBP target range between 130-139 mmHg may be considered, if well tolerated.aIn very frail patients, treatment targets should be individualized.

Office BP targets for drug treatment

figure24

11. ANTIHYPERTENSIVE DRUGS AND TREATMENT

In the 2018 ESC/ESH guidelines [4], five major drug classes were recommended as first-line agents for the treatment of hypertension i.e. ACEis, ARBs, CCBs, Thiazide/Thiazide-like diuretics and BBs. However, the recommendations included two particular weightings within this group of drug classes. First, the use of an RAS inhibitor (ACEi or ARB), if not contraindicated, was considered as a common component of the general combination treatment strategy, and second, the use of BBs was restricted to special clinical conditions or situations. The selection of these five drug classes was based on the following criteria:

  • 1. A proven ability to reduce BP as monotherapy.
  • 2. Evidence from RCTs that they reduce morbidity and mortality.
  • 3. A favorable tolerability and safety profile.

Based on RCTs and their meta-analyses [539–541], the 2018 guidelines concluded that all the above drug classes met the required criteria, i.e. they (i) effectively reduce SBP and DBP; (ii) reduce the major specific outcomes associated with hypertension when compared with placebo and (iii) exert an overall similar or only slightly different effect on the combined risk of major CV outcomes and mortality when given as the initial treatment step, albeit with some differences for some cause-specific outcomes (lesser stroke prevention with BBs and ACE-inhibitors, lesser HF prevention with CCBs and greater HF prevention with Thiazide diuretics) [539]. Additional considerations were that (iv) the reduction of events is due essentially to BP-lowering “per se” rather than to specific drug properties, which means that the greater the number of drug options, the greater is the opportunity of tailoring an effective BP-lowering treatment to the individual patient characteristics; (v) the antihypertensive effect of these five drug classes extends to ABPM and HBPM; (vi) their BP-lowering ability and protective effect include use in combination with other drugs as shown in RCTs in which BP-lowering treatment by multiple drugs was associated with a reduction of CV outcomes and (vii) side effects of drug classes are largely related to the doses employed and their between-class differences are minimized by use at lower doses in combination treatment. These considerations and recommendations are shared by the present guidelines, which also share the subordinate position attributed to other antihypertensive drugs (alpha-blockers, centrally acting agents and MRAs), because these drugs have been less widely studied in outcome-based RCTs, or are known to be associated with a higher risk of adverse effects. These drugs can be useful additions to the major antihypertensive armamentarium, in some specific cases, or when BP cannot be controlled by various combinations of the major drug classes.

Moreover, mention is made of new drug classes, such as SGLT2 inhibitors (SGLT2is) and nonsteroidal MRAs, which have become available and exhibit BP-lowering effects. These effects may be less pronounced than those of classical antihypertensive drugs [542], but there is now strong evidence from RCTs that they decrease CV and kidney events in patients with type 2 diabetes and – in the case of SGLT2i – also in patients without diabetes [543–547]. New criteria for drug performance are also discussed, such as the evidence of differences in the persistence and discontinuation rates of treatment between the major drug classes and even between drugs or drug combinations within a given class [548]. This has clinical relevance because antihypertensive treatment discontinuation leads to increased CV outcomes. Precise and correct prescriptions of drugs for documented CV or other medical conditions are among the most important decisions that can be taken by physicians to maintain or improve adherence and persistence to the prescribed drugs (see Section 21). A synopsis of the major drug classes and additional drug classed for BP-lowering therapy in hypertension is shown in Fig. 11.

F11
FIGURE 11:
Drug classes for BP-lowering therapy. (a) Use of Diuretics: Consider transition to Loop Diuretic if eGFR is between 30 to 45 ml/min/1.73 m2. If eGFR <30 ml/min/1.73 m2 use Loop Diuretic. (b) Non-DHP CCB should not be combined with BB. (c) BB should be used as guideline directed medical therapy in respective indications or considered in several other conditions (Table 16). Start with dual combination therapy including a RAS-blocker (either ACEi or ARB) plus a T/TLDiuretic or a CCB is recommended (thick blue lines). Triple therapy includes a combination of the three classes as indicated by the blue lines.

Compelling or possible contraindications for the selection of drug classes are summarized in Table 15.

TABLE 15 - Compelling contraindications and conditions requiring cautious use of BP-lowering drugs
Drug class Contraindications Cautious use
ACEi • Pregnancy
• Women planning pregnancy
• Previous angioneurotic edema
• Severe hyperkalemia (e.g. potassium >5.5 mmol/l)
• Bilateral renal artery stenosis or stenosis in solitary (functional) kidney
• Women of child-bearing potential without reliable contraception
ARB • Pregnancy
• Women planning pregnancy
• Severe hyperkalemia (e.g. potassium >5.5 mmol/l)
• Bilateral renal artery stenosis or stenosis in solitary (functional) kidney
• Women of child-bearing potential without reliable contraception
Beta-blocker • Severe asthma
• Any high-grade sino-atrial or atrioventricular block
• Bradycardia (e.g. heart rate <60 bpm)
• Asthma
• Glucose intolerance
• Athletes and physically active patients
DHP-CCB • Tachyarrhythmia
• Heart failure (HFrEF, class III or IV)
• Preexisting severe leg edema
Non-DHP-CCB (verapamil, diltiazem) • Any high-grade sino-atrial or AV block
• Severe LV dysfunction (LVEF <40%), HFrEF
• Bradycardia (e.g. heart rate <60 bpm)
• Co-medications susceptible to significant drug interactions mediated by P-gp or CYP3A4
• Constipation
Thiazide/Thiazide-like diuretics • Hyponatremia
• CKD due to obstructive uropathy
• Sulfonamide allergies
• Gout
• Glucose intolerance
• Pregnancy
• Hypercalcemia
• Hypokalemia
• Cancer patients with bone metastasis
MRA • Severe hyperkalemia (e.g. potassium >5.5 mmol/l)
• eGFR <30 ml/min/1.73 m2
• Co-medications susceptible to significant drug interactions mediated by P-gp or CYP3A4 for eplerenone

11.1 Blockers of the renin–angiotensin system (RAS)

11.1.1 Angiotensin-converting-enzyme inhibitors

ACEis are among the most widely used classes of antihypertensive drugs. Because of their earlier availability and, thus, their earlier evaluation against placebo in outcome-based RCTs, knowledge about ACEis is based on a large amount of RCT data, particularly in patients with HF, CAD and at high CV risk [549–551]. In these trials, ACEis have been shown to be associated with major benefits [550,551]. ACEis are associated with an increased risk of a very rare event such as angioneurotic edema, especially in people of Black African origin. They are associated with a cough that affects about 5– 10% of treated patients, although with a large variability (1.5– 11.0%) between studies and with a greater frequency in women and patients of Asian origin [552,553]. These side effects may affect their long-term persistence [554], although in a study on about half a million individuals, the discontinuation rate of people taking ACEis was considerably lower than that of BBs, CCBs and Diuretics and only modestly albeit significantly greater than that of people taking ARBs [514]. With the availability of several compounds, the class of ACEis represents a relatively homogenous class of drugs.

11.1.2 Angiotensin receptor blockers

ARBs have a similar antihypertensive efficacy and protective effect as ACEis [554–556], albeit with a somewhat different mechanism for RAS inhibition and a smaller RCT database. A difference between ACEis and ARBs is their tolerability profile, with ARBs having a rate of side effects similar to placebo. ARBs are associated with the lowest treatment-discontinuation rate when compared with all other antihypertensive therapies [514]. With several compounds’ availability, the class of ARBs represents a relatively homogenous class of drugs.

11.1.3 Renin inhibitors

The renin inhibitor aliskiren is a potent, long-acting antihypertensive drug when prescribed alone or in combination with a diuretic or a CCB [557]. Several RCTs on the impact of aliskiren on HMOD or CV outcomes have been conducted, but two of them were interrupted prematurely because of an increased incidence of adverse events, mainly when aliskiren was associated with another RAS blocker [558] Subsequently, aliskiren has almost disappeared from medical practice in European countries.

11.1.4 Combination of RAS inhibitors

ACEis, ARBs or aliskiren should not be combined because no added benefit on CV outcomes has been shown by dual RAS inhibition [559]. Although double RAS blockade has been shown to have a greater antialbuminuric effect and might have favorable effects in HF, the association may cause an excess of adverse events, with an increased risk of kidney function impairment (40%), hyperkalemia (44%) and hypotension (42%) [560] in patients with a high CV risk [559] or type 2 diabetes [561].

11.2 Calcium channel blockers

CCBs represent a heterogeneous class of drugs that can be divided into vascular-selective dihydropyridine (DHP) and non-DHP-CCBs. CCBs are widely used for the treatment of hypertension, and they are particularly effective in patients of African descent as well as in the general older population.

11.2.1 Dihydropyridine CCBs

Most RCTs demonstrating their benefits on outcomes have used DHP– CCBs, especially amlodipine [539]. DHP and non-DHP-CCBs differ in the tolerability profile and side effects. DHP-CCBs may be used, if necessary, to control an elevated BP in patients with HFrEF, although caution is needed because of their moderate negative inotropic effects. DHP-CCBs have only a limited potential for drug interactions.

11.2.2 Nondihydropyridine CCBs

Diltiazem and verapamil are not vascular-selective CCBs, but they are also efficacious in reducing BP. A smaller number of RCTs have compared non-DHP-CCBs with other drugs, while meta-analyses evaluating DHP and non-DHP-CCBs versus other drugs have not shown substantial differences in effectiveness [539]. Diltiazem and verapamil also belong to the class IV of antiarrhythmic drugs. Accordingly, they can delay atrioventricular conduction and slow heart rate in patients at sinus rhythm. DHP-CCBs are also used for heart rate control as an alternative to BBs in AF. Both drugs exhibit a potential for drug interactions because they inhibit the Cytochrome P450 3A4 enzyme and the drug efflux transporter P-glycoprotein, which can impair the tolerability and safety of other drugs, some statins (e.g. simvastatin, atorvastatin [562] or oral anticoagulants [563,564]) by increasing their plasma levels (see Section 17.3.4). Use of non-DHP-CCB is not recommended in HFrEF because of their pronounced negative-inotropic effect.

11.3 Diuretics

11.3.1 Thiazide/Thiazide-like

The effectiveness of Thiazide/Thiazide-like diuretics in preventing CV morbidity and mortality has been shown in RCTs and meta-analyses [539–541,565], with an effect similar to the effect of other major antihypertensive agents. In meta-analyses of RCTs, Thiazide/Thiazide-like appear to be more effective than other major drug classes in preventing HF, but this finding may be influenced by the results of the ALLHAT study [566], in which patients largely under a background diuretic treatment were rolled over to comparison drugs, with a possible emergence of HF symptoms previously under diuretic-based symptomatic control. The thiazide-like diuretics, chlorthalidone and indapamide, are more potent and have a longer duration of action compared with hydrochlorothiazide, but a greater incidence of side effects has been reported for chlortalidone in some studies [567]. A meta-analysis of placebo-controlled studies based on thiazides, chlorthalidone, and indapamide found similar effects for the three types of diuretics on CV outcomes [539]. A greater risk of CV events and HF has been reported with Thiazide diuretics in another meta-analysis [568]. Yet, no major difference between hydrochlorothiazide and chlorthalidone has been observed in a large observational cohort study [LEGEND] using a database spanning from January 2001 to December 2018 [569]. Furthermore, similar results have been obtained by a recent open-label study, DCP, on hypertensive US Veterans older than 65 years [570]. In this study, patients who were already on hydrochlorothiazide were randomized to either chlorthalidone (n = 6756) or hydrochlorothiazide continuation (n = 6767). Patients on treatment with hydrochlorothiazide 25 or 50 mg were converted to 12.5 or 25 mg chlorthalidone, respectively. No difference in CV outcomes between the two drugs was found, except for patients with a prior stroke in whom there was a greater benefit with chlorthalidone. Despite some limitations (in the last study, very few patients were on hydrochlorothiazide monotherapy at baseline, which means that the results could have been affected by concomitant medications and adherence to their use), the above-mentioned recent observations justify the recommendation of the present guidelines to still consider Thiazide/Thiazide-like diuretics both as suitable antihypertensive agents and as similarly effective in CV prevention. Both Thiazide/Thiazide-like can lower serum potassium and have a side-effect profile that is less favorable than RAS blockers. This may account for their higher rate of treatment discontinuation. Depending on the dose, they may also increase insulin resistance and, hence, the risk of new-onset diabetes. Potassium plays an important role in the metabolic effects of Thiazide/Thiazide-like, and evidence is available that these effects are reduced by the combination of Thiazide/Thiazide-like with a potassium-sparing diuretic [571,572] or with an RAS blocker. A recent placebo-controlled study [573] has demonstrated that chlorthalidone effectively lowers BP and albuminuria in patients with uncontrolled hypertension and CKD stage 4 (eGFR <30 ml/min/1.73 m2) when added to the therapy of these conditions, which often includes a loop diuretic (60% of the patients) (see Section 12).

11.3.2 Loop diuretics

Thiazide/Thiazide-like are considered less effective antihypertensive agents in patients with a reduced GFR (eGFR <30 ml/min/1.73 m2). Consequently, while loop diuretics (furosemide, bumetanide, torasemide/torsemide) are usually not indicated in the treatment of uncomplicated hypertension, they are recommended in patients with CKD stage 4 and 5 (eGFR below 30 ml/min/1.73 m2) and in patients with severe fluid overload/retention, e.g. in patients with HF or nephrotic syndrome. Recently, furosemide and torsemide were found to be equally effective on mortality, independently of ejection fraction in a RCT comprising 2859 HF patients receiving one drug or the other at discharge from hospital [574]. As indicated above, a loop diuretic can be combined with chlorthalidone to improve BP control in patients with advanced CKD stage 4 and uncontrolled or resistant hypertension [573].

11.3.3 Potassium-sparing diuretics

Amiloride is an agent that directly inhibits epithelial sodium channels at the luminal side of the late distal tubule and collecting duct. Hence, it is used as potassium-sparing diuretic in edematous states and for potassium conservation in combination with Thiazide or loop diuretics in hypertension or HF [575]. Amiloride was used for uncontrolled or resistant hypertension in the PATHWAY-3 [571] and PATHWAY-2 [576] studies. In the former study, the mean reduction in home SBP during a 24-week treatment did not differ significantly between patients taking amiloride (10– 20 mg once daily, -12.2 mmHg SBP) and those taking hydrochlorothiazide (25– 50 mg once daily, -12.9 mmHg SBP). However, the effect of the combination of low doses of amiloride (5– 10 mg) and hydrochlorothiazide (12.5–25 mg) was associated with greater reductions of BP than those obtained with the higher dose of hydrochlorothiazide alone. Amiloride had no impact on glucose tolerance in a sub-study of PATHWAY-2. Furthermore, a higher dose of amiloride, i.e. 10–20 mg once daily, as assessed during an optional 6– 12-week open-label runout phase reduced office SBP by 20.4 mmHg, compared with a reduction of 18.3 mmHg with spironolactone (25 mg once daily). No serious adverse events were recorded. Mean plasma potassium concentrations increased from 4.02 mmol/l on placebo to 4.50 mmol/l on amiloride (P < 0.0001) [576]. Triamterene is another potassium-sparing diuretic acting on the epithelial sodium channel. Its independent effects on BP have not been well studied, but the available evidence demonstrates an additive antihypertensive effect when associated with hydrochlorothiazide [577]. Triamterene is included as a compound of a quadruple combination that is available to treat hypertension in China [578].

11.4 Mineralocorticoid receptor antagonists

The steroidal MRAs spironolactone and eplerenone are established treatments in HFrEF based on outcome-based RCTs, but no outcome trial has been carried out in hypertension. Lack of outcome data and the risk of MRA-induced hyperkalemia and other side effects have restricted the use of MRA in the treatment of hypertension, except in specific conditions such as hyperaldosteronism or resistant hypertension. In resistant hypertension, a meta-analysis of 12 RCTs (1655 patients) [579] assessed the effect of spironolactone on BP compared with other therapies or placebo and showed a significant ability of spironolactone to lower BP, thus confirming the results of PATHWAY-2 [580] (see Section 12). Several new nonsteroidal MRAs are under investigation in patients with hypertension, type 2 diabetes or CKD. Among them, finerenone, a nonsteroidal MRA, has been shown to lower BP versus placebo [581] and to reduce cardiac and kidney outcomes in mostly (more than 95%) treated hypertensive patients with diabetic CKD [545–547]. In the Fidelio-DKD study, finerenone lowered SBP (-2.7 mmHg) when administered on top of an RAS blocker [582] mainly in patients who were hypertensive at baseline. A time-varying analysis revealed that 13.8 and 12.6% of the treatment effect of finerenone were attributable to the effect of the change in office SBP on the primary kidney composite outcome and the key secondary CV outcome, respectively [582]. With finerenone, hyperkalemia leading to treatment discontinuation occurred in 1.7% of patients [545] (see Section 19).

11.5 Beta-blockers

RCTs and meta-analyses have demonstrated that when compared with placebo, first-generation and second-generation BBs like propranolol, atenolol and metoprolol significantly reduce the risk of stroke, HF and major CV events in hypertensive patients. When compared with other BP-lowering drugs, BBs were almost equivalent in preventing major CV events, except for a less effective prevention of stroke [539–541,583,584]. It is possible that this difference on stroke data between BBs and other antihypertensive drug classes originates from small differences in achieved BP, including central SBP, to which cerebrovascular events may be especially sensitive. BBs are also associated with increased risk of new-onset diabetes in predisposed individuals (mostly those with the metabolic syndrome). They also exhibit a less favorable side-effect profile than that of RAS blockers, with a higher rate of treatment discontinuation when assessed in real-life conditions [514]. In previous guidelines [4,32,488], BBs were included among the five major antihypertensive drug classes. However, in the general treatment algorithm, they were recommended only when there is a specific indication, e.g. in patients with HF, angina, post-MI, AF or in younger hypertensive women of child-bearing potential or planning pregnancy. BBs do not constitute an homogeneous class but show several pharmacological differences, among which beta1-selectivity and an additional direct vasodilating property are of special interest. Third-generation BBs, such as nebivolol or carvedilol, exhibit direct vasodilating properties. Studies not only with nebivolol but also with bisoprolol, i.e. BBs with higher beta-1 selectivity and limited to nebivolol an added vasodilatation via increased release of nitric oxide, reported a more favorable side effect profile than other BBs, including fewer adverse effects on sexual function [585,586]. RCTs with carvedilol, bisoprolol, metoprolol and nebivolol showed improved outcomes in patients with HFrEF [587]. However, there are no outcome trials with vasodilating BBs in hypertensive patients, and the same applies to bisoprolol. There are also some recent large real-world studies with vasodilator BBs conducted in the USA, with inconsistent results. In one study, there was no statistically significant difference in CV outcomes between 118 133 patients receiving either nebivolol or carvedilol and 267 891 patients receiving atenolol [588]. In other studies, use of nebivolol led to greater CV protection compared with use of atenolol or metoprolol [589,590].

A recent pragmatic review scrutinized the use of BBs in medical treatments [591]. It was seen that, in addition to their compelling use as GDMT in specific diseases, BBs exhibit favorable effects in about 50 clinical conditions including (i) various cardiac diseases less or not related to hypertension, (ii) other vascular conditions and (iii) non-CV diseases (Table 16). In addition, concerns about the impact of BBs on psychological health should not affect their use in clinical practice, because a recent systematic review and meta-analysis of psychiatric adverse events during BB therapy indicated that BB use was not adversely associated with depression and most other psychiatric adverse events [592]. Finally, increased resting heart rate (>80 bpm) is common in hypertension, in which it reflects an increased sympathetic activity [593] A progressive increase in resting heart rate is accompanied by a progressive increase in the risk of AF, HF and mortality both in the general population and in hypertensive patients [594–596]. Although in hypertension, the advantage of reducing heart rate is limited to post hoc analysis of RCTs [597,598], the available evidence makes treated hypertensive patients with an increased heart rate a clinical phenotype supporting the use of BBs [599].

TABLE 16 - Selected diseases and conditions for the use of BBs in patients with hypertension [591]
Selected indications with guideline directed medical therapy for BBs
Chronic coronary syndromes, antiischemic therapy
Postmyocardial infarction: arrhythmias, angina, known incomplete re-vascularization, HF
Acute coronary syndrome
HFrEF and HFpEF if coronary disease (ischemia), arrhythmias and tachycardia
Atrial fibrillation: prevention, rhythm control, heart rate control
Women with child-bearing potential/planning pregnancy
Hypertension disorders in pregnancy
Selected other conditions in which therapy with BBs can be favourable
Hypertension with elevated resting heart rate >80 bpm
Emergency, urgency and parenteral administration
Perioperative hypertension
Major noncardiac surgery
Excessive pressor response to exercise and stress
Hyperkinetic heart syndrome
Postural orthostatic tachycardia syndrome
Orthostatic hypertension
OSA
Peripheral arterial disease with claudication
COPD
Portal hypertension, cirrhosis-related esophageal varices and recurrent variceal bleeding
Glaucoma
Thyrotoxicosis, hyperthyroidism
Hyperparathyroidism in uremia
Migraine headache
Essential tremor
Performance anxiety and anxiety disorders
Psychiatric disorders (posttraumatic stress)

11.6 Alpha-1 blockers

Treatment with the alpha-1 blocker doxazosin was equally effective as chlorthalidone in preventing the primary endpoint in ALLHAT [566], which was incident or fatal CAD. However, doxazosin was associated with a marked increase of incident HF, which led to stopping the doxazosin arm early. This was a controversial decision, because the increased incidence of HF in the doxazosin arm could have resulted from a HF misdiagnosis due to doxazosin-related fluid retention as well as discontinuation of diuretic treatment in patients with background HF in order to randomize them to doxazosin. In the ASCOT trial [600], doxazosin was given as a third-line therapy, and it showed no increase in the risk of HF. In the PATHWAY-2 study, it was more effective than placebo but less effective than spironolactone at lowering BP in resistant hypertension [580]. Alpha-1 blockers may also be required in specific conditions (e.g. treatment of benign prostatic hyperplasia). Orthostatic hypotension and fluid retention may be a concern with the use of alpha-1 blockers, especially in the older patients [601]. In a real-life study on older patients, administration of alpha-1 blockers in newly treated old hypertensive patients was followed by a significant increase of hospitalization for hip fracture over the following month, presumably as a result of orthostatic BP reductions and injurious falls [602].

11.7 Centrally acting drugs

In recent decades, centrally active drugs have been less frequently used, principally because of the lack of evidence by outcome RCTs and/or their poorer tolerability relative to the newer major classes of drugs. Thus, older compounds such as reserpine, alpha-methyldopa, clonidine, moxonidine or rilmenidine are no longer recommended for the routine treatment of hypertension and are primarily reserved for add-on therapy in the rare cases of resistant hypertension where other treatment options have failed or for specific conditions such as the use of methyldopa in pregnancy. It is worth noting that a recent study demonstrated a significant BP-lowering effect of clonidine in resistant hypertension that was similar to that of spironolactone [603](see Section 12).

11.8 Vasodilators

Vasodilators are a heterogeneous group of drugs, which exert a direct relaxing effect on vascular smooth muscle cells, thereby reducing BP via vasodilation and reduction of systemic vascular resistance. Powerful vasodilators, such as hydralazine and minoxidil were occasionally used in the past, but they have been now virtually abandoned because of their association with serious side effects. Side effects include marked baroreflex activation with tachycardia and increased activity of the sympathetic nervous system and activation of the RAS system resulting in tachycardia and fluid retention. Hydralazine may be occasionally considered in resistant hypertension that is unresponsive to multiple attempts to control BP, always in combination with BBs and diuretics to limit its side effects. A serious specific side effect for minoxidil is hirsutism. Nitrates and nitroprussiate also relax vascular smooth muscle cells and can lower BP. Intravenous nitroprussiate has a very efficacious and well controllable BP-lowering effect and is, therefore, used to manage hypertension emergencies (see Section 16.2).

11.9 Angiotensin receptor-neprilysin inhibitor (ARNI)

ARNI is a chemical combination of the ARB valsartan and the neprilysin inhibitor sacubitril, which simultaneously blocks the effects of angiotensin II at the AT-1 receptor (by valsartan) and inhibits the degradation of natriuretic peptides, thus promoting peripheral vasodilatation (by sacubitril) [604]. Initial studies including a successful phase 2 RCT, which showed compared to valsartan (320 mg) significant and fully additive reductions of SBP and DBP by treatment with sacubitril/valsartan [605], were performed in patients with hypertension, but the manufacturer later switched the focus to the treatment of HF. Currently, in most countries, the drug is therefore only approved for the treatment of HF (see Section 17.2.2 and 17.2.3). A recent meta-analysis of 10 studies including 5931 hypertensive patients confirmed the significant greater BP-lowering efficacy of sacubitril/valsartan when compared with other treatments including comparisons with valsartan, olmesartan, and amlodipine [604]. Sacubitril/valsartan is not approved for the treatment of hypertension in Europe or in the USA, while it was approved as an antihypertensive agent in China and Japan.

11.10 Antihypertensive drug combinations

11.10.1 Impact on hypertension drug treatment strategy

Guidelines have generated a variety of different strategies to initiate and increase BP-lowering medications in order to control an elevated BP. Before 2018, guidelines largely focused on the stepped care approach, initiating treatment with a variety of different monotherapies and then sequentially adding other drugs, until BP control was achieved. Starting with a two-drug combination was proposed only for patients with marked BP elevations and/or a high/very-high CV risk. Despite this, BP control rates have remained poor in Europe and worldwide. Failure to achieve BP control in most hypertensive patients, despite numerous iterations of guidelines, suggested that, whatever the reasons, the step care treatment strategy was not sufficiently effective and that a different approach was needed. The new strategy elaborated by the 2018 ESC/ESH guidelines was based on the following main considerations:

1. Efficacy of pharmacological therapies. Evidence from RCTs investigating BP responses to antihypertensive drugs demonstrates that BP control can be achieved in most patients and that no more than 5– 10% of these patients exhibit resistance to the selected treatment regimen [606]. Thus, ineffective drug therapy is unlikely to be the source of the problem.

2. Physician or treatment inertia. Evidence suggests that medical inertia, i.e. failure to adequately intensify or up-titrate treatment, contributes to failure or delay of treatment initiation but exerts an important adverse role also on suboptimal BP control with many patients remaining on monotherapy and/or suboptimal drug doses [607,608]. This was found to be the main reason for lack of BP control in major RCTs [609]. ACCOMPLISH achieved BP control in approximately 80% of study participants (the highest BP control achieved in major antihypertensive treatment RCTs from the Western World), but most of the remaining uncontrolled patients had not been up-titrated [609]. As expected, treatment inertia is quantitively much more important in real-world practice where it can be rated as a major factor responsible for poor BP control [608]. Addressing clinical inertia with specific measures is associated with an improvement in BP control [610].

3. Patient adherence to treatment. Evidence is accumulating that adherence to treatment is a fundamental factor to consider. Studies using urine or blood assays for the presence or absence of medication have shown that low or partial adherence to treatment is frequent, in particular, among patients with an uncontrolled BP [611]. This was also shown in studies of the general population, in which adherence to treatment based on prescription refilling was less than 50% in half of the patients [476,612]. Poor adherence has been associated with increased risk of developing CV complications in several studies, as reviewed recently [478,479,613].

4. Insufficient use of combination treatment. BP is a multiregulated variable depending on many pathophysiological pathways. Therefore, monotherapy is likely to be inadequate or insufficient to control BP in most patients, and combinations of drugs, working through different mechanisms, are necessary to achieve BP targets in most people with hypertension [608,614]. Indeed almost all patients in RCTs have required combinations of drugs to control their BP [615].

5. Complexity of current treatment strategies. There is also evidence that adherence to treatment is adversely affected by the complexity of the prescribed treatment regimen [479,616]. Several studies have demonstrated that adherence to treatment is strongly influenced by the number of pills a patient was prescribed for the treatment of hypertension as well as by the dosing frequency; the higher the number of pills, the lower the adherence.

The above considerations suggest that the most effective evidence-based treatment strategy to improve BP control would be one that:

  • 1. encourages the use of combination treatment in most patients, especially in the context of lower BP targets
  • 2. promotes the use of simplified single-pill-based combination therapies to support long-term persistence to treatment
  • 3. Recommends initial combination treatment in most hypertensive patients, as evidence is available that compared to monotherapy, initial combination treatment bypasses the problem of inertial monotherapy [608], improves long-term adherence to treatment [617–619], is accompanied by a better short-and long-term BP control [617,620] and, in observational studies, reduces the incidence of outcomes [612,621–623]

These are the main reasons why the 2018 ESC/ESH guidelines recommended a new simple and pragmatic treatment algorithm applicable to most patients, with the use of SPC therapy as first-line initial therapy. Exceptions are frail and very old patients, because of the impairment of the baroreflex and greater risk of hypotension, and very high-risk patients with high-normal BP, in whom the small BP reduction associated with monotherapy may allow to reach the BP target. For the same reason, initial monotherapy may also be considered in grade 1 hypertensive patients at low risk with a BP only modestly elevated above the BP threshold for drug treatment.

11.10.2 Drug combinations

Among the large number of RCTs of antihypertensive therapy, only a few have directly compared two different drug combinations, with systematic use of the two combinations in both arms. In most trials, treatment was initiated using monotherapy in either arm and another drug (and frequently more than one drug) was added, usually in a nonrandomized fashion, according to a prespecified treatment algorithm. In a few trials, such as ALLHAT [566], the design precluded the use of what could be optimal combinations, because multiple monotherapies were evaluated. With this caveat, Table 17 shows that a variety of drug combinations have been used in at least one active arm of placebo-controlled trials and have been associated with significant benefit on major CV events.

TABLE 17 - Major drug combinations used in trials of antihypertensive treatment in a stepped approach or as a randomized combination
Trial Comparator Type of patients SBP difference (mmHg) Outcomes (change in relative risk)
ACEi and diuretic combination
 PROGRESS [629] Placebo Previous stroke or TIA −9 –28% strokes (P <0.001)
 ADVANCE [630] Placebo Diabetes −5.6 –9% micro/macrovasc. events (P = 0.04)
 HYVET [502] Placebo Hypertensive; ≥80 years −15 –34% CV events (P < 0.001)
ARB and diuretic combination
 SCOPE [631] Diuretic + placebo Hypertensive; ≥70 years −3.2 –28% nonfatal strokes (P = 0.04)
 HOPE-3 [492] Placebo Patients at intermediate CV risk without CV disease (38% hypertensive patients) −6 NS overall difference in CV events but
-27% in CV events in patients with baseline BP>143.5 mmHg
ARB and CCB
 OSCAR [632] ARB Older, high-risk hypertensive patients −2.4 NS overall difference in CV events -31% events, patients with CV disease (P=0.02)
CCB and diuretic combination
 FEVER [633] Diuretic + placebo Hypertensive −4 –27% CV events (P < 0.001)
ACEi and CCB combination
 Syst-Eur [498] Placebo Older with ISH −10 –31% CV events (P < 0.001)
 Syst-China [145] Placebo Older with ISH −9 –37% CV events (P < 0.004)
BB and diuretic combination
 Coope and Warrender [634] Placebo Older hypertensive −18 –42% strokes (P < 0.03)
 SHEP [635] Placebo Older with ISH −13 –36% strokes (P < 0.001)
 STOP-Hypertension [636] Placebo Older hypertensive −23 –40% CV events (P = 0.003)
 STOP-Hypertension 2 [637] ACEi or conv. antiHT Hypertensive 0 NS difference in CV events
Combination of two RAS blockers/ACEi + ARB or RAS blocker + renin inhibitor
 ONTARGET [638] ACE inhibitor or ARB High-risk patients More renal events
 ALTITUDE [561] ACE inhibitor or ARB High-risk diabetic patients More renal events
Combinations versus placebo or monotherapy

In trials comparing different regimens [624] (Table 18), all combinations have been used in a larger or smaller proportion of patients, without major differences in benefits. The only exceptions are two trials in which a large proportion of the patients received either an ARB/diuretic combination or a CCB/ACEi combination, with both regimens being superior to a BB (atenolol)/diuretic combination in reducing CV outcomes [600,625]. Three outcome trials directly compared two different combinations, each involving a combination of an RAS blocker (ACEi or ARB) and a CCB with other combinations. In the ACCOMPLISH study [626], the ACEi/CCB combination was superior to the same ACEi in combination with a Thiazide diuretic at preventing major CV outcomes and CKD progression, despite only a small difference in BP between the two arms (SBP/DBP: 0.7/1.7 mmHg). The ACCOMPLISH finding [626] was not confirmed in the COLM and COPE trials, which reported no significant difference in CV events, when an RAS blocker– CCB combination was compared with an RAS blocker– Thiazide diuretic combination [627,628]. However, these two last trials were statistically underpowered, and their results on the outcomes, thus, have limited value.

TABLE 18 - Major drug combinations used in trials of antihypertensive treatment in a stepped approach or as a randomized combination
Trial Comparator Type of patients SBP diff (mmHg) Outcomes (change in relative risk)
ACE inhibitor and diuretic combination
 CAPPP [639] BB + diuretic Hypertensive +3 +5% CV events (NS)
 ACCOMPLISH [626] ACE inhibitor + CCB Hypertensive with risk factors +1 +21% CV events (P < 0.001)
ARB and diuretic combination
 LIFE [640] BB + diuretic Hypertensive with LVH −1 –26% stroke (P < 0.001)
Calcium channel blocker and diuretic combination
 ELSA [641] BB + diuretic Hypertensive 0 NS difference in CV events
 CONVINCE [642] BB + diuretic Hypertensive with risk factors 0 NS difference in CV events
 VALUE [524] ARB + diuretic High-risk hypertensive −2.2 –3% CV events (P = NS)
 COPE [627] CCB + BB Hypertensive +0.7 NS difference in CV events or stroke
 CREOLE [643] ACEi+CA
ACEi+D
Black Hypertensives uncontrolled −0.14
−3.14
No outcome data; CCB+D and ACE+CCB superior to ACEi+D in BP control
ACE inhibitor and CCB combination
 NORDIL [644] BB + diuretic Hypertensive +3 NS difference in CV events
 INVEST [645] BB + diuretic Hypertensive with CAD 0 NS difference in CV events
 ASCOT [600] BB + diuretic Hypertensive with risk factors −3 –16% CV events (P < 0.001)
 ACCOMPLISH [626] ACE inhibitor + diuretic Hypertensive with risk factors −1 –21% CV events (P < 0.001)
Beta-blocker and diuretic combination
 CAPPP [639] ACE inhibitor + diuretic Hypertensive −3 –5% CV events (P = NS)
 LIFE [640] ARB + diuretic Hypertensive with LVH +1 +26% stroke (P < 0.001)
 ALLHAT [566] ACE inhibitor + BB Hypertensive with risk factors −2 NS difference in CV events
 ALLHAT [566] CCB + BB Hypertensive with risk factors −1 NS difference in CV events
 CONVINCE [642] CCB + diuretic Hypertensive with risk factors 0 NS difference in CV events
 NORDIL [644] ACE inhibitor + CCB Hypertensive −3 NS difference in CV events
 INVEST [645] ACE inhibitor + CCB Hypertensive with CAD 0 NS difference in CV events
 ASCOT [600] ACE inhibitor + CCB Hypertensive with risk factors +3 +16% CV events (P < 0.001)
Beta-blocker and CCB combination
 COPE [627] ARB+CCB Hypertensive +0.8 NS difference in CV events or stroke
ARB and CCB combination
 COPE [627] CCB + diuretic Hypertensive −0.7 NS difference in CV events or stroke
 COPE [627] CCB + BB Hypertensive −0.8 NS difference in CV events or stroke
 COLM [628] ARB + diuretic Older hypertensive 0 NS difference in CV events
Combinations versus other combinations.

Based on the results of outcome RCTs, recent meta-analyses and evidence of BP-lowering effectiveness, all five major drug classes can in principle be combined with one another, except for ACEis and ARBs (see Section 11.10). However, we recommend that treatment of hypertension should be preferentially based on combinations of an ACEi or an ARB with CCB or a Thiazide/Thiazide-like diuretic. These combinations are now widely available in a single pill and in a range of doses, facilitating flexible prescribing and up-titration from lower to higher doses. They also (i) limit potential adverse effects associated with diuretic or CCB monotherapy, i.e. reduce the risk of hypokalemia due to diuretics, or the prevalence of peripheral edema due to CCBs and (ii) ensure that the RAS is inhibited as part of the treatment strategy, which is important for many patient groups (patients with diabetes, LVH, CKD with or without proteinuria etc.). Other combinations, such as CCB or BB plus a diuretic, also have RCT-based evidence supporting their use, as mentioned by the 2018 ESC/ESH guidelines. In addition, they can be the preferred combinations in a number of conditions. In black patients in sub-Saharan Africa, amlodipine plus either hydrochlorothiazide or perindopril was more effective than perindopril plus hydrochlorothiazide at lowering BP [643].

11.10.3 Rationale for initial two-drug combination therapy

As discussed above and with the emphasis in the present guidelines on achieving a BP target of <130/80 mmHg in most patients, the majority of patients will require combination therapy. Although no RCT has ever compared major CV outcomes between initial combination therapy and monotherapy, multiple arguments support combination of two antihypertensive drugs as the initial treatment step. One, initial combination therapy is invariably more effective at BP-lowering than monotherapy, and indeed even low-dose combination therapy is usually more effective than maximal dose monotherapy. Furthermore, the combination of medications targeting multiple mechanisms (i) reduces the heterogeneity of the BP response to initial treatment and (ii) provides a steeper dose– response effect than that observed with escalating doses of monotherapy and (iii) is safe and well tolerated, with no or only a small increase in the risk of hypotensive episodes, even when given to patients with grade 1 hypertension. Two, initial two-drug combination is associated with a faster BP reduction compared with monotherapy, and observational evidence suggests that the time taken to achieve BP control is an important determinant of clinical outcomes, especially in high-risk patients, with a shorter time to control associated with lower risk risk [646]. Three, evidence from the more general hypertensive population shows that compared with patients on initial monotherapy, those who start treatment with a two-drug combination reach more frequent BP control after 1 year [620], probably because initial combination treatment prevents therapeutic inertia [608,617], and initial two-drug combination is associated with a better long-term adherence and persistence [619] to the prescribed treatment regimen. Studies from large treated cohorts of patients under antihypertensive treatment have also shown that initial combination treatment resulted in a lower risk of CV events compared with initial monotherapy followed by the traditional stepped-care approach [608,622,623,647]. In one study, it was possible to analyze more than 2200 patients who experienced during the FU (1 year) a hospitalization for CV disease, while also showing a shift from initial combination treatment to monotherapy or vice versa. The results of this within-patient comparison (which removed a major limitation of observational studies, i.e. confrontation of external and possibly different patient groups) showed that the risk of hospitalization was much reduced when patients were on combination treatment compared to when they were on single-drug therapy [648].

General recommendations for antihypertensive drug treatment

figure26

11.10.4 Up-titration of treatment to three-drug combination

Evidence from RCTs shows that two-drug combination therapy will control BP in approximately half to two-thirds of patients [649]. For patients whose BP is not controlled by two-drug combination therapy, an option may be to use a different two-drug combination, or as suggested by the ISH guidelines [32], to use the same two-drug combination at higher doses of the combination components. A third logical option, however, is treatment with three-drug combination therapy, usually an RAS blocker, a CCB and a Thiazide/Thiazide-like diuretic. A three-drug combination can control BP in up to 90% of patients, which is a rate of BP control that is much greater than the current rate of BP control across Europe in treated hypertensive patients. There is also evidence that an SPC of three drugs achieves better BP control than usual care [650]. The present guidelines do not recommend to start treatment with a three-drug combination because of the risk of a BP reduction that is too fast and/or excessive, particularly in older patients.

11.10.5 Rationale for single-pill combination therapy

The 2013 and 2018 ESH/ESC guidelines [3,4] favored the use of two antihypertensive drugs as an SPC, because reducing the number of pills to be taken daily improves adherence to treatment and increases the rate of BP control. This recommendation is endorsed by the present guidelines. Use of SPCs is further supported by data from recent studies using various methods to assess adherence to treatment, including quantification of antihypertensive drugs in urine and blood, estimates such as prescription refills and calculation of the percentage of days covered by the treatments, which, although indirectly, enable to measure adherence on a prolonged basis, thereby accounting for its time-variable nature [479,616,621,647]. These studies have unequivocally shown a direct inverse relationship between the number of pills and the likelihood of adherence. This approach is now facilitated by the availability of several SPCs with a range of dosages, which eliminates the often-stated disadvantage of SPC therapy, i.e. the inability to increase the dose of one drug independently of the other. It is also convenient that the most widely available SPCs mirror the major drug class combinations recommended by the present guidelines. The major advantage of an SPC as the usual therapeutic approach for hypertension is that patients can progress from one, two or three drug treatments, remaining on a simple treatment regimen with a single pill throughout, thus increasing the likelihood of adherence to therapy while progressing to BP control. Such an approach can markedly increase the percentage of patients achieving high adherence to treatment (e.g. >80% of the treatment time covered by prescription) while markedly reducing patients characterized by low adherence to treatment (e.g. <20% of treatment time covered by prescription) with a clear reflection on patients’ protection independently from age, sex, co-treatment and clinical status [647]. It showed the potential to double BP control rates in treated patients from the present low level also with an improvement of outcomes [651]. SPCs of a BB plus a diuretic or a CCB have been available since many years, while at the time of the 2018 ESC/ESH Hypertension Guidelines, additional SPCs were almost exclusively limited to a RAS blocker (ACEi or ARB) plus a CCB or a diuretic. In the 5 years from the 2018 Guidelines, a large number of new two-drug SPCs have been developed and tested for their ability to improve adherence to treatment and reduce CV outcomes. Available two-drug SPCs now extend to most ACEis or ARBs in combination with a long-acting CCB or a diuretic belonging to the Thiazide (usually hydrochlorothiazide) or Thiazide-like (indapamide or chlorthalidone) class. Moreover, two-drug SPCs are now available for a RAS blocker (ACEi or ARB) with a BB, including SPC containing nebivolol with additional vasodilatory action, and a CCB with a diuretic (e.g. amlodipine plus indapamide or nifedipine plus a Thiazide). The availability of three-drug SPCs has also grown and, although almost invariably based on a diuretic, a RAS blocker and a CCB, it now extends to different compounds within each of the three drug classes involved. This enables to tailor SPC treatment to different clinical requirements [647,651–653].

11.10.6 The quadpill concept

Another innovative therapeutic approach to increase BP control while improving tolerability is to use combinations of low or even ultralow doses of the recommended antihypertensive drugs. With this approach, the ability to effectively reduce an elevated BP appears to be maintained while most side effects are avoided. A proof-of-concept study and a systematic review of quarter-dose BP-lowering drugs was reported in 2017 [654,655]. The systematic review included 36 trials (n = 4721 participants) of one drug at quarter-dose and 6 trials (n = 312) of two drugs at quarter-dose against placebo. The pooled placebo-corrected SBP/DBP-lowering effects were, respectively, 5/2 mmHg and 7/5 mmHg with no reported side effects. The BP-lowering effect was even greater when quarter-doses of four drugs, i.e. quadpill, were used. These preliminary data have been supported by a phase 3 study by the same authors, in which 591 patients were randomized to the quadpill or to full-dose monotherapy with an ARB. BP changes were assessed at 12 weeks and 12 months. The BP control rate was greater in the quadpill group at both time points. Studies in broader populations are needed, and more information is required on a number of issues such as the strategy to adopt when side effects occur or hypotension develops using the quadpill approach.

11.10.7 The polypill concept

Polypills consist of SPCs of one or two antihypertensive agents and a statin with or without low-dose acetyl salicylic acid (aspirin) [656]. Different doses of antihypertensive agents, usually including an ACEi, are available. The rationale is that (i) hypertensive patients often also have dyslipidemia and an elevated CV risk and (ii) treatment simplification, i.e. a single pill rather than multiple pills daily, improves adherence to treatment, which is low in hypertension [479,612,657–659]. Bioequivalence studies suggest that, when combined in the polypill, different agents maintain their expected effect [660], including the BP-lowering efficacy. Furthermore, studies performed in the setting of secondary CV prevention, particularly in patients with a previous myocardial infarction, have shown that use of the polypill is accompanied by better adherence to treatment compared with separately administered medications [656]. This is true for treatment simplification in general. Based on this evidence, the use of the polypill has been recommended for the management of myocardial infarction [660]. However, present data also document that the polypill reduces the risk of CV outcomes. This was initially reported by large observational studies in patients with established atherosclerotic CVD [656] and has more recently been proven by the results of large outcome-based RCTs in patients with and also without previous CV events [661,662]. In an individual-participant meta-analysis of three primary prevention trials, a combination of two antihypertensive agents and a statin at low doses reduced the risk of CV outcomes by 38%. A polypill including low-dose aspirin was associated with a nearly 50% outcome reduction. The benefits were seen across various subgroups (different lipid and BP levels, diabetic patients, smoking, obesity) with the smallest effect in patients <55 years of age [663]. In a fourth RCT in patients with a recent myocardial infarction, the polypill (aspirin, ACE-inhibitor and statin) treatment strategy reduced the risk of the primary outcome (CV death, myocardial infarction, ischemic stroke and urgent revascularization) by 24% compared with usual care, again across subgroups with different clinical characteristics and with the additional evidence of an improved adherence to treatment in the polypill group. Adverse events were similar between groups and the most common adverse event in the polypill group was dizziness [662]. The above evidence supports use of currently available polypills in hypertensive dyslipidemic patients at elevated CV risk. Polypills without low-dose aspirin may be used in primary prevention, while use of those with aspirin should be restricted to secondary prevention. The previously issued recommendation to check the efficacy of the combination components in separate tablets before switching to the polypill appears impractical [4,664]. Potential inconveniences may be the limited dose flexibility of the polypill components as well as the limited potential of the available polypills to reach the lower LDL-cholesterol and BP targets at present recommended by guidelines. This may require the separate administration of additional drugs in a number of patients, with partial loss of the polypill advantages.

11.10.8 Choice of drug combinations for initiation of treatment

Reflecting on the evidence discussed above and recognizing the need to avoid or minimize the factors contributing to poor BP control in treated hypertensive patients, the following few simple and pragmatic recommendations for the treatment of hypertension can be listed (Figs. 11 and 12):

  • 1. In most patients, treatment should be initiated with an SPC of two drugs to improve the speed, efficiency and predictability of BP control.
  • 2. Although several two-drug combinations can be used, the preferred two-drug combinations should be an RAS blocker with a CCB or a Thiazide/Thiazide-like diuretic.
  • 3. A BB can be used at any step of combination with any drug from the other major drug classes as GDMT or in several other conditions (Table 16).
  • 4. Initial monotherapy is recommended for very-high-risk patients with a high-normal BP as well as (for cautionary reasons) for very old and frail patients. It may also be considered in low-risk patients with stage 1 hypertension whose SBP is more modestly elevated (<150 mmHg).
  • 5. An SPC comprising an RAS blocker CCB Thiazide/Thiazide-like diuretic, should be used if two-drug combinations do not achieve BP control (at the maximum tolerated doses) and a BB is not indicated.
  • 6. Regardless the initial treatment choice, ultimately most patients should be on combination treatment, using SPC whenever possible.
F12
FIGURE 12:
General BP-lowering strategy in patients with hypertension.aUse of Diuretics:–Consider transition to Loop Diuretic if eGFR is between 30 to 45 ml/min/1.73 m2–If eGFR <30 ml/min/1.73 m2 use Loop DiureticbBB should be used as guideline directed medical therapy in respective indications or considered in several other conditions (Table 16)cControlled below 140/90 mmHgd When SBP is ≥140mmHg or DBP is ≥90 mmHg provided that:–maximum recommended and tolerated doses of a three-drug combination comprising a RAS blocker (either an ACEi or an ARB), a CCB and a Thiazide/Thiazide-like diuretic were used–adequate BP control has been confirmed by ABPM or by HBPM if ABPM is not feasible–various causes of pseudo-resistant hypertension (especially poor medication adherence) and secondary hypertension have been excluded (see Section 12).

11.10.9 Tolerability and side effects of drugs

All antihypertensive drugs can elicit side effects, which can be modest or, in some instances, serious, leading to treatment discontinuation. Side effects play a major role for treatment nonadherence and discontinuation [665] and can be either related to BP-lowering “per se” or because of class-specific effects [666,667]. Nevertheless, the recommended major antihypertensive drug classes show by and large a good tolerability, which is one of the criteria that supports the recommendation for their use, in addition to their BP-lowering effect and proven outcome reduction. Side effects vary not only between different classes but also within a drug class, e.g. between different BBs and between DHP and non-DHP-CCBs. Fortunately, the potential for pharmacokinetic drug– drug interactions that may influence plasma concentration of drugs is marginal for any of the major BP-lowering drug classes, thereby providing the basis for their safe use in combination therapy. It should also be mentioned that there are clinically favorable interactions between major antihypertensive drugs, a most important one being their additive BP-lowering effect. Furthermore, there are interactions that increase drug tolerability, i.e. reduced incidence or intensity of ankle edema by adding an RAS blocker to a CCB or reduced hypokalemia by adding an RAS blocker to a Thiazide/Thiazide-like diuretic [665].

11.10.10 Prescribing of antihypertensive drugs

11.10.10.1 Standard drug administration

The primary goal of antihypertensive treatment is to provide BP control over the 24 h period (short-term) and to maintain this control over time (long-term). To achieve this goal, clinicians must consider (1) the half-lives and dose of the used drugs, as well as their interactions with the concomitantly used drugs (pharmacokinetics), (2) the mechanism of action of the drugs (pharmacodynamics) and (3) patient characteristics that might influence absorption, metabolism or elimination of the drug [668]. It is generally accepted that use of multiple daily doses during the day may achieve BP control at the expense of reduced patient adherence to treatment [479,612,657,659,669]. For this reason, drugs that have a duration of action that covers the 24 h period with a single daily administration should be preferred. This is not fulfilled by all agents [670] within major antihypertensive drug classes. This shortcoming has been accounted for by giving a drug at higher doses, which can also prolong the effect of short-acting agents albeit with the risk of inducing hypotension at the time of the peak effect. To allow once-daily drug administration, extended release formulations have also been developed. Different durations of action of antihypertensive drugs given once daily [670] may also affect short-term or long-term BP variability [180] and perhaps outcomes, but the extent of these influences is still unclear.

Most of the available evidence on the outcome benefit of BP-lowering therapy has come from RCTs using morning dosing of the drugs. However, recent evidence that night-time hypertension is not rare as well as that a nondipping profile may have adverse prognostic consequences [671] has favored the hypothesis that bedtime administration of antihypertensive drugs should be preferred to more effectively reduce night-time BP and CV outcome risk [671,672]. The limitations of the supporting data have been discussed by some critical articles [671,672] and by a systematic review of eight studies on the effects of morning versus bedtime dosing of antihypertensive agents. Reaching a conclusion was considered problematic because of major methodological limitations and bias [673]. In this regard, the recent TIME pragmatic trial [670] has provided important data. In TIME, 21 104 participants from the United Kingdom were randomized in a 1: 1 ratio to take their usual BP medications in the morning or in the evening. The average age of participants was 65 years and 58% were men. The median FU was 5.2 years, but some patients were followed up for more than 9 years [670]. Overall, no safety concerns were detected in the study. The reported nonadherence to therapy was significantly higher with evening versus morning dosing (39.0 versus 22.5%, P < 0.0001). However, there was no significant difference in the primary outcome (hospitalization for major CV events and vascular death) between the evening-dosing and the morning-dosing groups. Thus, data do not support preferential use of antihypertensive drugs at bedtime, which is, however, also not harmful. Based on TIME, patients have a choice as to when to take their medication, while physicians may consider bedtime dosing in patients with documented high night-time BP [674]. In general, the present guidelines recommend taking BP in the morning, as adherence to antihypertensive medication is worse at bedtime [670,674,675].

11.10.10.2 Partial treatment reduction or complete withdrawal

Complete withdrawal of antihypertensive drug therapy, because of side effects or other reasons, is accompanied by a more or less rapid return of BP to the pretreatment elevated values [676]. With centrally acting agents, especially clonidine, an abrupt rebound BP increase may occur, whereas abrupt withdrawal of BBs in patients with CAD may result in angina or other symptoms or complications of CAD, e.g. arrhythmias. Headache, joint pain, palpitations, edema and a general feeling of being unwell have also been reported [676]. A marked increase in CV risk after antihypertensive treatment withdrawal has been documented in many studies [494,677,678], although no adverse consequences have also occasionally been reported [676,679]. Some studies observed no or little BP increases after treatment withdrawal in a sizeable fraction (up to 25%) of the hypertensive population, but the interpretation of this finding is uncertain and possibly related to an erroneous hypertension diagnosis or withdrawal of hypertensiogenic risk factors such as overweight [676] during the treatment period. The possibility that a long-term effective antihypertensive treatment reverses the structural changes of hypertension and favors prolonged BP normality after treatment cannot be excluded. To date, little or no evidence exists on the BP and outcome effects of partial deprescribing of BP-lowering drugs (Table 19).

TABLE 19 - Withdrawal of BP-lowering drugs
Drug class Effects of abrupt withdrawal
CCBs • Risk of angina
BBs • Risk of angina and other complications in CAD patients
Centrally acting agents • Sympathetic overactivity (nervousness, tachycardia, headache, agitation and nausea for 36–72 h after drug cessation)
• Rapid rebound BP increase even above pretreatment levels
• Angina and other complications in CAD patients
Diuretics • Fluid retention, edema, HF decompensation

11.10.10.3 Antihypertensive drugs and cancer risk

Whether hypertension “per se” or treatment with antihypertensive drugs may influence the risk of cancer has been a matter of debate for many years [680–682]. More recently, two case–control studies suggested that the use of hydrochlorothiazide is associated with an increased risk of developing squamous cell carcinoma in the skin and lip [683,684]. This was not the case with other diuretics or antihypertensive drugs. However, although it cannot be denied that most diuretics (including hydrochlorothiazide) are potentially photosensitizing drugs [681,685] and that this can be a basis for an adverse influence on skin cancer, the studies had methodological limitations [681] and were confirmed by some, but not by all, subsequent studies carried out in other countries or ethnicities [686]. The issue is confusing also because a recent study from the UK Clinical Research Database again also reported an increased risk of skin squamous cell carcinoma in users of hydrochlorothiazide. However, the increased risk was seen only in women, and no biological rationale was offered for this finding [687]. Finally, a large propensity-matched cohort study in Germany found that hydrochlorothiazide was associated with not only a small increase of skin cancer risk but also with an overall lower risk for any cancer compared with other diuretics [688]. Further well designed observational studies are needed to provide a more solid evidence on the association between hydrochlorothiazide and skin cancer risk. In the meantime, important evidence to quote is that in a recent individual participant data meta-analysis of 33 RCTs, involving 260 447 participants and 15 012 cancer cases, no significant association between Thiazides (including hydrochlorothiazide) and skin cancer was observed [689]. Furthermore, no significant association with any cancer was found for ACEis, ARBs and BBs. Only for CCBs, a small but significant effect size was found for all cancer risk (hazard ratio 1.06, 95% CI 1.01–1.11). Although the relatively short duration of the trials (slightly more than 4 years) represents a limitation, this supports the recommendation not to consider cancer risk a barrier to any drug management of hypertension including hydrochlorothiazide, which is a frequent component of SPCs [690] and has a documented protective effect.

11.10.11 Concomitant medications

11.10.11.1 LDL-cholesterol lowering

Hypertension and dyslipidemia are highly prevalent in the general population and often coexist, contributing to CV risk in an additive way [531]. Lipoproteins in plasma are classified according to size, and differ with respect to lipid content and apolipoprotein expression [691]. Several different lipid-and lipoprotein measures (including total cholesterol, non-HDL cholesterol, LDL cholesterol and triglycerides) have been associated with the risk of CVD in a log-linear fashion. Current evidence suggests that all ApoB-containing lipoprotein with a diameter <70 nm may cross the endothelial barrier and contribute to the formation of atherosclerosis. This explains why different lipid measures are used in different contexts, such as non-HDL cholesterol in the SCORE2 and SCORE2-OP models to estimate CV risk, whereas treatment targets are generally guided by levels of LDL-cholesterol. These guidelines support the concept of risk-based lipid-lowering treatment suggested by the 2021 ESC guidelines on CVD prevention [33]. Apparently healthy adults should undergo a CV risk assessment using the SCORE2 or SCORE2-OP tools [33]. In low-risk patients, no specific treatment is needed except for general lifestyle advice. For high-risk or very-high risk primary preventive patients, it is recommended to initially aim for a LDL-cholesterol level below 2.6 mmol/l (100 mg/dl). Depending on 10-year and lifetime risk, comorbidities, frailty and patient preference, the ultimate goal may be an LDL-cholesterol level below 1.8 mmol/l (70 mg/dl) in high-risk patients, and below 1.4 mmol/l (55 mg/dl) in very high-risk patients [33].

The ASCOT trial [692] demonstrated that in people with hypertension and moderately elevated CV risk, treatment with 10 mg/day of atorvastatin reduced the risk of composite CV outcome by 36%. Numerous RCTs and multiple meta-analyses have shown that statin treatment is associated with a reduction of CV outcomes that is proportional to the LDL-cholesterol reduction [693]. In recent years, evidence from RCTs on ezetimibe and PCSK9-inhibitors have accumulated, adding to the overall body of evidence supporting the causal role of LDL-cholesterol reduction for CV protection [694]. Importantly, the relative benefit of lipid-lowering treatment in patients at moderately elevated risk seems to be independent of BP level [695,696], although the absolute benefit is more pronounced in people with hypertension because of elevated CV risk [697]. Based on the results of the ASCOT and HOPE-3 trials [692,695], among others, people with hypertension and elevated CV risk should be treated with a moderate dose of a statin, whereas hypertensive patients classified as having a high or very high CV risk, thus fulfilling the criteria for intensive LDL-cholesterol-lowering treatment, the required lower LDL-cholesterol goals for CV prevention should be attained by uptitrating statins to the maximally tolerated dose [691]. It is of note that most side effects, including muscle symptoms, are nonspecific and not related to statin treatment “per se” [698]. According to the recent guidelines on dyslipidemia, ezetimibe should be added if LDL-cholesterol control is not achieved (preferably as SPC to improve adherence to treatment) [699] and PCSK9 inhibitors or siRNA may be considered in very high-risk patients to attain the LDL-cholesterol target.

In isolated triglyceridemia, statin therapyshould be the initial drug class of choice to reduce CV risk, and maybe considered if triglyceride levels are 2.3 mmol/l (200 mg/dl), especially in diabetic patients. Treatment with fenofibrate has been suggested to provide additional benefit, proportional to its effect on non-HDL-cholesterol [691,694]. The role of n-3 polyunsaturated fatty acids (PUFA) is uncertain because of the conflicting results of two major trials published in recent years [700,701].

On the antihypertensive treatment side, BBs and diuretics may be regarded as less preferable in difficult-to-treat dyslipidemia because of their modest dyslipidemic effects, more evident in combination treatment. However, their ability to reduce CV risk through BP-lowering greatly outweighs their metabolic downsides, and they should be used to control BP if necessary.

Recommendations for LDL-cholesterol-lowering therapy in hypertension

figure28
11.10.11.2 Antiplatelet therapy

Common complications of hypertension are related to atherothrombotic diseases, i.e. CAD, ischemic stroke, and LEAD [702]. The decision to recommend antiplatelet therapy in hypertension should be based on the individual CV risk, similarly to normotensive patients, i.e. according to their belonging to the primary prevention versus the secondary prevention setting, and to the bleeding risk. In secondary prevention, use of antiplatelet therapy [usually low-dose acetylsalicylic acid (aspirin)] is required, because in patients with established CVD, low-dose aspirin is associated with clinically important reductions of major CV events although with an increase of bleeding risk, especially from the gastrointestinal tract [703]. In primary prevention, a Cochrane systematic review [704] comprising 61 015 patients included in six trials (four trials in primary prevention, n = 41 695 patients; and two trials in secondary prevention, n = 19 320 patients) investigated the effects of antiplatelet agents and anticoagulants in hypertension. Four studies compared low-dose aspirin versus placebo and found no evidence of a difference in all-cause or CV mortality. However, aspirin treatment reduced the risk of all nonfatal CV events, albeit increasing the risk of major bleedings. The authors conclude that there is currently no evidence that antiplatelet therapy has a protective effect on hypertensive patients in the setting of primary prevention. The same conclusion had been reached more than 25 years ago in the HOT trial in which, however, some protective effect of low-dose aspirin was shown in a subgroup of patients with no previous CV events but with a high CV risk due to advanced kidney disease [705].

The benefits and harms of the newer drugs, i.e. clopidogrel, prasugrel and ticagrelor, have not been sufficiently studied in clinical trials on patients with hypertension.

Recommendations of antiplatelet therapy in hypertension

figure29

12. TRUE RESISTANT HYPERTENSION

12.1 Definition, prevalence, pathophysiology and cardiovascular risk

In the 2018 Guidelines, hypertension was defined as resistant to treatment when appropriate lifestyle measures and treatment with optimal or best tolerated doses of three or more drugs (a Thiazide/Thiazide-like diuretic, an RAS blocker and a CCB) fail to lower office BP to <140/90 mmHg [4]. The inadequate BP control should be confirmed by out-of-office BP measurement showing an uncontrolled 24 h BP (≥ 130 mmHg SBP or ≥ 80 mmHg DBP) values. Evidence of adherence to therapy and exclusion of secondary causes of hypertension are required to define true resistant hypertension, otherwise resistant hypertension is only apparent and termed as pseudoresistant hypertension (Fig. 13).

F13
FIGURE 13:
Characteristics of true resistant hypertension. RAAS, renin – angiotensin aldosterone system.

The prevalence of resistant hypertension has been difficult to quantify because of its dependence on a number of factors: the clinical setting (e.g. general population, tertiary referral center, clinical trial), (ii) the classes and optimal doses of the antihypertensive drugs used, (iii) the exclusion or retention of patients not adhering to treatment, (iv) the method of BP measurement and (v) the definition of the target BP value representing BP control [606,706,707]. Resistant hypertension can be very common in patients with advanced CKD [708]. Obviously, it is more common using the <130/80 mmHg than the <140/90 mmHg office BP values to define BP control [709]. It has a lower prevalence after removal of patients with normal out-of-office BP values, and the percentage of patients removed for this reason has been found to vary from about 14% to about 37% in meta-analyses of observational studies, randomized trials, surveys and registries [707,710,711]. After applying the strict criteria outlined above, a reasonable estimate of the resistant hypertension prevalence is that it might involve about 5% of the overall hypertensive population. These patients are referred to as having a true resistant hypertension phenotype. Contributing demographic and clinical factors are obesity (or large weight gains), excessive alcohol consumption, high sodium intake, advanced HMOD and atherosclerotic disease as well as older age, male sex, Black African origin, low income, depression, high BP values at hypertension diagnosis and a 10-year CV risk score >20% [712–715]. The pathophysiology of true resistant hypertension involves an interplay between multiple neurohumoral factors such as increased levels of aldosterone [580,716,717], endothelin-1 [718], vasopressin [719] and increased sympathetic activity [720]. These factors contribute to volume and sodium overload, increase in peripheral vascular resistance, arterial stiffness and more advanced HMOD, including cardiorenal damage [721–723]. Patients with resistant hypertension are at higher risk of HMOD [724], CKD [725] and premature CV events [726,727] (Fig. 13).

12.2 Diagnostic work-up

The diagnostic work-up should rule out pseudoresistant hypertension and, once true resistant hypertension is identified, characterize the patient's clinical status by history, physical examination and laboratory and imaging analysis for the assessment of risk factors and HMOD (see Section 5.5). Exclusion of pseudoresistant hypertension requires (i) the demonstration of an elevated ABPM; (ii) the exclusion of an origin of the BP elevation from inaccurate BP measurement, e.g. the spurious BP increase associated with marked brachial artery calcification, especially in older patients or in patients with advanced CKD; (iii) the exclusion of a secondary cause of hypertension (see Section 6) or (iv) the exclusion of poor adherence to the prescribed treatment regimen (see Section 21). The prevalence of secondary hypertension, especially primary aldosteronism [728] and atherosclerotic renal artery stenosis (particularly in older patients or patients with CKD) can be as high as 10– 20% of patients with resistant hypertension [716]. Search for OSA (by sleep history and specific tests) should not be omitted because of the frequent involvement of this condition in resistant hypertension, including night-time hypertension [729–733].

Poor adherence to antihypertensive drug treatment is common in hypertension [479], and in resistant hypertension, this is even more the case [616], because adherence is inversely related to treatment complexity and the number of prescribed daily tablets [479,612,616,657,659,669]. Precise information on adherence can be difficult to obtain. Careful inquiry about a patient's medical habits, if necessary with the help of the patient's relatives is the first step. Help can also be sought from some objective measures of adherence, which is unfortunately not easy to do [616]. Confirmation of adherence to antihypertensive medications may be provided by drug screening of urine or blood whenever available [616,659] or by pharmacodynamic markers of exposure to medications [bradycardia on BBs, increase blood levels of uric acid on diuretics, increase in plasma renin concentration on diuretics or RAS blockers, increases in urine N-acetyl-seryl-aspartyl-lysyl-proline (AcSDKP) concentration on ACEi [734] and specific drug-related side effects] [616]. History should include accurate information on use of drugs or substances potentially interfering with BP control either by impairing the efficacy of antihypertensive drugs or by increasing BP (Table 20). Careful evaluation of the drugs taken by patients is made easier by use of standardized questionnaires [735] or drug– drug interaction-checking applications or web-tools.

TABLE 20 - Medications and other substances that may increase BP
Medication/substance Proposed mechanism Comments
NSAIDs Inhibition of COX-1 and 2 decreasing PG I2 and E2 synthesis with subsequent reduction in urinary Na excretion and an increased systemic vascular resistance. Mild, dose-dependent increase in BP. Increased risk with age, preexisting hypertension, salt-sensitive patients, patients with renovascular hypertension.
Paracetamol (acetaminophen) Presumably via inhibition of cyclooxygenases and reduced production of prostaglandins. Increased relative risk of 1.34 of hypertension with almost daily paracetamol use.
Estrogens and progestins Increased renin synthesis (by estrogens) leading to RAS activation and subsequent Na+ and water retention. Mild, sustained increase in BP (6/3 mmHg increase with high doses of estrogen >50 μg of estrogen and 1–4 μg progestin) but can be severe, common in premenopausal women, cause hypertension in 5% of women.
Glucocorticoids Enhanced Na+ reabsorption and fluid retention via stimulation of mineralocorticoid receptors.
Increased systemic vascular resistance due to upregulation of AT1 receptors on vascular smooth muscle cells.
Dose-dependent, low doses have less effect on BP, more common in older patients, or with a family history of primary hypertension.
Calcineurin inhibitors Reduced NO production, ET-1 overproduction, systemic and renal vasoconstriction, renal Na+ retention. Dose-dependent, mild-to-moderate increase in BP. Severe hypertension has been reported. Increased risk with preexisting hypertension, elevated creatinine levels and maintenance therapy with corticosteroids. See Section 20.8.2
Antidepressants
SNRIs
Increased noradrenaline release causing adrenergic activation and increased SNS activity. Dose-dependent, mild (2/1 mmHg) increase in BP.
Nasal decongestants Vasoconstriction due to stimulation of alpha-1 receptors on vascular smooth muscles. Dose-dependent, sustained increase in BP.
Erythropoietin-stimulating agents Increased thromboxane, reduced prostacyclin levels and activation of the local RAS.
Increased ET-1 production, decreased NO synthesis with subsequent vasoconstriction.
Dose-dependent, mild increase in BP, increased risk with preexisting hypertension, or when the initial hematocrit level is low. See Section 20.8.2
Stimulants

- Modafinil
- Amphetamines
- Methylphenidate


Block noradrenaline or dopamine reuptake.
Promote release of catecholamines
VEGF inhibitors Decreased NO production via VEGFR-2 antagonism and stimulation of ET-1 receptors promoting vasoconstriction. A class effect. The incidence of hypertension is dose-related, risk is increased by preexisting hypertension, old age and overweight. See Section 20.8.2.
Substances of abuse

- MDMA
- PCP
- Methamphetamine

- Cocaine

- Alcohol
Increased release and inhibited reuptake of monoamine neurotransmitters with subsequent SNS activation.
Increased CNS catecholamine release with decreased neuronal uptake.

Cocaine induces acute sympathomimetic effects and chronic HMOD, i.e. an increase in arterial wall stiffness.

Alcohol increases SNS and RAS activity.
Cocaine induces both acute and chronic increases in BP.

Alcohol causes a dose-dependent, sustained increase in BP.
Herbal products

- Licorice
- Ephedra
- St. John's wort
- Yohimbine
- Ginseng (high doses)
- Ma huang
Chronic excessive liquorice use mimics hyperaldosteronism by stimulating the mineralocorticoid receptor and inhibiting cortisol metabolism.
Ephedra activates the alpha-1 receptor increasing SNS activity.
Licorice: Dose-dependent, sustained increase in BP characterized by hypokalemia, metabolic alkalosis and suppressed plasma renin activity and aldosterone levels
Yohimbine causes acute, dose-dependent increase in BP.
Diet pills

- Sibutramine
- Phenylpropanolamine
Increased levels of norepinephrine with subsequent activation of noradrenergic transmission Mild increase in BP.

12.3 Optimizing lifestyle changes and ongoing drug therapy

Effective treatment of true resistant hypertension should combine (i) lifestyle changes (particularly reduction of sodium and alcohol intake, implementation of regular physical activity and weight loss in overweight or obese patients) (ii) discontinuation of interfering substances, (iii) rationalization of current treatment and (iv) the sequential addition of antihypertensive drugs to the existing triple therapy.

Replacing current drugs with a more rational and possibly simpler treatment regimen is based on use of combination therapies that are appropriate to a patient‘s age, ethnicity, compelling indications for certain drug classes, comorbidities and risk of drug– drug interactions. Drugs should be used at the maximal tolerated doses and SPCs should be preferred when available to reduce pill burden and improve adherence to treatment [617]. Because volume retention of multifactorial origin is frequent, reducing sodium intake (<2 g/day) or NaCl intake (<5 g/day) and increasing the intensity of diuretic therapy, particularly in older patients, patients of Black African origin or CKD patients, should be implemented. If eGFR is 30 ml/min, BP control may be improved by increasing the dose of the existing Thiazide diuretic or by switching to a possibly more potent and longer acting Thiazide-like diuretic (indapamide or chlorthalidone). If eGFR is <30 ml/min, a loop diuretic (furosemide, bumetanide and torsemide) should replace Thiazide/Thiazide-like diuretics, although even under this circumstance, Thiazides may retain their natriuretic and antihypertensive effects. In the CLICK trial [573], patients with stage 4 CKD (eGFR 15–29 ml/min/1.73 m2) and poorly controlled hypertension, showed an about 10 mmHg 24 h SBP reduction with chlortalidone versus placebo, and the BP-lowering effect was particularly evident in patients already on loop diuretics [736]. Furosemide and bumetanide should be administered twice daily, because of their short duration of action, whereas longer acting agents, such as torsemide, can be administered once daily [737]. The dose or intake frequency of the loop diuretic may be increased in patients with severe CKD and/or albuminuria [737]. Careful monitoring of kidney function, serum electrolyte levels and fluid status is required to detect dehydration, hypokalemia, hyponatremia, hypovolemia or worsening of kidney function. After optimizing the ongoing therapy, a stepwise addition of other antihypertensive drugs should be considered if BP is still not at goal.

12.4 Fourth and subsequent lines of antihypertensive therapy

In patients with true resistant hypertension, the fourth line treatment should include the MRA spironolactone, based on the evidence of its efficacy in the PATHWAY-2 trial [580] as well as in meta-analyses [579], including those in patients with HFrEF (Fig. 14). A secondary analysis of the TOPCAT trial has shown beneficial effects of spironolactone also in patients with HFpEF [738], a condition in which difficult-to-control hypertension is frequent [193]. In the 2018 guidelines, it was recommended that spironolactone (25– 50 mg/day) should be used with caution in patients with an eGFR <45 ml/min and a plasma potassium concentration >4.5 mmol/l, as these were exclusion criteria in the PATHWAY-2 trial. Thus, the efficacy and safety of spironolactone in patients with more advanced CKD or higher potassium levels at baseline have not yet been established. The spironolactone-associated risk of hyperkalemia is greater in patients with CKD, particularly if the drug is added to a treatment regimen that usually already includes an RAS blocker [739], making it necessary to closely monitor plasma potassium and eGFR after treatment initiation and, depending on individual risk and the CKD stage, at least annually or at three to 6 month intervals thereafter. The use of newer potassium binders such as patiromer [740] or sodium zirconium cyclosilicate [741] can reduce the risk of hyperkalemia, without increasing sodium overload (in the case of patiromer) or decreasing antihypertensive drug absorption as observed with sodium polystyrene sulfonate. Not all patients will be able to tolerate spironolactone because of its antiandrogenic side effects resulting in breast tenderness, gynecomastia and sexual impotence in men, and menstrual irregularities in women. The other steroidal MRA, eplerenone, has lesser potential to interfere with progesterone or androgen receptors and can, thus, be used alternatively to lower BP, but it is less potent than spironolactone [742]. Furthermore, in many countries, eplerenone is only approved in patients with HF, and both eplerenone and spironolactone are not approved for use in hypertension in some European countries. A suboptimal tolerability profile as well as restrictions of its use, may partly explain the low prescription rate and the low persistence on treatment of spironolactone in real-life settings. Only 9.0% of patients with apparent resistant hypertension were treated with an MRA in a survey carried out in the USA [711], and only about 30% of patients were prescribed spironolactone at enrolment in the RADIANCE TRIO trial testing the efficacy of RDN in patients with resistant hypertension [743]. Alternative drugs can be amiloride, to be used at high dosages (10– 20 mg per day), which was as effective as spironolactone (25– 50 mg per day) in reducing BP in an open-label extension period of the PATHWAY-2 trial [576]. However, this can lead to an increased pill burden as the marketed dose of amiloride is only 5 mg, and the drug is not available as a single agent but only in combinations (usually 5 mg) in many countries. Finally, new more selective nonsteroidal MRAs such as finerenone (approved for the treatment in diabetic kidney disease), esaxerenone (approved for the treatment of hypertension in Japan), and ocedurenone (KBP-5074, in development for resistant hypertension in CKD) might provide future alternatives to spironolactone for patients with resistant hypertension [742]. Ocedurenone (0.25– 0.50 mg/day) reduced BP in patients with resistant hypertension and stage 3b/4 CKD with a higher incidence of hyperkalemia at the highest dose [744]. Finally, the use of selective aldosterone synthase inhibitors such as baxdrostat has been shown to effectively lower BP in patients with resistant hypertension in a phase 2 trial [717] and may, thus, develop into an additional treatment. This approach will avoid the noxious overall effects of aldosterone by reducing its synthesis instead of blocking its effects on mineralocorticoid receptors. Spironolactone as well as all above discussed alternatives should be used with caution in patients with reduced eGFR and baseline potassium levels >4.5 mmol/l.

F14
FIGURE 14:
BP-lowering strategy in true resistant hypertension according to renal function. (a) When SBP is ≥140mmHg or DBP is ≥90 mmHg provided that: maximum recommended and tolerated doses of a three-drug combination comprising a RAS blocker (either an ACEi or an ARB), a CCB and a T/TLDiuretic were used, adequate BP control has been confirmed by ABPM or by HBPM if ABPM is not feasible, various causes of pseudo-resistant hypertension (especially poor medication adherence) and secondary hypertension have been excluded (see Section 12). (b) Use of Diuretics: Use T/TLDiuretic if eGFR >45 ml/min/1.73 m2. Consider transition to Loop Diuretic if eGFR is between 30 to 45 ml/min/1.73 m2. Use loop Diuretic if eGFR <30 ml/min/1.73 m2. (c) MRA contraindicated. (d) Caution if eGFR <45 ml/min/1.73 m2 or serum potassium >4.5 mmol/l. (e) Should be used earlier at any step as guideline directed medical therapy in respective indications or considered in several other conditions (Table 16).

When spironolactone and other MRAs are not tolerated or contraindicated (i.e. in CKD stage 4, eGFR <30 ml/min), bisoprolol (5– 10 mg/day), doxazosin extended release (4– 8 mg/day) or a centrally acting agent such as the alpha-adrenergic receptor agonists (clonidine, 0.1– 0.3 mg twice a day) [603] can be used as alternatives. However, bisoprolol and doxazosin reduced BP less effectively than spironolactone in the PATHWAY-2 trial [580], while clonidine has shown BP-lowering effects similar to spironolactone in a head-to-head comparison open-label RCT in patients with resistant hypertension [603]. Depending on approval and availability, the dual endothelin antagonist aprocitentan may also be used, because this drug had a sustained BP-lowering effect in patients with resistant hypertension as compared with placebo [718]. However, dose-dependent increase in the incidence of mild-to-moderate edema and fluid retention was also observed and may impair tolerability and safety [718]. Direct vasodilators, such as hydralazine or minoxidil, should be used parsimoniously because they may cause severe fluid retention and reflex sympathetic activation with tachycardia. Recent RCTs have shown that endovascular RDN can be associated with a significant, albeit not marked, office and ambulatory BP reduction in patients with uncontrolled hypertension [743,745,746]. In a large registry of renal denervated patients, the BP reduction was long-lasting and devoid of significant safety problems [191,747]. RDN can thus be proposed as an adjunctive therapy to patients with resistant hypertension provided eGFR >40 ml/min/1.73 m2, in whom BP control cannot be achieved or serious side effects cannot be avoided with antihypertensive medications [743,748] (see Section 13.1).

It is important to emphasize that in resistant hypertension, dedicated trials such as the PATHWAY-2 [580] and ReHOT [603] were short-term efficacy studies showing the effect on BP-lowering after 12 weeks of treatment. Both studies included a relatively small number of patients (314 and 187, respectively) and in a relatively large fraction of patients, BP was still not controlled after treatment with spironolactone (about 40% in PATHWAY-2 based on HBPM measurements and about two-thirds of patients in ReHOT based on office BP) [749]. Finally, and most importantly, outcome RCTs in resistant hypertension are lacking. A recent real-world-evidence study in the United States included a total of 80 598 patients with resistant hypertension and compared the effectiveness of newly prescribed MRA treatment (6626 patients, 98% of whom were spironolactone users) with newly prescribed BB treatment (73 972 patients) as fourth line drugs [750]. In propensity score matched analysis, a 23% nonsignificant (95% CI 0.50– 1.19) reduction in favor of spironolactone for the combined primary outcome of stroke and myocardial infarction was found. The risk of hyperkalemia and worsening of kidney function was significantly greater for spironolactone. An earlier similar and smaller observational study in a UK primary care database involved 8639 patients with resistant hypertension receiving new prescriptions of MRAs (n = 350), BBs (n = 2869) and alpha-1 blockers (n = 5420) [751]. The risk of the primary outcome (combined all-cause mortality, stroke and myocardial infarction) indicated a nonsignificant risk difference in favor of BBs compared with spironolactone and a significant risk difference (-32%) in favor of alpha1-blockers [751]. Taken together. these observational studies show a substantially lower use of MRAs (spironolactone) in clinical practice, while data supporting a benefit for CV outcomes or mortality with use of spironolactone remain lacking. RCTs to identify the most protective medical therapy in true resistant hypertension are needed.

Treatment of true resistant hypertension includes the patients’ frequent comorbidities, for which additional treatment options may apply. Hence, for patients with OSA, continuous CPAP may be of moderate benefit [752,753], especially when this condition is severe (AHI >30 events/h), baseline BP is high and adherence to CPAP is good [754]. In obese patients, GLP1 receptor agonists can reduce body weight [755], modestly lower BP [756,757] and improve CV prognosis in patients with type 2 diabetes or with established CVD [758]. Bariatric surgery can lower BP, CV risk factors and risk of CV events in severely obese patients [759] and may, thus, reduce the burden of antihypertensive medication when these patients have resistant hypertension [760]. In patients eligible for treatment with SGLT2is, their use may add a moderate BP-lowering effect to the background antihypertensive therapy for resistant hypertensive patients [761]. Finally, compared with valsartan alone, the sacubitril/valsartan combination did not lower hospitalization for HF and death in patients with HFpEF [762], but did reduce significantly the NYHA class of the patients. Furthermore, in a post hoc analysis of the same study, its use reduced BP in patients with resistant hypertension, despite treatment with at least four antihypertensive drugs, including an MRA [763]. This confirmed the BP-lowering effect of this compound reported in a phase 2 trial in hypertension [605].

Given the association with multiple comorbidities and the need for multiple and complex drug therapeutic regimens, we recommend to address patients with true resistant hypertension to a hypertension specialist or, if necessary, to a hypertension specialist center. A dedicated tertiary BP clinic can be useful to perform the necessary diagnostic steps, optimize the multidrug treatment regimen, reduce the likelihood of drug-related adverse effects and increase adherence to treatment. Patients should receive a dedicated program of FU.

True resistant hypertension

figure32

13. DEVICE-BASED TREATMENT OF HYPERTENSION

13.1 Renal denervation (RDN)

Increased activity of the SNS (Fig. 3) is one of the important factors in the pathophysiology of hypertension, especially in obesity, OSA and CKD [764]. Efferent sympathetic nerves to the kidneys increase renin release via beta1-adrenergic receptor activation at the level of the juxta-glomerular cells, decrease renal perfusion and GFR, increase tubular reabsorption of sodium and induce a rightward shift of the BP-natriuresis curve. Conversely, increased afferent sensory nerve signaling to the central nervous system in response to kidney ischemia, injury or inflammatory, fibrotic processes and other alterations of the tissue environment leads to reflex sympathetic activation, with peripheral vasoconstriction, increased BP and aggravation of HMOD [765–768]. The rationale of RDN is to modulate the overactive signaling between the kidneys and the central SNS, which may be at least partly responsible for the sympathetic hyperactivity of resistant hypertension [720]. The introduction of endovascular catheter-based RDN devices has allowed to obtain RDN in a minimally invasive fashion [769]. RDN has been shown to reduce whole-body and renal sympathetic activity in humans [770–772], although not in all studies [773]. However, a recent meta-analysis of available studies has reported a limited relationship between the RDN-dependent reduction of sympathetic activity as measured by microneurography and the BP reduction [774]. This is compatible with the possibility that more than just a neural factor is responsible for the RDN-dependent therapeutic effects.

13.1.1 Clinical Evidence of the BP-lowering effect of RDN

Proof-of-concept human studies applying radiofrequency energy, high frequency unfocused ultrasound energy and perivascular injection of alcohol found substantial decreases in BP in patients with uncontrolled, treatment-resistant hypertension [748,769,775]. However, when in 2014, the sham-controlled SYMPLICITY-3 HTN trial failed to show a significant BP reduction after RDN compared to sham treatment, clinical investigations of RDN almost stopped [776]. Lack of significant BP reduction in the SYMPLICITY-3-HTN study was later attributed to energy delivery at the proximal (instead of distal) location of the innervated renal artery, incomplete noncircumferential denervation, high rates of drug changes during the run-in and treatment phase and lack of adherence to drug treatment [777]. As a result, clinical consensus conferences in Europe and the US recommended to (i) use optimized techniques, ensuring complete circumferential ablation of renal nerves, (ii) apply strict criteria for patients inclusion and during the run-in phase and (iii) objectively measure medication adherence individually [778].

To categorize the scientific quality of published RCTs, the following quality criteria were applied: (i) sham-controlled, multicenter trial, (ii) adequate blinding of patients and outcome assessors, (iii) ambulatory BP change as the primary outcome, (iv) study completion as planned and (v) use of advanced RDN systems [748,769]. Of the 18 RCTs already published or ongoing, 9 RCTs fulfilled all these criteria [748,769]. The REQUIRE trial, which included patients with uncontrolled resistant hypertension, failed to fulfill all quality criteria because of incomplete medication blinding of treating physicians and coordinators combined with a lack of objective measurements of medication adherence [779]. Two RCTs with perivascular injection of alcohol are not yet published [748,769]. Thus, at present, an RDN should be based on the procedures that applied RF energy to main, accessory, and distal arteries or high frequency unfocused ultrasound energy to main and accessory arteries.

13.1.2 Off-medications studies

Patients with uncontrolled office and 24 h ABPM in the absence of antihypertensive drugs, with suitable renal artery anatomy, and eGFR 40 to 45 ml/min/1.73 m2 were investigated in four of the nine RCTs. In the SPYRAL HTN-Off MED Proof-of-Concept Study, radiofrequency-based RDN decreased 24 h SBP (primary objective) by 5.0 mmHg (P = 0.041) compared with the sham procedure [780]. In the SPYRAL HTN-Off Med Pivotal Study comprising 331 patients off any antihypertensive medication, 24 h SBP reduction was in favor of radiofrequency-based RDN [group difference − 3.9 mmHg (P < 0.001)] [746]. In the RADIANCE HTN SOLO trial, ultrasound RDN decreased daytime SBP by 6.3 mmHg (P= 0.001) versus a sham procedure [745]. In the pivotal RADIANCE-II trial randomizing 224 patients with uncontrolled BP in the absence of antihypertensive drugs, the daytime SBP reduction (primary objective) was greater with ultrasound RDN than with the sham intervention [between-group difference of 6.3 mmHg (P < 0.001)] [781]. Thus, multiple evidence exists that RDN using radiofrequency and ultrasound energy reduces BP significantly in hypertensive patients in the absence of antihypertensive medication.

13.1.3 On-medications studies

In the SPYRAL HTN-ON Med study in 80 patients with uncontrolled office BP and 24 h ABPM on one to three antihypertensive drugs, radiofrequency-RDN was associated with a greater BP reduction than sham procedure at 6 months (between-group difference of 7.0 mmHg, P= 0.0059), while antihypertensive therapy was unchanged [782]. In the RADIANCE HTN-TRIO Study, 136 patients with uncontrolled hypertension resistant to more than three antihypertensive medications were switched to a single pill, triple antihypertensive drug combination during 4 weeks run-in and prior to the RDN intervention. Ultrasound RDN reduced daytime SBP to a greater extent than the sham procedure at 2 months (between-group difference of 4.5 mmHg, P= 0.022) [743]. The triple therapy remained unchanged and adherence to treatment was stable (about 80% in both groups).

Published meta-analyses including RCTs of high, medium and even low scientific quality showed variable results, albeit data from RCTs showed consistent 24 h BP reduction ranging from 3.9 to 7.0 mmHg SBP and from -3.7 to -6.9 mmHg DBP after RDN. Remaining gaps in knowledge concern a more precise definition of the prevailing magnitude of BP reduction after RDN, an issue that might be fruitfully addressed by individual patient-level meta-analysis [769], although pooled analysis of trials conducted by ultrasound RDN showed a rather consistent BP-lowering effect [783]. Head-to-head comparisons of RDN to properly conducted intensified pharmacotherapy studies are also needed. The Prague-15 study, reporting similar effects between RDN and optimized pharmacotherapy mainly by adding spironolactone, was inconclusive, since after 6 months, 25% of the RDN group were also prescribed spironolactone, and in the pharmacotherapy group, only 61% were still on spironolactone [784]. Finally, no solid predictor of future BP reduction after RDN has yet been identified, with the exception of pretreatment BP [748,769]. The latter finding is not specific for RDN, as it has been almost invariably observed in virtually all trials on the BP-lowering effect of antihypertensive drugs. Nevertheless, the lack of a diagnostic measure that predicts the BP response to a device-based treatment such as RDN represents a relevant limitation.

13.1.4 Safety

The main concerns regarding the safety of RDN is potential damage to the arterial endothelium, intima and media by the applied energy, renal artery dissection, contrast-induced nephropathy in the short-term, occurrence of de-novo renal artery stenosis and eGFR loss in the long-term [748,769]. No safety signals emerged in any of the sham-controlled RCTs in which there were similar rates of major adverse events in the RDN and control groups [743,745,746,778,780–782,785]. This conclusion is limited to the studied population with an eGFR >40 ml/min/1.73 m2. A meta-analysis of 50 trials comprising 10 249 patient-years of data estimated a 0.20 annual incidence of renal artery stenting following RDN, which is similar to the reported natural incidence of renal artery stenosis in hypertension [786]. RCTs systematically using noninvasive renal artery imaging 1 year after the procedure confirmed vascular safety of RDN. Likewise, a meta-analysis of 48 studies with 2381 patients showed no significant change in eGFR after a FU of 9.1 months [787]. No acute kidney injury or time-dependent decrease in kidney function was reported. In the Global Simplicity Registry, the observed eGFR decrease being within the expected time-dependent eGFR decline, with the limitation that so far, the FU is limited to 3 years [747]. Since with currently available RDN devices, femoral arterial access is needed, access-site vascular complications, e.g. hematoma, pseudoaneurysm, fistula, bleeding among others may occur [748,769].

13.1.5 Durability

Long-term data are available from sham-controlled RCTs, although unblinding of both patients and physicians generally took place after 6 and 12 months, respectively, and antihypertensive medications were added, changed or stopped by primary care physicians. Based on the analysis of patients enrolled in the Global Simplicity Registry (n= 2652 patients) [747], the BP decrease documented after RDN appears to be clearly maintained up to 36 months after RDN. Similarly, in the RADIANCE HTN-SOLO FU, the RDN-related BP reduction was found to be maintained over a FU of 3 years although, because of the high rate of cross-over to RDN, no formal comparison with the sham group was provided [788]. Similar limitations involve the 3-year durability of the BP effects of RDN reported by the SPYRAL HTN-ON MED trial [789] as well as the long-term increase of the antihypertensive effect of RDN reported by the SYMPLICITY HTN-3 trial [790,791], because of the different magnitude of the long-term BP differences, depending on inclusion or exclusion of patients who crossed over to RDN in the comparison group [792]. It should be mentioned that the durability of the BP reduction associated with RDN argues against a functional reinnervation of the kidneys in hypertensive patients after RDN, even though nerve regrowth has been reported in experimental animal models. However, the functionality of regrown nerves is uncertain in these models and long-term human data yet need to be explored. Notably, in patients after renal transplantation no clinically meaningful reinnervation was observed in long-term studies.

13.1.6 Application

Endovascular RDN with radiofrequency energy or high frequency unfocused ultrasound energy represents a treatment option, that is additive or alternative to increasing medication in patients with uncontrolled resistant hypertension confirmed by ABPM after excluding secondary causes of hypertension [748,769]. Patients who are repeatedly nonadherent (if this reflects the unwillingness of the patient to take drugs) or intolerant to multiple antihypertensive drugs, may be considered for RDN after information about the potential lack of effect and benefits, and also the risks associated with the procedure [748,769]. These patients may be on fewer than three drugs at the time of their selection for RDN. To date, no prospective multicenter, blinded, randomized, prospective outcome trial exists for RDN. Unfortunately, this also applies to all drug treatment strategies for true resistant hypertension. This means that whether, and to what extent, BP reduction in resistant hypertensive patients translates into CV protection is unknown, and the value of BP reduction as a protective marker is necessarily extrapolated from the large body of evidence obtained by RCTs in nonresistant hypertensive individuals. Based on the outcome reductions calculated for SBP reductions in a large meta-analysis of RCTs [80,793] the Global Simplicity Registry concluded that RDN might reduce the relative risk of stroke by 43%, while the absolute risk of major adverse CV events might decline from 11.7% in the control group to 8.6% in the RDN group [793]. Likewise, in 3077 patients of the Registry, the conclusion was reached that a 10% increase of time at BP target (120– 140 mmHg SBP) during the first 6 months after RDN might reduce by 15% the risk of major CV events in the following 6–36-month period [191]. As mentioned above, these conclusions suffer from the limitations that the BP reduction– outcome relationship reported in the meta-analysis did not concern patients with resistant hypertension [80].

RDN should be performed only in experienced and specialized centers that have established a multidisciplinary team with a structured pathway for evaluating hypertensive patients [748,769]. Understanding the patients’ perspective, exploring their preference and expectation is crucial prior to RDN. Benefits and risks of RDN need to be addressed in a shared-decision making process [748,769]. In this regard, roughly 1/3 of hypertensive patients were prone to prefer RDN instead of pharmacotherapy to have the elevated BP controlled [794]. This applied particularly to younger patients, male patients, those with experience of side effects, and those who admitted to be nonadherent [794].

Use of renal denervation

figure33

13.2 Carotid baroreceptor stimulation

Stretch-sensitive baroreceptors located in the carotid sinus and the aortic arch are involved in short-term and long-term BP regulation. Carotid baroreceptor external stimulation via a pacemaker-like device or baroreflex neuromodulation by endovascular deployment of a self-expanding nitinol implant in the carotid artery has been investigated for treatment of resistant hypertension [795]. Carotid baroreceptor stimulation was associated with a reduction of sympathetic nerve activity in studies on hypertensive patients [796,797], and a sympathoinhibitory effect was also shown when the stimulation was applied to HF patients [797]. The first-generation bilateral electrical stimulation device (Rheos, CVRx) was tested in a double-blind, randomized, sham-controlled pivotal trial, which included 265 patients with resistant hypertension [798]. At 6 months, the office BP fall was significantly larger in the treatment group compared with the sham group. However, the study failed to meet two of the five co-primary endpoints, and safety was not established. Therefore, the Rheos device did not receive approval from the Food and Drug Administration for use in patients with resistant hypertension [798]. A second-generation unilateral stimulation device has been developed to reduce the complexity, complications and costs of the procedure (Barostim). However, no RCT is currently available with this new device in patients with resistant hypertension. The endovascular baroreflex amplification therapy is achieved via implantation of a dedicated stent, which aims at passively increasing wall stretch by increasing the vessel-effective radius while preserving pulsatility [799]. In a small, noncontrolled, open-label, first-in-human CALM-FIM study, 30 patients underwent implantation of the MobiusHD system (Vascular Dynamics). At 6 months, there were significant reductions in both office and ambulatory BP compared with baseline, which appeared to be maintained through 36 months [800]. Several RCTs investigating this approach are ongoing.

13.3. Other device-based treatments

The creation of a fixed-diameter iliac arteriovenous anastomosis with a catheter-based device (ROX coupler; ROX Medical) was investigated in resistant hypertension to lower peripheral vascular resistance [795]. The creation of such a shunt significantly decreased BP in the prospective, open-label, randomized, controlled ROX CONTROL HTN trial [801]. However, 33% of all patients undergoing implantation of the arteriovenous coupler developed late ipsilateral venous stenosis requiring treatment. Because of a potential HF risk following treatment with the coupler, the development of this device has been stopped.

The Moderato system (BackBeat Cardiac Neuromodulation Therapy, Orchestra BioMed) is a dual-chamber, rate-responsive, implantable pulse generator that variably shortens and lengthens the atrioventricular interval [795]. By shortening the atrioventricular coupling interval, left ventricular filling can be reduced, and BP falls. The device intermittently and asymmetrically introduces short sequences of one to three beats of longer atrioventricular delay with the aim of preventing a compensatory baroreflex-mediated activation of the SNS [802]. Following the initial proof-of-concept study (MODERATO I) [802], which included 35 patients with uncontrolled hypertension despite 2 antihypertensive medications, the results of the MODERATO II study [803], a prospective, multicenter, randomized, double-blind study were reported. In this trial, 68 patients with uncontrolled hypertension despite treatment with at least one antihypertensive medication and an indication for the implantation or replacement of a dual-chamber pacemaker underwent the Moderato device implantation. In the treatment group, 24 h SBP immediately dropped after the device activation and the BP-lowering effect was maintained through 6 months of FU. Although the primary efficacy endpoint (difference of 24 h SBP change between groups) was met, and treatment appeared to be safe through 6 months, the long-term consequences of cardiac neuromodulation therapy need to be investigated in larger trials. Device-based therapies such as carotid baroreceptor stimulation, arteriovenous anastomoses and cardiac neuromodulation therapy (via a dedicated pacemaker) are not recommended for antihypertensive treatment.

14. SPECIFIC HYPERTENSION PHENOTYPES

14.1 Sustained hypertension and true normotension

Use of out-of-office BP measurements by HBPM and/or ABPM allows identifying BP phenotypes that were unknown when BP measurements were limited to office BP. One phenotype is termed sustained hypertension and consists of an elevation of both office and out-of-office BP. Another is true normotension, which is characterized by office and out-of-office BP normality. These phenotypes may refer not only to untreated but also to treated individuals, where they indicate extended office and out-of-office BP control or no control of all these BP values. Available studies have defined out-office BP normality or elevation by either HBPM or ABPM and, at present, only the PAMELA study (a study based on a population sample from northern Italy) provides office BP, ABPM and HBPM in each individual, although with a restricted protocol for home BP measurements [142]. This study allowed defining the sustained hypertension and true normotension phenotypes based on all three BP values and found that, compared to true normotension, sustained hypertension is associated with a clearcut increase in the prevalence and incidence of CV mortality. Outcome differences have also been reported between patients with one versus two or three elevated BP values [159], suggesting a clinical relevance for even more complex BP-based phenotypes.

14.2 White-coat hypertension

WCH refers to the untreated condition in which BP is elevated in the office but is normal when measured by ABPM, HBPM or both [4]. The term white-coat effect is used to describe the difference between an elevated office BP and a lower home or ambulatory BP, which is believed to mainly reflect the pressor response to an alerting reaction elicited by office BP measurements by a physician or a nurse [804,805]. However, other factors are probably also involved, as shown by the poor correlation between the office and out-of-office BP difference and the white-coat effect measured directly with beat-to-beat BP recording [806,807]. Although the prevalence varies between studies, WCH can account for about 30% of people attending hypertension clinics [62] and up to 30– 40% among patients with an elevated office BP. It is more common with increasing age (>50% in the very old patients), in women and in nonsmokers. Its prevalence is lower when office BP is based on repeated measurements or when the attending physician or nurse are not involved in the BP measurement [95]. A significant white-coat effect can be seen at all grades of hypertension (including true resistant hypertension), while the prevalence of WCH is greatest in grade 1 hypertension.

There has been much debate in the literature about whether WCH should be considered an innocent condition. HMOD is less prevalent in WCH than in sustained hypertension, and several studies have shown that this is the case also for the risk of CV events [62]. However, compared with true normotensives, patients with WCH have increased adrenergic activity [808], a greater prevalence of metabolic risk factors and a more frequent asymptomatic HMOD [809]. In the PAMELA population, cardiac and renal asymptomatic HMOD was detected in about one of three patients with WCH compared to one of ten individuals with normal office and out-of-office BP [217]. Furthermore, compared to normotensive individuals, white-coat hypertensive individuals have shown a greater long-term risk of new-onset diabetes, progression to sustained hypertension and CV mortality [217,810–814]. The increased CV risk and mortality have been reported with (i) diagnostic use of both HBPM and ABPM, (ii) in the absence of HMOD at baseline and (iii) in ISH, older patients and other conditions [815–818]. In addition to the above-mentioned factors (dysmetabolic risk profile, more common HMOD and increased risk of new-onset diabetes), this can probably also be accounted for by a greater prevalence of nocturnal hypertension and by 24 h BP values that, albeit normal by definition, are a few mmHg higher than in nonwhite-coat hypertensive individuals [814]. Due to its limited reproducibility, the diagnosis of WCH should be confirmed by repeated office and out-of-office BP measurements. Ideally, out-of-office BP measurements should include both ABPM and HBPM because the two values can give discrepant results, i.e. one value can be normal and the other elevated or vice versa, and the CV risk appears to be lower (and close to sustained normotension) in white-coat hypertensive individuals in whom ABPM and HBPM are both normal [814]. Thorough assessment of CV risk factors and HMOD are recommended. Treatment should consider lifestyle changes to reduce CV risk and a closer FU compared with true normotensive individuals. Antihypertensive drug treatment effectively lowers office BP, while the effect on out-of-office BP is small and variable [819,820]. Whether or not patients with WCH should receive antihypertensive drugs is still unresolved, because, although WCH patients have been a considerable proportion of virtually all RCTs proving the benefits of antihypertensive treatment [821], no specific outcome-based trial has been performed. Drug treatment may be considered in patients with HMOD and a high CV risk.

White-coat hypertension (WCH)

figure34

14.3 Masked hypertension

MH refers to untreated patients in whom the BP is normal in the office but elevated when measured by HBPM or ABPM [822]. About 10– 20% of patients attending hypertension clinics have MH [62], with out-of-clinic BP measured either with ABPM or with HBPM. A noticeable prevalence has been found in population-based studies, especially in Asian and African American patients [822,823]. The prevalence of MH also varies when different ABPM periods (daytime, 24 h or night-time) are used to define the out-of-office hypertension status. The optimal approach for the detection of MH has not been established. Screening all individuals with nonelevated office BP for MH is impractical. An office BP in the high-normal BP range is associated with a higher likelihood of MH. The prevalence is greater in younger people, men, smokers and those with higher levels of physical activity, alcohol consumption, anxiety and job stress [824,825]. Obesity, diabetes, low HDL-cholesterol, CKD, family history of hypertension, are also associated with an increased prevalence of MH [824]. An exaggerated BP response to exercise and to the orthostatic posture have also been found to be predictors of MH [826]. A CV risk-based approach, limiting the use of out-of-clinic BP measurement to those individuals with multiple risk factors for MH, has been proposed [822]. MH has been associated with HMOD such as impaired kidney function, LVH, carotid intima– media thickness and large artery stiffness [815,826–828]. People with MH have an increased risk of developing metabolic abnormalities and diabetes as well as sustained hypertension [812,829] and have increased sympathetic activity [830,831]. Meta-analyses and recent studies have shown that the risk of CV events is substantially greater in MH compared with normotension, and intermediate risk or even close to the risk of sustained hypertension [822,832–834]. Both ABPM-based and HBPM-based MH has been found to be independently associated with CV events and mortality [822,832,833]. An increase in risk of CV and kidney events has also been observed in diabetic patients with MH, especially when the BP elevation occurs during the night [835]. A systematic review and meta-analysis has shown a slight to fair reproducibility of MH that was better for ABPM than for HBPM [836], while a limited reproducibility has been reported in other studies [62]. Thus, the diagnosis of MH requires confirmation with at least a second set of office and out-of-office BP measurements. No RCT has ever been performed on MH, which means that the effects of antihypertensive treatment are unknown. Given the adverse prognostic importance of out-of-office BP elevations, it seems appropriate to recommend that in patients with confirmed MH, stringent lifestyle modifications and a closer FU are implemented. Antihypertensive drug treatment may be considered if CV risk is particularly elevated, and in patients with HMOD.

Masked hypertension (MH)

figure35

14.4 White-coat uncontrolled hypertension (WUCH) and masked uncontrolled hypertension (MUCH)

Originally referred to untreated patients, WCH and MH now also include patients on antihypertensive treatment, defined as (i) WUCH (white-coat uncontrolled hypertension), i.e. BP control by treatment during the 24 h but not in the office and MUCH (masked uncontrolled hypertension), i.e. BP control by treatment in the office but not outside the office. Although patients with WUCH have shown greater large artery stiffness than those with controlled BP [837], several studies have shown that their CV risk does not differ significantly from that of the treated population in which both office and out-of-office BP achieved control. Consistent results were provided by two large studies in which WUCH was assessed with HBPM [133,838]. In the IDACO study, analyses were stratified by the presence or absence of antihypertensive therapy and found that WCH was associated with an increased CV risk only in untreated patients [133]. Data provided by large meta-analyses have confirmed that patients with WUCH do not have an increased CV risk compared to those in whom both office and out-of-office BP are controlled [811]. Patients taking antihypertensive medication have been found to have a higher prevalence of MUCH compared with untreated individuals. In the large IDACO database, the prevalence of MH (ABPM elevation) was about 1.7 times higher for individuals taking versus those not taking antihypertensive medications (31.9 versus 19.2%)[133]. Similar results have been obtained in African Americans by the Jackson Heart Study [839], while in participants with controlled office BP of the SPRINT study, MUCH was present in 62 and 56% in the intensive and standard groups, respectively [840]. MUCH has been found to be associated with a worse metabolic profile [841], HMOD [842] and unfavorable clinical outcomes, regardless of the out-of-office BP monitoring technique [832,833]. In a meta-analysis of observational studies, MUCH patients had rates of CV events higher than those in patients with both office and out-of-office BP control, and similar to those of patients with treated uncontrolled office and out-of-office hypertension [843], regardless whether ABPM or HBPM was used for MUCH identification. A limitation of all available studies is that MUCH and WUCH identification was based on a single set of office and out-of-office BP measurements, because post hoc analyses of clinical trials have shown that both phenotypes are extremely inconsistent, irrespective of the type of antihypertensive treatment [172,844]. In the analysis of the ELSA data, only about 5% of the WUCH or MUCH patients exhibited the same condition throughout the 4 years of the trial [172]. Thus, occasional rather than consistent phenotypes may have been addressed in outcome analyses. As observed for MH, MUCH is more frequent with smoking habits, alcohol consumption, overweight, BP response to physical activity, psychological stress and some comorbidities [824,845]. The role of medication nonadherence is equivocal [846]. There is evidence that patients with MUCH have higher level of sympathetic activity assessed in daily life conditions than those in whom both office and out-of-office BPs are controlled by treatment [845,847]. This may be responsible at least in part for the higher CV risk found in MUCH. It is recommended to identify WUCH and MUCH with repeated office and out-of-office BP measurements. Considering the limitations of the available evidence, it seems wise to advise treatment uptitration in both WUCH and MUCH if treatment is well tolerated, so as not to keep the patient above the recommended target BP values generally recommended for antihypertensive treatment, without solid supporting evidence.

White-coat uncontrolled hypertension (WUCH) and masked uncontrolled hypertension (MUCH)

figure36

14.5 Isolated systolic hypertension of the young

ISHY, defined as an SBP 140 mmHg and a DBP <90 mmHg, is present not only in older persons but also in young and very young individuals (ISHY), more commonly in male individuals [121]. ISHY may be present also in children and adolescents and is often associated with overweight and obesity. There is debate in the literature about the clinical significance of ISHY, i.e. whether this condition associates with worse outcomes and needs antihypertensive treatment. The reason is due to the different pathogenetic backgrounds of this BP phenotype. Mechanistic research has documented that in these people, an isolated SBP elevation may be associated with or caused by multiple factors that can operate in isolation or interact. Increased cardiac output, heart rate and stroke volume are the predominant hemodynamic abnormalities in ISHY [121,848] and may explain why peripheral pulse pressure is higher and ISHY is more common in athletes than sedentary people [849]. However, some studies showed that about 20% of ISHY patients had a normal stroke volume and an increased pulse wave velocity, indicating that ISHY may be associated with premature aortic stiffening [848]. In the NHANES [850] as well as in other studies [121], obesity, male sex, high salt intake and smoking, were associated with higher odds of ISHY. In contrast, some investigations have identified subgroups (more commonly tall men, nonsmokers and people active in sports), in which ISHY was associated with no risk factors, and there was coexistence of pulse pressure amplification with a normal central BP. This low-risk condition has been termed spurious systolic hypertension [851,852]. In a large multiethnic population, individuals with ISHY had central SBP and pulse wave velocity values lower than in individuals with isolated diastolic or systolic– diastolic hypertension and similar to those of individuals with a high-normal BP [853]. These conflicting data suggest that ISHY is a very heterogeneous condition that may include individuals with very different genetic and pathophysiological backgrounds as well as clinical characteristics. Central BP measurement can help to identify ISHY patients at lower risk [121,851,852] and a good prognosis, as shown in longitudinal studies with long-term FU [854,855]. However, different results were reported in two more recent studies in which ISHY has been associated with increased CV risk, limited to men in one study [856] and to women in two other reports [855,857]. It should be noted that between-sex comparison is difficult in ISHY, because all studies have shown a clear predominance of this phenotype and a higher event rate in men. Clinical evaluation of an individual with ISHY should consider the possible presence of WCH because one of the strongest determinants of high pulse pressure in this condition is a pronounced white-coat effect [858]. This means that all individuals with ISHY should be assessed with out-of-office BP measurement. If ISHY is confirmed, assessment of central BP, other central hemodynamic parameters and arterial distensibility may provide additional useful information, although with the limitation that central BP lacks documented cut-off values that differentiate normal from high values (see Section 4.6). Other problems are the current limitation of available prognostic data and the uncertain prognostic superiority of central versus brachial BP in the younger population. Young individuals with ISHY should receive recommendations on lifestyle modifications, particularly cessation of smoking, sodium (NaCl) restriction and hypocaloric diet in the presence of overweight [121]. In addition, they require a close FU because those with high central BP are prone to develop sustained systolic– diastolic hypertension over time [859]. In individuals who present with high out-of-office BP, high central BP and other risk factors, pharmacological treatment should be considered.

Isolated systolic hypertension in the young (ISHY)

figure37

14.6 Isolated systolic hypertension in older persons

In older people, several functional and structural abnormalities including endothelial dysfunction, vascular remodeling and fibrosis cause an increase in stiffness of large elastic arteries. The increased size of the forward pressure wave together with the earlier arterial wave reflection from the peripheral to the large arteries lead to an increase of SBP and pulse pressure [860]. As a consequence, aging is accompanied by a steady increase in SBP while a plateau of DBP occurs at the age of 50– 60 years, followed by a decrease [861,862]. The increased arterial load due to the high SBP promotes vascular atherosclerosis and LVH, ultimately leading to CAD, cerebrovascular disease and HF. ISH, defined as SBP ≥ 140 mmHg and DBP <90 mmHg, becomes the most common form of hypertension after 50 years of age and is present in most patients with hypertension who are >70 years of age [863]. ISH is also more common in women and overweight people. A large body of evidence has shown that SBP has a greater impact on outcomes than DBP after 50 years of age [864,865] and that ISH, either assessed with office BP [860,866] or by ABPM [867], is associated with a high risk of CV outcomes and mortality. The risk of CV outcomes and mortality is increased also in patients with grade I ISH, i.e. with SBP values between 140 and 159 mmHg [868] and is greater in the ISH subgroup with orthostatic hypertension [869].

Diagnosis of ISH is particularly challenging in old individuals because of the high BP variability that characterizes this condition and the frequent occurrence of a pronounced white-coat effect at office BP measurement [860]. Thus, BP assessment with repeated office visits or out-of-office BP measurement is recommended. The use of central BP measurement may also be of help because it allows the identification of those ISH individuals in whom the aortic SBP elevation is much less pronounced than that seen with peripheral BP measurement. RCTs have demonstrated the benefit of treating ISH [498,635,870] even in the oldest segment of the population [502]. An individual-patient meta-analysis of older patients with ISH showed that active treatment reduced all-cause mortality by 13%, CV mortality by 18%, and all CV outcomes by 26% [871]. This meta-analysis also showed that DBP was inversely associated with total mortality, highlighting the role of pulse pressure as a risk factor. The benefit of treatment was larger in men, in patients aged ≥ 70 years and in those with previous CV complications. The above studies also showed that early versus late initiation of treatment after the detection of ISH led to a persistently greater CV protection at all ages. Based on the data in aggregate, CCBs and Thiazide-like diuretics emerged as the drugs of choice for the management of ISH, whereas ACEis/ARBs showed less efficacy, suggesting that they should be used as first-line agents when there are compelling indications such as HF, coronary artery disease, CKD, metabolic syndrome and diabetes [860,866]. Because the rate of BP control with monotherapy is low in patients with ISH, the general recommendation to start with dual combination therapy applies also to older patients with ISH, if they are not frail.

Target SBP and DBP values in older patients with hypertension have been an issue of intense debate. In dedicated RCTs, documenting the protective effect of antihypertensive treatment in ISH, on-treatment SBP remained in the 140–150 mmHg range [145,497,498]. This has been confirmed by a recent RCT meta-analysis [144], which supports the primary recommendation to have, as target SBP, values between 150 and 140 mmHg. However, in a large number of trials on older patients in whom treatment reduced CV outcomes by lowering SBP below 140 mmHg, the number of patients with ISH or a prevalent SBP elevation was considerable [499]. Furthermore, in a recent meta-analysis that included 23 RCTs in each of which a mean baseline SBP was 140 mmHg and mean baseline DBP < 90 mmHg, antihypertensive treatment that lowered SBP to <140 mmHg was associated with a significant outcome reduction [500]. This justifies the additional recommendation to try to reduce SBP in the 140–130 mmHg range, provided that this is well tolerated and DBP is not too low (see below). SBP values <140 mmHg have been found to be safe in the ISH patients of the VALISH trial [872], while in dedicated ISH trials, little information has been provided on the relationship between on-treatment DBP and outcomes.

Nevertheless, in ISH patients, a marked decrease in DBP should be avoided, because the risk reduction obtained by lowering SBP may be jeopardized by the increased risk arising from an excessive decrease of DBP, as recently suggested by the SPRINT study [873]. Intensified BP treatment may be particularly harmful in patients with severe stenosis of large arteries.

To prevent organ hypoperfusion, DBP should not be reduced below 70 mmHg by drug treatment, although compliance with this recommendation is often difficult as a large number of ISH patients already have a DBP <70 mmHg [874], and a considerable number exhibits values within the 70– 80 mmHg range. In population studies largely based on untreated patients, a very low DBP in ISH patients has been associated with a very high prevalence of CVD [874], although the causative factor may not only be poor vital organ perfusion but also a marked arterial stiffness, of which the very high pulse pressure value is a reflection. Treating ISH with low DBP remains a challenging task because of the difficulty to decrease SBP without reducing DBP at the same time, and to apply rigid safety diastolic boundaries, which may limit the achievement of SBP control, with its proven protective effect. Physicians should thus aim at achieving a balance between the best achievable SBP reduction with DBP values that do not raise suspicion of reduced organ perfusion and affect treatment tolerability. In older patients, ISH is a condition in which antihypertensive treatment individualization, based on patient characteristics, risk profile, and level of both SBP and DBP, is particularly important. However, aiming at SBP control remains the primary goal to improve outcome and, if well tolerated, this goal should be pursued also in patients with a low DBP. In the SHEP trial in patients with ISH, a treatment-induced reduction of SBP was accompanied by a clear reduction of major CV events, despite a DBP reduction that brought its average value to 68 mmHg [497].

14.7 Isolated diastolic hypertension

Isolated diastolic hypertension (IDH) is a hypertension phenotype characterized by an SBP <140 mmHg and a DBP >90 mmHg. In the general adult population, IDH prevalence has been reported to be between 2.5 and 7.8% [875,876], with a peak between 30 and 39 years, a decrease in the fifth and sixth decades and almost no case above 70 years of age [877]. IDH has a greater prevalence in men than in women [875]. Some reports have shown that awareness and treatment of hypertension are very low among IDH patients [875,878]. In the PEACE study, only 10.3% of untreated patients knew that they had IDH, and 86.1% of patients with IDH did not receive treatment [875]. IDH is more frequent among people with overweight and obesity, particularly central obesity, and is linked with other components of the metabolic syndrome [875]. When compared with other hypertension phenotypes, IDH patients are generally younger, of male sex, consume more alcohol and tobacco, and are more frequently diabetic patients [875,879]. Some longitudinal studies have shown that IDH patients have higher odds than normotensive individuals of developing systodiastolic hypertension [880,881], and large longitudinal studies on Asian, American or European people have shown that over long FU (up to 31 years), IDH is associated with a greater CV risk compared with normotensive individuals [882–884]. Longitudinal studies have almost invariably reported that the IDH-related risk is age-dependent, i.e. it is seen in patients aged <60 or even <50 years but not above these ages. This has been observed by a cohort-based study on 107 599 patients, which showed that while the relationship between CV events and SBP was age-independent, the relationship between CV events and DBP was significant only in patients <50 years of age [885]. Confirmation is provided by a meta-analysis of 15 cohort studies including 489 814 participants [886] as well as by the results of the IDACO study in which diagnosis of IDH was made by ABPM [887]. In conclusion, available data suggest that IDH is a CV risk factor mainly among younger adults. Unfortunately, this does not satisfactorily clarify whether younger adults with IDH need antihypertensive treatment because, although in early antihypertensive treatment trials enrolment of hypertensive patients was mainly based on DBP levels, treatment reduced both DBP and SBP, and no differentiation was possible between the effects of one versus the other BP reduction [888]. Thus, there is no evidence on the protective effect of antihypertensive medications in the IDH phenotype. At any rate, based on the available epidemiological data and the high risk of transition of IDH to ISH or systodiastolic hypertension, periodic BP evaluations and initiation of lifestyle interventions should be implemented in all IHD patients. Because the evidence of an association of IDH with adverse CV outcomes is more evident in younger than in older patients antihypertensive drug treatment may be considered in patients aged <50 years. In contrast, in older patients with IDH, the low IDH prevalence and the uncertain association with CV events may support treatment limitations to lifestyle interventions and close FU. Antihypertensive treatment may be generally considered in patients with high CV risk.

Isolated diastolic hypertension (IDH)

figure38

14.8 Night-time hypertension and dipping

Night-time hypertension is defined as an average BP ≥ 120/70 mmHg recorded during the night hours with ABPM. In recent years, dedicated home BP devices have emerged as a new method for obtaining BP values during sleep, and they could be a practical alternative to ABPM [889,890]. Evidence in favor of this approach has been obtained with either upper arm or wrist devices measuring BP three times during sleep. Despite the limited number of values, there is evidence that there may be a good agreement between BP measuredat night with ABPM and home devices (mean SBP and DBP differences +1.4 and −0.2 mmHg, respectively) [150]. In several original studies and meta-analyses, night-time BP proved to be more predictive of adverse outcomes, including CV events and mortality, than daytime [891] or even 24 h BP [892]. Patients with nocturnal hypertension are more likely to develop cardiac and carotid structural changes than people with nocturnal normotension [893]. Two large studies have shown that having nocturnal hypertension in people with normal office BP and normal daytime BP, a condition called isolated nocturnal hypertension, is associated with higher risk of HMOD and adverse outcomes [890,893,894]. Isolated nocturnal hypertension has been found to be present in 9.2–12.9% of adults [895,896] and to be more prevalent in men with high-normal BP and a high CV risk profile, African Americans, older individuals, obese and diabetic patients, and patients with CKD [897]. A steady increase in asleep BP and a decline in the nocturnal BP fall occur with aging [898]. An increase of CV risk occurs not only with nocturnal hypertension but also with a reduced magnitude of the nocturnal BP fall, regardless the absolute night-time BP values. The night-to-day ratio is a significant predictor of outcome, and allows subdivision of patients into dippers (night-day ratio ≤ 0.9 or ≥ 10% of the daytime average BP) and nondippers (night–day ratio ≥ 0.9 or ≤ 10%), with the latter group exhibiting an increased CV risk [62]. Moreover, in patients in whom there is no night-time BP reduction or even a higher night-time than daytime average BP (reverse dippers), the increase in risk is particularly marked [62,159,898]. In patients aged 70 years or older, a U-shaped relationship between the night-to-day ratio and CV has been observed [898]. In this age group, not only nondipping or reverse-dipping but also extreme-dipping (>20% nocturnal BP fall) were found to be associated with increased risk of CV events. In contrast, in younger patients, the nocturnal BP reduction/adverse outcome association appeared to be linear. It is important to know that reproducibility of night-time BP patterns is low in both untreated and treated hypertensive patients [173,899], which suggests that night-time phenotype typification should be based on repeated ABPM readings rather than only on single recordings. Several different mechanisms can lead to nocturnal hypertension, including increased SNS activity, autonomic dysfunction, impaired baroreflex sensitivity, salt sensitivity, increased plasma volume, RAS hyperactivity, OSA and other sleep disturbances, increased stress and renal dysfunction [732,900]. Selective treatment of nocturnal hypertension is not available, and no solid evidence exists on the effects of enhancing nocturnal BP reduction on outcomes in nondippers or reverse-dippers. Several therapeutic strategies have been proposed to achieve this goal, in particular, the chronotherapeutic approach with bedtime administration of antihypertensive medications, salt restriction, treatment of sleep disturbances (e.g. sleep apnea), RDN [732,900] and specific drug classes. However, criticism has been raised about thequality of some of thesestudies and attention has been directed to possibleinconveniencesof a nightime BP that is too low, as well as to the possibility of a lower adherence to treatment of bedtime drug administration rather than the more usual morning drug administration (which is always adopted in trials) [671,672]. In a recent, large, prospective, open-label and blinded-endpoint RCT, bedtime drug administration did not lead to any difference in CV outcomes compared with morning drug administration (which was associated with a better adherence to treatment), although with bedtime administration, there was a greater morning and, thus, possibly a greater night-time BP reduction [670]. Based on available evidence, morning or evening intakeof theprescribedmonotherapyor SPC of drugs shouldbeleft to thepatient preference. In thenot infrequent case of a treatment based on multiple drug tablets, patients may prefer to split drug intake into morning and bedtime doses, which may favor a smoother BP profile over the 24 h [901], but some drugs, i.e. diuretics, are not suitable for bedtime dosing.

Night-time hypertension and BP phenotypes

figure39

14.9 Orthostatic hypertension and hypotension

BP should also be measured in the standing position at the first visit, and regularly at each visit in patients who are older, under antihypertensive treatment, have diabetes, or with specific causes or factors that may favor an orthostatic BP fall. At least two BP measurements should be taken 1 and 3 min after standing. Orthostatic hypotension is defined as a decline in SBP of at least 20 mmHg or in DBP of at least 10 mmHg within 3 min of standing, and is associated with an increased risk of mortality and CV events events [902,903] as well as with a greater incidence of marked BP reductions, with, thus, a greater risk of injurious falls, particularly, when patients are treated with vasodilators. However, it is important to remember that an exaggerated BP increase or a BP elevation on standing is also associated with an increased risk of adverse outcomes both in young [831,902,904] and older [869,905] individuals. An important role for the genesis of this condition can be played by unhealthy lifestyle behaviors, including smoking, heavy coffee drinking and alcohol intake [831]. Measuring the BP response to standing may facilitate the identification of people with MH [831]. A neurohumoral overshoot seems to be the driving mechanism of orthostatic hypertension in young adults whereas vascular stiffness seems to be the main pathogenetic factor in older patients [906]. There is no generally accepted definition of orthostatic hypertension in the literature. According to some authorities, an exaggerated pressor response is a sustained SBP increase of at least 20 mmHg when changing from the supine to the standing position. Another definition is an exaggerated pressor response associated with an SBP of at least 140 mmHg while standing [779].

14.10 Baroreflex failure and efferent autonomic failure

Arterial baroreceptors sense changes in vascular stretch resulting from BP fluctuations. The information is conveyed through afferent nerves to the brainstem and elicits counter-regulatory adjustments in sympathetic and parasympathetic efferent activity, thereby stabilizing BP. Given the importance of the baroreflex in BP control, damage to the afferent portion, the integrating brainstem centers and/or the efferent portion of the baroreflex results in severe BP abnormalities. Despite anatomical preservation of the aortic baroreflex arch, baroreflex failure occurs following bilateral damage to carotid baroreceptors or baroreflex afferent nerves [907–909]. Damage to the efferent portion of the baroreflex is usually part of autonomic failure [909]. Both conditions differ dramatically in their clinical presentation and in either case BP and more general management is not supported by trials with hard clinical endpoints.

Clinically, baroreflex failure is characterized by extreme BP variability with dramatic hypertensive surges, hypotensive episodes and orthostatic hypotension in some but not all patients [907–909]. Causes of baroreflex failure include neck dissection or radiation therapy, bilateral resection of carotid body paragangliomas, familial dysautonomia (hereditary sensory and autonomic neuropathy type 3), and, very rarely, brainstem lesions [907–910]. The diagnosis of baroreflex failure should be confirmed by pharmacological baroreflex testing. Because hypertensive episodes in patients with baroreflex failure are mediated through unrestrained sympathetic activation, which are exacerbated by psychological and physiological stress, long-acting central sympatholytic agents are the mainstay of therapy [909]. Vasodilators and diuretics can dramatically lower BP in patients with baroreflex failure and should be avoided whenever possible. Hypotensive episodes are usually managed using nonpharmacological means. Bradycardia through unrestrained cardiac parasympathetic activation may require pacemaker implantation [911].

The clinical presentation of autonomic failure is characterized by a variety of manifestations of loss of autonomic BP control, including posture-related BP abnormalities with orthostatic hypotension and, in many patients, supine hypertension [912]. Causes of autonomic failure should be differentiated to identify potentially treatable conditions, such as autoimmune-mediated autonomic ganglionopathies [913]. In most cases of severe orthostatic hypotension due to autonomic failure, symptomatic therapies are required. Nonpharmacological therapies, such as venous compression garments, increased salt ingestion, and sufficient water drinking should be tried first [914]. Drugs that could worsen orthostatic hypotension should be discontinued whenever possible. Pharmacological therapies, such as alpha-adrenoreceptor agonists, may be required in patients who remain symptomatic on nonpharmacological therapy [915]. All current antihypotensive drugs, particularly longer acting mineralocorticoids will worsen supine hypertension. Supine hypertension increases urine excretion during the night and worsens orthostatic hypotension the next morning, but data on CV risk in this condition is limited. Sleeping with the whole bed tilted head-up is recommended to lower BP during the night and to decrease orthostatic symptoms the next morning [916]. Antihypertensive therapies, given before sleep, can be considered in selected patients. However, their potential long-term benefits on CV risk have to be weighed against the risk of falls and the prognosis of the condition causing autonomic failure [916].

Baroreflex failure and autonomic failure

figure40

15. HYPERTENSION IN DIFFERENT DEMOGRAPHIC SITUATIONS

15.1 Blood pressure in children, adolescents and transition period

15.1.1 Blood pressure measurements in children and adolescents

There is a general consensus that starting from the age of 3 years (or earlier in children with risk factors for high BP), BP should be regularly measured [917–920]. Normality BP tables, derived from studies using manual auscultatory measurements, provide BP distributions according to sex, age and height percentile [917,921,922]. In children and adolescents, BP should be measured by manual auscultatory devices [917,922]. Automated devices are increasingly used, but they should be validated specifically in children (www.stridebp.org) [62], and the detection of a child's BP elevation by an electronic device, needs confirmation with auscultatory BP measurements [917,922]. A cuff with appropriate size for the individual child's arm circumference should be used. For manual auscultatory devices, the inflatable bladder length should cover 75–100% of the child's middle upper arm circumference and 37– 50% of the upper arm width [62,917]. Use of automated electronic devices should follow the same rules as for standard BP measurements and physicians should follow the device's instructions.

15.1.2 Hypertension in children and adolescents

In children and adolescents, BP increases with age and body size, making it impossible to use a single BP cutoff value to define hypertension. Hypertension in the age range of 1– 15 years is diagnosed when office BP values are found to be equal or above the 95th percentile of the normative BP distribution for age, sex and height percentile, persistently on at least three separate occasions [62,917,923]. Because of its superior reproducibility and association with HMOD compared to office BP [62,917,919,924,925], ABPM can be a valuable source of additional information, its elevation being also based on 24 h mean values ≥ 95% percentile [62,917]. ABPM is also indicated for the evaluation of BP control during treatment, the confirmation of resistant hypertension [62,917] and the identification of WCH and MH, which are also not uncommon in children [917,926]. Information on HBPM in children is limited [927], but it should be considered that in primary care, HBPM is more easily accessible and practical, particularly for monitoring children treated for hypertension [919,927]. Other diagnostic steps involve detection of secondary hypertension, which is more frequent than in adults, as well as of additional CV risk factors and HMOD. In the absence of outcome trials, the treatment goal is to reduce office BP below the 95th percentile, but lower BP targets (below the 90th percentile) are regarded as desirable in children with HMOD or secondary hypertension. Stricter BP targets, i.e. below the 75th percentile or below the 50th percentile for 24 h mean BP, are recommended for CKD either without or with proteinuria [917,928,929]. Treatment should start with lifestyle changes, within which loss of body weight has a primary importance because of the close association of hypertension and obesity in adolescents [917,930,931]. The decision to use antihypertensive drugs should be based on failure to reach BP control and also, concomitantly with lifestyle interventions, on the BP level (grade 2 hypertension), the presence of signs or symptoms related to the BP elevation and the evidence of HMOD. Drug treatment should be implemented according to a step care approach. The same five major drug classes validated in adult hypertensive patients are recommended [917,930,931].

15.1.3 Transition period to adulthood

In hypertensive adolescents aged 16 years or older, the consensus is to shift to the diagnostic and treatment criteria largely similar to those used in adult hypertensive patients. That is, to (i) identify hypertension by office BP values ≥ 140 mmHg for SBP and/or ≥ 90 mmHg for DBP, (ii) pursue an office BP target of <130/80 mmHg aiming at <125/75 mmHg in the presence of HMOD or CKD [4] and (iii) lower BP by the same nonpharmacological and pharmacological treatment strategies used in adults. The transition of an adolescent with hypertension from pediatric to adult-like care is a well recognized difficult period for a variety of reasons. Full acceptance of new doctors by the patient may take time. Diagnostic steps and therapeutic adjustments may interfere with the educational challenges, intense social and community life of adolescents. The asymptomatic nature of the BP elevation and the remoteness of its possible adverse consequences may favor underappreciation of the risks and poor adherence to the prescribed treatment. Adherence has been reported to be frequently low in the transition period [932]. During transition, close collaboration and sharing of clinical information between pediatricians and adulthood physicians is of crucial importance. The age and the duration of the transition period should be flexible, depending on the maturation of individual adolescents, family, socioeconomic characteristics and the disease severity and comorbidities. Parents’ involvement plays an important role. After transition is completed, patients should be closely followed to detect the BP trajectories in the subsequent years, which include, in some cases, spontaneous BP normalization.

BP measurement in children and adolescents

figure41

15.2 Hypertension in young adults

Recent large-scale longitudinal studies have expanded knowledge on the relationship of high BP with nonfatal CV events and mortality [933,934] in young people. In a subgroup of young patients (n= 5000, age <40 years) from the CARDIA study followed for about 19 years [935], the risk of CV events increased progressively with a BP increase, and in hypertension, (SBP ≥ 140 mmHg) it was 8.4 times greater than in normotension [936]. This was the case also in another study, based on almost 25 million young Korean individuals (20– 39 years old, median age 31 years) in whom an SBP between 130 and 139 mmHg and a DBP between 80 and 89 mmHg (grade 1 hypertension according to the ACC/AHA classification) were accompanied by a 25% increase in the risk of CV events [43]. The increase was seen in men and women, it was related to either SBP or DBP, and it was visible also in conditions of ISH and IDH [857], although in either case, it was less than when the elevation involved both SBP and DBP values. Finally, the association between BP and CV events has been documented also by a meta-analysis of 17 observational studies, including more than 4.5 million young adults (age18– 45 years) followed for about 15 years. This leaves no doubt that the relationship between BP and adverse CV outcomes is similar in young as compared with middle-aged and older people, although predictably in studies on younger people, the absolute number of events was lower. On the other hand, whether and how much in young people (i) the hypertension-related risk can be reduced by BP-lowering treatment and (ii) some drugs, e.g. those which more effectively delay, prevent or regress subclinical HMOD, are more protective than others remains an unresolved problem. This is the case because, although in most outcome-based RCTs in hypertension recruitment criteria included patients aged 18 years or above, data on young people have invariably been extremely scarce. In a large individual-participant meta-analysis of RCTs on hypertension in the youngest group from which data were available (<55 years of age), the median age was around 50 years[484]. Unfortunately, this evidence gap can hardly be reduced by a study on young people from Korea [43], in which the risk of CV outcomes was modestly and not significantly greater in the treated fraction of the population than in normotensive controls, because of the potential confounding of observational studies in which therapeutic interventions are addressed. While waiting for the urgently needed outcome-based RCTs in young hypertensive patients, the only option left is to extrapolate the results obtained in middle-aged and old people to hypertensive young patients and treat them according to the same recommendations. This finds support by the now clearly demonstrated relationship of BP to CV outcome at a young age. In this context, it is also urgent to develop specific risk charts for younger people, given that the SCORE2 system provides risk quantification only starting from 40 years of age [55].

15.3 Hypertension in older persons

Hypertension is extremely frequent in older individuals, where it is accompanied by an increase in the risk of CV and kidney outcomes almost throughout the entire old age range. In old people, SBP is prognostically much more important than DBP, and ISH is the predominant hypertension phenotype, particularly above 70 years of age. Some studies have shown that in old people, a pulse pressure >65 mmHg could be an independent risk factor for CV morbidity and mortality[937]. Although chronological age is not invariably the most important criterion for defining diagnostic and therapeutic strategies for hypertension, a number of considerations suggest that two age thresholds might be usefully considered. One age threshold is 65 years, i.e. the age at which an acceleration of arterial aging is observed, leading to a sharp increase in SBP and pulse pressure with a concomitant decrease in DBP [937]. Moreover, after this age, most people become inactive professionally, with a significant impact on somatic and psychological health. However, the majority of people aged 65– 79 years have a good functional status and do not need help for most of their daily living activities. For individuals of this age group with significant alterations of functional status and autonomy, the strategies described for the group aged 80+ years (see Section 15.3.2) should be applied. The second age threshold is 80+, an age range within which a large number of individuals present with several comorbidities, frailty and loss of functionality [938]. At the same time, a significant percentage of this population maintains well preserved somatic, cognitive and psychological health. The consequence is that the 80+ years are a most heterogeneous age group in terms of functionality while being at the same time the most markedly growing age group in the world, especially in Europe [939].

15.3.1 Patients 65–79 years old

15.3.1.1 Threshold and target for drug treatment

There is strong evidence from RCTs that in this age range, antihypertensive therapy significantly reduces CV morbidity as well as CV and all-cause mortality. In addition, antihypertensive treatment can generally be well tolerated [940,941]. The previous guidelines [4] supported BP thresholds for drug treatment similar to those of younger people, i.e. 140 mmHg SBP or 90 mmHg DBP. Target SBP values within the range of 130–139/80 –89 mmHg aiming at a BP close to 130/80 mmHg, if well tolerated, was further recommended. However, more recent data suggest that lower BP targets might be considered. In hypertensive people aged 60– 80 years, the STEP trial [496] has shown that treatment to an SBP target of 110– 129 mmHg (mean 126.7 mmHg) resulted in a lower incidence of CV events compared to standard treatment, with a 130–149 mmHg SBP target. A reduced number of events has also been shown in older patients of the SPRINT [503] trial to reach an average on-treatment SBP of about 122 mmHg compared an on-treatment value between 130 and 139 mmHg. However, the SPRINT trial [942] has the limitations discussed above and in the previous guidelines [4]. This is the case also for the STEP trial, which made use of a reference group that included patients with uncontrolled BP (i.e. >140 mmHg), possibly amplifying the incidence of outcomes [943]. Furthermore, STEP used drug discontinuation to randomize patients with initially lower BP values to the higher target BP group, again with possible outcome amplification. Nevertheless, in a meta-analysis of 32 RCTs in 96 549 patients largely confined to a 65– 80 years of age group (i) antihypertensive treatment was associated with a reduction of CV outcomes when patients had a baseline SBP 140 mmHg; (ii) reducing SBP to <130 mmHg was associated with an incremental reduction of CV events and mortality compared to patients in whom the SBP reduction left on-treatment SBP values in the 139–130 mmHg range and (iii) this was also the case when DBP was reduced to <80 mmHg compared with remaining in the 80– 89 mmHg range. Data in older patients were similar to what was seen in younger patients and indeed, the linear relationship between BP reductions and outcomes over a 40 mmHg SBP change was almost superim-posable in the two age groups [144,145,489]. Thus, it seems appropriate for the present guidelines to somewhat modify the previously recommended BP targets in hypertensive patients aged 65– 79 years. That is, (i) to emphasize the recommendation to have as the initial goal a reduction of SBP to between 140 and 130 mmHg, because this guarantees an appropriate trade-off between the degree of CV protection and the incidence of side effects (ii) to consider, if treatment is well tolerated and no overt signs of organ hypoperfusion emerge, a further SBP reduction to <130 mmHg (iii) to reduce DBP, which is of lesser importance in this age range, to <80 mmHg and (iv) to make no attempt to decrease SBP/DBP to <120/70 mmHg. It is important to mention that threshold and target BP values are not identical to those of patients with ISH aged 65–79 years. In trials documenting the benefit of antihypertensive treatment in ISH, initial SBP was 160 mmHg while on-treatment SBP was confined to the 149–140 mmHg range [144,145,497,498], which supports the recommendation to treat patients with grade 2 or 3 ISH and to reduce their SBP to <150 mmHg, a conservative target that might diminish the risk of an excessive DBP reduction. However, as addressed in Section 14.6, a considerable number of older patients with ISH or a prevalent SBP elevation was included in RCTs showing CV protection at lower SBP threshold and target values, i.e.140– 159 and <140 mmHg SBP, respectively [499,500]. This provides a rationale for the additional recommendation that in ISH patients (i) physicians can consider treatment also with a grade 1 SBP elevation and (ii) cautiously aim at on-treatment SBP values between 130 and 139 mmHg, if treatment is well tolerated and DBP does not show a reduction below the safety values.

15.3.1.2 Antihypertensive treatment strategies

Treatment of older patients should make use of lifestyle interventions as in younger patients. However, in subjects 80 years or older the measures indicated for younger patients (see Section 7) may have to be adapted. Although overweight and obesity remain deleterious for CV and metabolic health, weight loss programs may lead to muscle mass loss, sarcopenia and malnutrition. Therefore, except for the case of severe obesity or with robust old people, weight loss is not recommended. Also, salt restriction may contribute to loss of appetite with deleterious effects on nutritional status and therefore should not be adopted, except in cases of very high salt consumption (e.g. NaCl >10 g day). In these patients, a particular effort should be devoted to the promotion of physical activity, adapted to the individual's capacity and cultural context. Collective physical activities (e.g. dance, tai-chi or walking) should be preferred to also promote social contacts and fight loneliness and social isolation in older people. In older patients (even those in healthy conditions) who experience drug-related side effects [944], medical efforts should aim at achieving the BP goal and at the same time avoid side effects, the risk of which increases with the number of prescribed drugs. In old patients and particularly frail patients, initial monotherapy may be considered the first treatment step more frequently than in younger patients, especially with grade 1 hypertension. However, because controlling the SBP elevation of older patients is particularly difficult, even in grade 1 hypertension, physicians should consider an initial two-drug combination treatment in the most fit patients or be prepared to uptitrate treatment with the addition of a second and, if needed, a third drug. Combination treatment is necessary in the vast majority of the patients with grade 2 or 3 hypertension, and in these patients, it can usually be considered as first step treatment because initial combination treatment favors better adherence to treatment and reduced treatment inertia [608], also in old patients. This is particularly important because due to their multimorbidity, more than half of older patients are exposed to polypharmacy, i.e. are treated with at least five agents [945,946]. Thus, in the very old and in the most frail patients, initial monotherapy should be considered in order to avoid adverse side effects of multiple drug regimens. As a general rule, and although delayed achievement of BP control should be avoided, it may be wise to start antihypertensive drugs with lower doses and uptitrate treatment somewhat more slowly in older than in younger patients.

15.3.1.3 Antihypertensive drugs

In the absence of specific indications, there is no evidence of more pronounced long-term benefits or harms of any specific drug class. Therefore, any of the five major drug classes can be used. However, older patients may be more susceptible to side effects associated with BBs, most importantly fatigue or sleep-related disorders (unusual dreams or insomnia) [592] that can negatively impact on the quality of life. Therefore, in older individuals, BBs should not be a general first choice for treatment in the absence of GDMT indications or other conditions where their use is recommended (Table 16). However, in clinical practice there are many cardiac, vascular and non-CV conditions, for which BBs are indicated, and their prevalence is high in old people as well [591]. Specific prescription rules for older patients are available and may be of major help for reducing drug-related adverse effects [947].

15.3.1.4 Monitoring the effects of treatment

The search for orthostatic hypotension should be systematic in people aged 65– 79 years, even in the absence of symptoms. HBPM should be implemented and help to better define the usual BP values in the face of the higher SBP variability in this age group. These measurements can be proposed even in patients with mild-to-moderate dementia, including patients with Alzheimer disease and other neurocognitive disorders. ABPM can be useful, especially in patients with polypharmacy, to identify hypotensive episodes and obtain information on the presence and magnitude of the night-time BP reduction.

15.3.2 Patients 80 years old or beyond

There is only one outcome-based RCT in this age group, i.e. the HYVET trial [502], showing that in hypertensive patients aged 80 years or more (mean age 83 years), antihypertensive treatment was accompanied by a reduced risk of CV events and all-cause mortality [compared to placebo (relative risk – 21%), with a major benefit for HF (relative risk – 64%). Globally, the benefits of BP-lowering treatment in moderately frail octogenarians seems not to be different compared with those observed in fit older adults. For example, in a post hoc analysis of its database, the HYVET study showed that frailty did not modify the favorable impact of antihypertensive treatment [948]. Furthermore, the SPRINT study reported that in a subgroup of patients aged >75 years, the benefit of intensive BP control was observed independently of their frailty level [503]. In both studies, the therapeutic strategies and the threshold and target BP values were superimposable to those of the entire study populations. Finally, a recent analysis of a very large real-life population from northern Italy, old patients defined as frail by a large number of comorbidities and a high risk of mortality within few years, exhibited a reduced risk of death, if their adherence to antihypertensive drug prescription was high [512]. In very old hypertensive patients definable as fit for their age or only moderately frail, antihypertensive drug treatment should be implemented together with lifestyle modifications, including promotion of mild physical activity, preferentially in a collective group context. As mentioned in Section 15.3.1.2, social contacts should also be favored in order to avoid isolation and depression. Unless very old patients are robust or severely obese, weight loss programs should be considered with caution because of possible deleterious consequences such as loss of muscle mass and malnutrition.

No RCTs are available in patients close to 90 years of age or beyond. Data on therapeutic strategies and BP targets are also missing in patients with documented severe loss of autonomy due to the fact that trials on antihypertensive therapy in older people target healthier individuals and exclude those with loss of autonomy and limited life expectancy [949,950]. Several observational studies have shown that the relationship between BP and mortality is influenced by the level of frailty [504,505] and that in the most frail old patients, the rates of morbidity and mortality are higher at low BP (mainly SBP <130 mmHg), especially when low BP levels are observed in the presence of antihypertensive treatment [506,507,509–511,945]. In these patients, the treatment strategy is frequently referred to as ‘start low and go slow’. Overall, the evidence appear to favor the beneficial effects of BP-lowering interventions in old, moderately frail hypertensive individuals, albeit the opposite may happen in very frail individuals. In general, no information is available on whether and to what extent treatment BP thresholds, treatment BP targets and drugs should differ between, e.g. a robust 85-year-old patient person with high physical, social and cognitive capacities, and a very frail 85-year-old person with total loss of autonomy. The amount of loss of functionality and increased risk of death that should lead to changes in treatment strategies are also not established [38,945]. Despite these limitations, in patients aged >80 years, frailty and functionality should be part of the diagnostic work-up. How to assess frailty and functionality at clinical practice level is reported below together with some suggestions on how to adapt treatment strategies to the frailty level.

15.3.3 How to assess the level of frailty/functionality to better personalize therapeutic strategies

Tools that can be used by physicians, nurses or other healthcare professionals in daily practice need clinical validation, standardization, limited requirement of time to be completed (less than 15 min), no need of complicated/expensive specific devices and no requirement of specialized skills (except for a short course for a minimal theoretical education and practical training). The clinical frailty scale [951] is a validated scale classifying the 80+ year-old people according to their frailty level, in an easy, rapid and standardized way (Table 21). This tool should be employed before the initiation of treatment and repeated annually in order to monitor the evolution of a patient's functionality/autonomy and personalize the treatment strategies (Table 21) [38]. For fit patients with an ADL (Activities of Daily Living) score of 5/6 [952], absence of clinically significant dementia and capacity of routine walking activities, treatment strategies and objectives should be similar to those of the 65–79 years group. For patients with intermediate functional status, i.e. patients with moderate functionality impairment and partial loss of autonomy, antihypertensive treatment strategies should be more conservative, i.e. treatment may start when SBP ≥ 160 mmHg, targeting an SBP range between 140 and 150 mmHg. Progressive reduction of antihypertensive drug treatment could be considered if SBP is <120 mmHg or in the presence of orthostatic hypotension [953], although little information is available on the consequences of ‘deprescribing’ from studies in old patients [954]. In this group, a more detailed evaluation of the functional/autonomy status should be conducted to better personalize therapeutic strategies. For Group 3, i.e. patients with severe loss of functionality/autonomy (ADL 2, with severe dementia, chronic bedridden or end of life), indication of treatment should be individually decided according to symptoms, comorbidities and polypharmacy. Progressive deprescribing should be considered if SBP is <130 mmHg or in the presence of orthostatic hypotension. The results of ongoing RCTs in very frail individuals will provide more evidence on the benefits/risks of reduction of treatment in these patients.

TABLE 21 - Adapting BP-lowering strategies in patients older than 80 years according to their functional/autonomy status (adapted from [38])
Group 1 Group 2 Group 3
Characteristics Fit Slowed but autonomous for most activities Severely dependent
Diagnosis -ADL (Katz) ≥5/6
and
-absence of clinically significant dementia (MMSE≤20/30)
and
-routine walking activities
-Profile between Groups 1 and 3 -ADL (Katz) ≤2/6
or
-severe dementia (MMSE ≤10/30)
or
chronic bedridden
or
-end of life
Therapeutic strategy - Initiate drug treatment if office SBP ≥160 mmHg
- However, in most cases initiation of treatment in the 140 to 159 mmHg range maybe considered
- Office SBP should be lowered in the 140 to 150 mmHg range
- However, reduction of office SBP between 130 to 139 mmHg may be considered if well tolerated, albeit cautiously if DBP is already below 70 mmHg
- Consider to start with monotherapy
- Initiate drug treatment if office SBP ≥160mmHg
- However, a lower office SBP threshold in the 140 to 159 mmHg range may be considered
- Office SBP should be lowered in the 140 to 150 mmHg range
- However, reduction of office SBP between 130 to 139 mmHg may be considered if well tolerated, albeit cautiously if office DBP is already below 70 mmHg
- Consider to start with monotherapy and titrate antihypertensive medication cautiously
- Consider treatment reduction if office SBP is very low (<120 mmHg) or in patients with orthostatic hypotension
- Make a more detailed evaluation of the functional status: SPPB (mobility), handgrip (muscular force), mini-GDS scale (depression), and MNA-SF (nutritional status)
- Priorize therapeutic strategies according to comorbidities and polypharmacy issues
- Consider treatment if office SBP ≥160 mmHg
- Target treatment of office SBP to a range of 140–150 mmHg
- Reduce treatment if office SBP is very low (<120 mmHg) or in patients with orthostatic hypotension
- Correct other factors and medications decreasing BP
ADL[952]: Activities of Daily Living (Katz Index) scaled rated from 0 (completely dependent) to 6 (completely autonomous). This scale comprises 6 ADL: Bathing, Dressing, Toileting, Transferring, Feeding and Continence. For each ADL ‘0’ means that the person is unable to do it without assistance, 0.5 need of some assistance, 1 no need of any assistance.MMSE[955]: Mini mental status evaluation. Score 0-30, 30 best, 0-10 severe dementia, 11-20 moderate dementia.SPPB[956]: Short Physical Performance Battery. Combined test including a balance test, a gait speed test and a 5-time chair stand test. Each one of the 3 tests is scored from 0 to 4, 4 best. The total score is from 0 to 12, 0-6 high risk of falls, 10-12 low risk of falls.Mini GDS[957]: Geriatric depression scale. 4 questions score 0 to 4 (classically score 3-4 indicates a high risk of depression).MNA-SF[958]: Mini nutritional assessment short form. Six items, score 0-14 (0-10 possible malnutrition, 14 best).

Treatment strategies in older persons

figure42

15.4 Sex and gender aspects in hypertension

15.4.1 Epidemiology and pathophysiology

In hypertension, as in many other diseases, there are sex (a biological characteristic) and gender (a social construct) differences that have an important impact on its pathophysiology, epidemiology and clinical management. In 2019, the worldwide age-standardized prevalence of hypertension, defined by SBP 140 mmHg and/or DBP 90 mmHg, or taking medication for hypertension, was 32% in women and 34% in men [37]. However, large disparities were observed within world regions. Thus, in Western Europe, the prevalence of hypertension in individuals aged 30–79 years ranged between 17.5 and 30% in women and 26 and 43% in men, whereas in Eastern Europe, it ranged between 34 and 46% in women and 43 and 56% in men [37]. The main difference between men and women is the life trajectory of BP that is apparent already from age 12 [37,959]. Hypertension prevalence increases with age in both sexes [959], but it tends to be lower in premenopausal women than in men of the same age, with a marked rise in women after menopause [39]. After the age of 65, the prevalence of hypertension in female exceeds that of male individuals [37,39,959].

The differences exhibited by women in BP development over the life course and the age-related hypertension can be explained by differences of BP regulatory mechanisms in male and female individuals – most likely a combination of sex and gender-specific factors [37,959]. In premenopausal female individuals, estrogens contribute to lower BP in the context of their general vascular protective effect. Protection is mediated through different mechanisms including endothelial vasodilatation via upregulation of the nitric oxide pathway and inhibition of the activity of SNS and RAS. Moreover, estrogens decrease endothelin production, decrease oxidative stress and reduce inflammation [960]. After the menopause, the marked decrease in estrogen levels partially explains why BP and the risk of hypertension increase [37,960]. In postmenopausal women, androgens may contribute to increased BP and the age-associated CV risk as they do in men [961]. In this context, young women with estrogen/androgen imbalance and conditions such as premature ovarian insufficiency, polycystic ovarian disorders and infertility are at increased risk of developing hypertension [962]. Although estrogens have a protective role in premenopausal women, the administration of exogenous estrogens to menopaused women has no effect on BP and does not affect the risk of CV outcomes.

15.4.2 Blood pressure and cardiovascular risk

In the pooled data of the IDACO study, the absolute CV risk was lower in women than in men, while the increase in risk associated with 24 h and night-time BP was steeper in women than in men. The proportion of events potentially preventable by BP-lowering may, therefore, be greater in women [963], making a wider implementation of ABPM in women desirable. In a meta-analysis of 27 542 individuals without baseline CVD (54% females), the increased risk for CV events, including myocardial infarction, HF and stroke, associated with SBP elevations was visible at lower SBP ranges in female than in male patients, suggesting that the definition of optimal SBP might differ between men and women [964]. One important issue is that several sex-specific events occurring in women, including hypertensive disorders of pregnancy or the polycystic ovary syndrome, are associated with a greater long-term risk of hypertension and CVD. Unfortunately, a large proportion of women are not aware of their increased CV risk due to insufficient screening, particularly among younger women, ethnic minorities and low socioeconomic groups [965].

15.4.3 Differences in clinical phenotypes

15.4.3.1 White-coat hypertension and masked hypertension

The IDACO study reported that WCH exponentially increased from individuals aged 18– 30 years to individuals aged 70 years, with limited differences between men and women (8.0 versus 6.1%; P= 0.0003). However, data from national and international registries consistently report a higher prevalence of WCH in women [966]. A high prevalence of WCH was observed in older individuals and pregnant women [967]. In contrast, MH is generally more prevalent in men than in women as shown by the Spanish ABPM registry (43 versus 26%) [968] and the IDACO registry (21.1 versus 11.4%) [969].

15.4.3.2 Hypertension-mediated organ damage

Female and male patients develop HMOD, but differences in HMOD frequency, severity and reversibility occur according to sex [970]. LVH is more prevalent and less modifiable by antihypertensive treatment in women than in men [971,972]. In the Campania Salute Network project, new LVH was more frequently detected in women [973], and women with LVH had about the same CV risk as men. LA dilatation, an early sign of hypertensive heart disease, has been reported to be more common in women than in men with hypertension [974,975].

Regarding arterial stiffness, PWV decreases in women after puberty. Thereafter, both sexes experience an increase in arterial stiffness with age, with women showing a more rapid increase after menopause. Older women experience greater aortic stiffness and arterial pulsatility than their male counterparts, seemingly contributing to ISH, uncontrolled hypertension, HFpEF and aortic stenosis, all being more common in women [976]. A greater adverse prognostic significance in women of arterial stiffness has been observed in patients with CAD [977].

There is also evidence that the impact of hypertension on kidney function and disease progression may have a sex-dependent component [978]. The prevalence of albuminuria was lower in postmenopausal female individuals than in male individuals. In a cohort of 2379 Chinese individuals, the association between BP and eGFR differed in male and female individuals, suggesting that men were more sensitive to hypertension-induced changes in kidney function over time than women [979]. Moreover, in a meta-analysis of 68 studies of patients with various forms of CKD, a less rapid decline in renal function was observed in women [980].

15.4.4 Sex differences in hypertension outcomes

The development of CV outcomes in relation to hypertension is also influenced by gender. An important observation is that hypertension-related outcomes, e.g. myocardial infraction, HF, stroke and CVD, may occur at lower SBP levels in women [964]. In the INTERHEART study, the increased risk of myocardial infarction associated with hypertension was greater in older female than in male patients [531]. At a similar age, Norwegian population-based study that enrolled 33 859 individuals (51% women), men were at greater risk of myocardial infarction as BP increased [981]. However, both for SBP and DBP increases, the association with BP was closer for female patients. Similar conclusions were drawn from a large UK cohort study [982] as well as from studies in younger individuals [857,983]. In the Hordaland Health study, which had a 16.2 years FU, stage 1 hypertension in the fourth decade doubled the risk of myocardial infarction in midlife female individuals, while this relationship was not observed in male individuals [983]. Hypertension increases the risk of HF in both sexes [984], but the increase in risk is greater in female than in male individuals [985]. Clustering of female patients was observed among patients with HFpEF, where females represent 55–70% of patients. This is different for HFrEF, in which females have been reported to be 30– 40% of the overall number of patients with HFrEF [203,986–988]. Although differences in the age distribution of the patients at risk (because of the longer life expectancy of female individuals) may have contributed, hypertensive female patients have been reported to develop more LVH, vascular and myocardial dysfunction compared with hypertensive male patients [960,976], with, thus, a possible sex-related contribution to the development of CAD and HF [973,986]. A Norwegian study reported a stronger association of an elevated SBP with incident AF in female than in male patients [989], but this finding has not been consistently confirmed by other studies [990,991]. Recent studies also suggest that, in line with previous evidence, stroke risk starts to increase at a lower BP in female patients [964,992]. Hypertension also seems to be a stronger risk factor for dementia and cognitive decline in female individuals [993,994].

15.4.5 Benefits of antihypertensive treatment and target blood pressure

Do women receive the same CV outcome benefits from antihypertensive treatment than men? Because of the global under-representation of women in trials, answering this question is difficult. Moreover, the statistical analyses of interventional trials commonly used a binary approach that does not appropriately capture the clinical specificities of men and women. Only few clinical trials in hypertension report treatment results stratified by sex. The NORDIL study showed similar treatment effects in both sexes [995]. The LIFE trial found similar treatment effects in both sexes, but female participants randomized to treatment with losartan had a greater reduction of the primary endpoint, all-cause mortality, and new-onset diabetes [996]. Three other studies showed sex differences favoring treatment in women. In the HOT trial, the target DBP of <80 mmHg reduced myocardial infarction in women but not in men. On the other hand, low-dose acetylsalicylic acid reduced incident myocardial infarction only in men but not in women[997]. In the ANBP-2 trial, the benefit of ACEi treatment was only demonstrated in male participants [998] and in the VALUE trial, amlodipine lowered BP and reduced the primary endpoint (composite of cardiac mortality and morbidity) more effectively in female than in male individuals. A meta-analysis of RCTs (103 268 men and 87 349 women) found comparable reductions in BP and incidence of CV events in both sexes for treatments based on BBs, ACEis, ARBs, CCBs or diuretics [941]. Limited information from RCTs is available on whether BP targets should be different in women compared with men, in part because no trial was adequately powered to investigate outcomes at different on-treatment BP levels separately in the two sexes, a limitation that was particularly evident in women because of their lower number of CV outcomes [97]. In a post hoc analysis of SPRINT, in which data were analyzed separately for men and women, the primary composite CV outcome was achieved in men but not in women [999]. In another post hoc analysis of the same study in which a propensity score matching was used to equalize patients’ baseline characteristics (in SPRINT, randomization was not stratified according to sex), women also did not exhibit a significant outcome difference at standard versus intensified treatment. The low baseline CV risk in female patients may have accounted for this finding [1000]. Given the low number and the limitations of the available studies, there is insufficient evidence to recommend different BP targets in women compared to men.

15.4.6 Sex differences in hypertension management

Lifestyle modifications are important for nonpharmacological management of hypertension, and sex differences in their effects have been noted. In the DASH trial, dietary sodium restriction induced pronounced BP reductions only in female individuals. Regarding physical activity, a meta-analysis of 93 trials assessing the impact of a structured aerobic exercise therapy found that exercise induced a greater BP reduction in male than in female participants [357,1001]. Regarding drug therapy, it is important to mention that there are disparities in the prescription of antihypertensive drugs between male and female individuals, indicating that female patients are less likely to receive antihypertensive therapies than male patients for comparable BP values [1002]. In a recent real-world analysis conducted in Sweden, hypertensive women had a higher BP, less antihypertensive treatment and a worse BP control, with female sex being a significant predictor of less intensive antihypertensive therapy [1003]. There are no established differences in pharmacokinetics of antihypertensive drugs between women and men that warrant sex-specific dosing. This applies also to the unclear relevance of differences in body weight and body composition between men and women. Adverse effects from antihypertensive drugs are reported more often for women than for men, even when women are taking fewer drugs [1004], and women have a 50% greater risk of suffering from adverse reactions compared to men [1005]. A higher incidence of ACEi –induced cough and CB-induced ankle edema has been observed in women [1006]. Women were more likely to experience hypokalemia and hyponatremia with diuretics, although less likely to experience gout [1007]. There is no consistent data on sex differences in the efficacy of antihypertensive drugs. Therefore, drug selection and dosing should not be based on the sex, but treatment individualization should consider also sex-specific adverse drug reactions. Clearly, specific caution and contraindications must be considered for women planning pregnancy or who are pregnant (see Section 16.1). Whether drug adherence is better or worse in female compared with male individuals remains controversial. No difference in adherence was found between sexes in a meta-analysis of 82 studies [1008]. In studies on resistant hypertension, where adherence was assessed using highly sensitive methods measuring drugs in urine, drug adherence was lower in women [669]. This may be explained, at least in part, by depression, a known risk factor for nonadherence, which has a greater prevalence in women.

15.4.7 Infertility treatments and hypertension in women

Following any invasive assisted reproductive therapy, the risk ratios for gestational hypertension, preeclampsia and a combination of both, increase by 54% independently of the gestation order [1009]. Moreover, a recent meta-analysis has confirmed that pregnancies conceived by in-vitro fertilization, with or without intracytoplasmic sperm fertilization, are at higher risk of being complicated by HDP including preeclampsia, compared with spontaneously conceived pregnancies [1010].

15.4.8 Oral contraceptive pills and hypertension

Physicians counselling women with CV risk factors, including hypertension, should balance the potential risks of contraceptive methods against those of an unintended pregnancy. Older studies have demonstrated a relationship between the use of a combined estrogen– progestin oral contraceptive pill (mostly with a higher dose of estrogens), venous thromboembolism, and, to a lesser extent, myocardial infarction (especially with a concomitant smoking history) and stroke [1011,1012]. However, the dose or type of estrogen in the combined pill may change the strength of this association [1012–1014]. The evidence so far supports that progestin-only pills do not increase thrombotic events, while injectable forms of progestins require further evaluation [1015–1017]. In normotensive women, formulations of oral low-dose estrogen– progestin combinations are associated with a 5 mmHg increase of either SBP or DBP [1018]. The development of hypertension is almost 2% over 4 years [1019]. Initiation of low-dose estrogen– progestin formulations can be advised for normotensive women without CV complications or additional risk factors after careful BP monitoring that should be regularly repeated (every 3–6 months) [1020]. In newly diagnosed women with grade 1 hypertension or treated hypertensive women with BP levels within the target range, a combined estrogen–progestin pill may be considered, if no other method is appropriate. Newly diagnosed women with (i) higher hypertension grades, (ii) on-treatment uncontrolled hypertension or (iii) a history or a high risk of CVD, should not receive estrogen-based contraceptive pills [1021], and alternative forms of contraception should be advised. Discontinuation of combined estrogen– progestin oral contraceptives in women with hypertension may improve BP control because BP usually decreases promptly following pill withdrawal [1022]. Progestin-only pills have no significant effect on BP levels, and measurement of BP at initiation or during contraception is not required [1021,1023]. However, in women with more severe forms of hypertension, the initiation of progestin-only pills should be individualized within the context of additional risk factors. At the same time, it should be carefully considered in women with a history of CVD irrespective of BP levels [1021]. Combined hormonal contraceptives are not recommended in smokers aged 35 years or older [1023]. In premenopausal women, estrogen use with an oral contraceptive pill increases BP. Although SNS and RAS activation may be underlying mechanisms, whether this is because of the effects of estrogen, progesterone or a complex interaction between the two is not well understood. Studies on the effects of progesterone are limited, with short-term FU and mostly observational designs. Nevertheless, as of today, data do not suggest any increased risk of hypertension or short-term cardiometabolic outcomes with progesterone-only contraceptive pills.

15.4.9 Hormone-replacement therapy and hypertension

Cross-sectional studies have long established that menopause doubles the risk of developing hypertension, even after adjusting for factors such as age and BMI [1024]. Although hormone-replacement therapy contains estrogens, there is no convincing evidence that BP will increase significantly in menopausal women with or without hypertension [1025]. However, after the initiation of hormone-replacement therapy, it is reasonable to measure BP to confirm persisting normotensive BP values or regular BP control by treatment. In the case of uncontrolled BP, hormone-replacement therapy should be stopped. In summary, current evidence suggests that the use of hormone-replacement therapy is not associated with an increase in BP. If BP levels can be controlled with antihypertensive medications, women may receive hormone-replacement therapy.

15.4.10 Gender-affirming hormone therapy and hypertension

According to the Global Health 50/50 definition, gender refers to the socially constructed norms that impose and determine roles, relationships, and positional power for all people across their lifetime [1026]. Gender-affirming therapy (i.e. testosterone, estrogens, antiandrogens and gonadotropin-releasing hormone analogs) is used in transgender individuals to favorably modulate their secondary sex characteristics in line with gender identity. There is conflicting evidence from small studies of limited duration on the direction or extent of BP changes in transgender individuals receiving gender-affirming therapies [1027]. It is also undetermined whether transgender hormonal treatments are associated with increased CV risk. As for other individuals receiving sex hormone treatment for different purposes (e.g. hormone replacement therapy, oral contraceptive agents), it seems reasonable to recommend that transgender individuals should regularly measure their BP during gender-affirming treatments to ensure persistency of BP normality or, if hypertensive, BP control by treatment [1028].

15.5 Hypertension and ethnicity

15.5.1 Nomenclature and relevance

Given the degree of African and Asian migration to Europe, attention to factors and differences in the management of hypertension related to ethnicity is mandatory. In the United States, the notion is consolidated that in African Americans, hypertension is more common, possibly more severe, more risky for CV and kidney events, and more difficult to control. A major research agenda to improve the hypertensive burden among African-Americans was proposed some years ago [1029]. The intensive search for genetic links to excess ‘African’ hypertension yielded less success than expected [1030,1031], but nevertheless several specific new pathophysiological, clinical and even genetic traits have emerged. With regard to European immigrants, early arterial stiffening (indicating premature vascular aging independent of BP), has been observed in African Americans but only inconsistently in African origin groups in Europe [1032–1035]. Varying degrees of hyperaldosteronism have been found to raise BP disproportionally in people of SSA, African Caribbean and African American origin, probably via underlying salt-sensitivity [1036,1037]. In a recent diagnostic trial for primary aldosteronism, MATCH, 40% of patients were of African origin compared with <20% in the local population [1038]. Somatic (not germ-line) adrenal mutations of the Ca channel were more frequent in African-origin patients, with K+ channel mutations more common in Europeans [1038]. Clearly, the present scarcity of European research needs to be amended and expanded beyond the present general knowledge that, as in USA, African-origin people in Europe have more hypertension associated with greater CV risk than the European ancestry population. So far, some evidence has been obtained that differences may be related to in utero and postnatal development and growth, including overt malnutrition, malaria in pregnancy or epigenetic influences [1039–1042]. Persistent social disadvantages from individual and structural racism, including neighborhood segregation [1043–1045], ethnic disparities or inequalities in access to care are increasingly recognized as driving factors for resulting differences in hypertension and increased CV morbidity and mortality [1046,1047]. Disparities in hypertension control have also been attributed to patient-level rather than provider or system-level factors, both converging to promote therapeutic inertia [1048]. Other studies on extra-European origin migrant communities reflect the USA-derived knowledge that hypertension is more prevalent in European immigrants of heterogeneous African origin than in local populations[1049]. Ethnic and socioeconomic results from Dutch and UK cohorts [1046,1050] have found major risks of CV events (not only stroke but also CAD) in Caribbeans and Asian minorities affected by hypertension.

15.5.2 Management

Lifestyle modifications are vital for reducing BP and risk across the life-span in all hypertensive patients. Differences in BP response to antihypertensive drugs by ancestry have been associated with BP levels rather than ‘race’ [466], and no consistent interethnic factors have been detected in antihypertensive drug pharmacokinetics. BP associations with CV risk have also been broadly consistent across ethnic groupings [1051]. Based on systematic reviews [1052] in people of SSA ancestry, CCB and diuretic monotherapies appear to be more effective than BB and ACEi monotherapies, with combination therapy between major drug classes allowing a substantial group of patients (50– 70%) to reach BP control [1052,1053]. This is similar to the results of the ALLHAT trial in the USA, which included 33% African Americans [566,1054] and found that BP control at 4 years was 63, 60 and 54% with chlorthalidone, amlodipine, and lisinopril-based therapy, respectively. However, 24, 28 and 41% of patients of the corresponding groups received 3 drugs, including the BB atenolol. A recent systematic review and meta-analysis [1052] indicated that, in hypertensive adults of African ancestry, BP reduction was similar across initially administered dual combination therapy containing an RAS inhibitor with a CCB, or a diuretic while with a BB in combination therapy, SBP showed a 3.8 mmHg higher SBP compared with other combinations [1048,1052]. An RCT with African patients from seven SSA countries has also shown an effective BP reduction (ABPM) using a CCB in combination with either a Thiazide or an ACEi [1055].

To date, only the ACCOMPLISH trial (n= 11 506, mean age 68 years) has provided outcome data with initial combination therapy [626]. The trial was not powered for separate analysis of the African American subgroup (12%), However, the combined morbidity/mortality was 8.9% for Thiazide diuretic plus ACEi treatment and 6.6% for CCB plus ACEi treatment, with a significant 35% outcome benefit in African Americans (n= 1414) versus only a 23% benefit in the total study population (N= 11 506).

The above limited evidence suggests that in hypertensive adults of African ancestry, antihypertensive treatment should be largely based on CCBs but also that CCB plus ACEi and CCB plus diuretic combinations can both effectively lower BP, with some suggestion of a greater CV protection by the former combination. A systemic review has shown no different response to antihypertensive drugs in south Asian patients with a relatively high CV risk [1052].

16. HYPERTENSION IN SPECIFIC SETTINGS

16.1 Hypertension disorders in pregnancy

Hypertension disorders affect almost 10% of pregnancies worldwide and are the major cause of maternal, fetal or neonatal morbidity and mortality [1056]. Maternal risks include the following: placental abruption, stroke, pulmonary edema, thromboembolic events, multiple organ failure and disseminated intravascular coagulation. The fetus is at high risk of intrauterine growth retardation (25% of cases of preeclampsia), prematurity (27% of cases of preeclampsia) and intrauterine death (4% of cases of preeclampsia). Neonates are exposed to preterm birth with low birth weight, prolonged high-level neonatal care and postnatal death [1057]. The definition and classification of hypertension disorders in pregnancy are summarized in Table 22 and addressed in detail below.

TABLE 22 - Classification of hypertensive disorders in pregnancy
A. Preexisting (chronic) hypertension
Hypertension either preceding pregnancy or developing before 20 weeks gestation, usually persisting for more than 42 days postpartum, and may be associated with proteinuria.
1. Primary hypertension
2. Secondary hypertension
3. White-coat hypertension
4. Masked hypertension
B. Gestational hypertension
Hypertension develops after 20 weeks gestation and usually resolves within 42 days postpartum.
Transient gestational hypertension
– Usually detected in the clinic but then settles with repeated BP measurements taken over several hours, it is associated with a 40% risk of developing true gestational hypertension or preeclampsia in the remainder of the pregnancy, thus requiring careful follow-up.
Preeclampsia is gestational hypertension accompanied by one or more of the following new-onset conditions at or after 20 weeks gestation:
– Proteinuria (urinary albumin excretion in a 24 h urine sample >0.3 g/day or UACR in a random spot urine sample >30 mg/mmol (0.3 mg/mg)
– Other maternal organ dysfunction
– Acute kidney injury (serum creatinine ≥90 μmol/l; 1 mg/dL)
– Liver involvement (elevated ALT or AST >40 IU/l; 0.67 >μkat/l with or without right upper quadrant or epigastric abdominal pain)
– Neurological complications (e.g. eclampsia, altered mental status, blindness, stroke, clonus, severe headaches, persistent visual scotomata)
– Hematological complications (platelet count <150000/μl, DIC, hemolysis)
– Uteroplacental dysfunction (fetal growth restriction, abnormal umbilical artery Doppler waveform analysis, or stillbirth)
C. Preexisting hypertension + superimposed preeclampsia
Preexisting hypertension associated with any of the above maternal organ dysfunctions consistent with preeclampsia or a further increase in BP with new-onset proteinuria
D. Antenatally unclassifiable hypertension
When BP is first recorded after 20 weeks gestation, and hypertension is diagnosed, reassessment is necessary at or after 42 days postpartum. If hypertension resolves, it should be reclassified as gestational hypertension, whereas if hypertension persists, it should be reclassified as preexisting hypertension.

16.1.1 Definition and classification of hypertension in pregnancy

The definition of hypertension in pregnancy is based on office BP values, i.e. SBP ≥ 140 mmHg and/or DBP ≥ 90 mmHg [1057,1058] and is classified as mild (140–159/90 –109 mmHg) or severe (≥ 160/110 mmHg), at variance from the general hypertension grading (Table 1). HDPs are further classified according to the onset of hypertension in pregnancy. Hypertension known before pregnancy or present in the first 20 weeks can be distinguished as preexisting (chronic) hypertension (primary or secondary), WCH (elevated office and normal out-of-office BP), and MH (normal office and elevated out-of-office BP). Hypertension that develops de novo at or after 20 weeks of pregnancy is classified as gestational hypertension, transient gestational hypertension, or preeclampsia de novo or superimposed on preexisting hypertension. Prepregnancy or early first-trimester BP measurements should be available to avoid misclassification of hypertensive disorders in pregnancy. Early second-trimester BP measurements in women without previous BP measurements should be interpreted with caution because of the physiological second-trimester BP fall. However, women with hypertensive BP levels after 20 weeks and unknown BP levels before 20 weeks should be managed as those with gestational hypertension. In these women with prepartum pregnancy unclassifiable hypertension, BP reassessment 6 weeks postpartum will help to distinguish preexisting hypertension from gestational hypertension. Transient gestational hypertension (diagnosed at the clinic after 20 weeks) is characterized by BP normalization after consecutive BP measurements over several hours. However, almost 40% of women with transient gestational hypertension develop sustained gestational hypertension [1059].

The present guidelines adopt the broader definition of preeclampsia previously endorsed by the International Society for the Study of Hypertension in Pregnancy [1058]. Accordingly, preeclampsia is a gestational hypertension in the presence of one or more of the following new-onset conditions at or after 20 weeks of gestation: (i) significant proteinuria (ACR ≥ 30 mg/g or albuminuria >300 mg/24 hour), (ii) maternal organ dysfunction, [i.e. acute kidney injury (serum creatinine ≥ 1 mg/dl; 90 μ/l); liver injury (elevated transaminases >40 UI/l; 67 μkat/l) with or without right upper quadrant or epigastric pain; neurological manifestations (convulsions, altered mental status, blindness, scotoma or headache); hematological manifestations (platelet count <150 000/μl, disseminated intravascular coagulation, hemolysis)] and (iii) uteroplacental dysfunction (i.e. fetal growth restriction, abnormal umbilical artery Doppler waves or stillbirth). The combination of hemolysis, thrombocytopenia and elevated transaminases defines the HELLP syndrome and, therefore, additional features of preeclampsia should be evaluated. Clinicians should always consider preeclampsia a serious disease with rather unpredictable consequences. In clinical practice, it is not recommended anymore to use the previous classification of preeclampsia based on clinical features such as mild or severe.

Among women with preexisting hypertension, almost 25% will develop superimposed preeclampsia [1060]. In these women, the diagnosis is made when a de novo development of proteinuria is detected, or other maternal organ dysfunctions develop after 20 weeks. This is usually associated with an abrupt or progressive BP elevation.

16.1.2 Blood pressure measurement in pregnancy

During pregnancy, BP should be measured in the sitting position (or in the left lateral recumbent position during labor) with an appropriately sized arm cuff at heart level using the manual auscultatory method and Korotkoff phase V for DBP [1061]. Manual auscultation remains the gold standard for BP measurement in pregnancy, because automated devices tend to under-record the BP and are unreliable in severe preeclampsia. Only devices validated specifically for pregnancy should be used [1061]. ABPM is superior to office BP measurement for predicting pregnancy outcomes [1062], and ABPM devices recommended for use in pregnancy are more accurate than those used for office measurement or HBPM. ABPM helps to avoid unnecessary treatment in WCH and is useful in the management of high-risk pregnant women with hypertension and those with diabetic or hypertensive nephropathy. According to the BUMP-1 trial [1063], among pregnant individuals at higher risk of preeclampsia, HBPM did not lead to earlier clinic-based detection of hypertension. However, the BUMP-1 trial also suggests that HBPM and office BP measurements may be used alternatively or in complement to diagnose hypertensive disorders during pregnancy in women at risk of preeclampsia. In the BUMP-2 trial [1064], HBPM was not associated with better BP control among pregnant individuals with preexisting or gestational hypertension compared with scheduled office BP measurements. Again, the BUMP-2 trial suggests that BP control according to HBPM can be used alternatively or complementarily to office BP measurements because both methods achieved similar rates of BP control.

16.1.3 Laboratory examinations in pregnancy

Basic laboratory investigations are recommended for monitoring pregnant hypertensive women, including urine analysis, blood count, hematocrit, liver enzymes, serum creatinine and serum uric acid (increased in clinically evident preeclampsia). Hyperuricemia in hypertensive pregnancies identifies women at increased risk of adverse maternal and fetal outcomes [1065]. All pregnant women should be assessed for proteinuria in early pregnancy to detect preexisting renal disease and, in the second half of pregnancy, to screen for preeclampsia. However, the presence of proteinuria is no longer a ‘sine qua non’ criterion for the diagnosis of preeclampsia. Occasionally, proteinuria may anticipate a subsequent rise of BP in the natural course of preeclampsia. A dipstick test of at least 1+ should prompt evaluation of ACR in a single spot urine sample, and a value of less than 30 mg/g can reliably rule out proteinuria [1066,1067]. Other investigations to be considered are: (i) renal ultrasound if serum creatinine or any of the urine testing is abnormal and (ii) Doppler ultrasound of uterine and umbilical arteries (performed after 20 weeks of gestation) to detect those at a higher risk of gestational hypertension, preeclampsia and intrauterine growth retardation.

16.1.4 Prediction and prevention of preeclampsia

Women at high or moderate risk of preeclampsia should be advised to take 100–150 mg of aspirin daily (at bedtime), preferably before 16 weeks and ideally from weeks 11 to 14 until 36 weeks of gestation [1068–1071].

High risk of preeclampsia includes any of the following:

  • 1. Hypertensive disorders during a previous pregnancy
  • 2. Chronic hypertension
  • 3. Chronic kidney disease
  • 4. Type 1 or type 2 diabetes mellitus
  • 5. Autoimmune diseases such as systemic lupus erythematosus or antiphospholipid syndrome
  • 6. Assisted reproductive therapy in the current pregnancy

Moderate risk of preeclampsia includes two or more of the following risk factors:

  • 1. Nulliparity
  • 2. Age 40 years or older
  • 3. Pregnancy interval of more than 10 years
  • 4. BMI of 35 kg/m2 or more at the first visit
  • 5. Family history of preeclampsia
  • 6. Multifetal pregnancy

Beyond clinical and ultrasonographic parameters, several laboratory markers have been tested during early pregnancy for the prediction of preeclampsia:

  • 1. Angiogenic factors [endoglin, PIGF, soluble fms-like tyrosine kinase-1 (sFlt-1) and sFlt-1/PIGF ratio]
  • 2. Pregnancy-associated plasma protein A (PAPP-A) in association with clinical (e.g. BP, maternal risk factors) and ultrasonographic characteristics (e.g. uterine artery Doppler) [1072].

However, more studies are desirable to refine the role of the above markers separately or in combination with clinical characteristics for predicting preeclampsia. Although prospective evidence is limited [1073,1074], a sFlt-1/PIGF ratio of at least 38 may be considered to exclude the development of preeclampsia when clinically suspected [1074].

16.1.5 Lifestyle interventions

Unless contraindicated, aerobic exercise (three to four times per week for 30– 60 min sessions until delivery) should be recommended in pregnant women to maintain ideal body weight and reduce adverse pregnancy outcomes, including hypertensive disorders [1075–1077]. In addition, calcium supplementation at a dose of at least 1 g/day may be considered to reduce preeclampsia risk in women with low calcium intake (i.e. <600 mg/day) [1078]. Finally, although salt restriction is not advised to reduce HDP, it is reasonable that women with preexisting hypertension should continue pursuing a limited salt intake diet [1070].

16.1.6 Clinical management of hypertension in pregnancy

16.1.6.1 Mild preexisting essential hypertension

During the first trimester, all RAS blockers, i.e. ACEis, ARBs or direct renin inhibitors should be stopped. Clinicians should decide on an individual basis whether drug treatment during fetal organogenesis (until week 16) overcomes the risk of fetal drug effects because any drug may be potentially hazardous during the first trimester, including alpha-methyldopa [1079]. The decision to use or discontinue antihypertensive treatment during the first and early second trimester may be individualized based on (i) the prepregnancy untreated BP levels, (ii) the early on-treatment first trimester BP values, the presence of HMOD and (iv) the BP values after a potential short-term trial of antihypertensive treatment withdrawal in selected cases. In the early first trimester, for a woman with office BP levels of <130/80 mmHg, BP-lowering treatment may be discontinued or de-escalated under a careful FU of BP levels until week 16. Antihypertensive treatment should restart in case of BP >140/90 mmHg at any gestational age. In preexisting hypertension, the absence of mild antihypertensive treatment during the early second trimester may prevent a profuse BP drop, potentially accompanied by miscarriage because of the physiological BP reduction during this pregnancy period.

In the large CHIPS [1080] and CHAP [1081] trials, tight versus less tight DBP control or drug treatment versus placebo in women with preexisting hypertension was more beneficial and carried no harm. Furthermore, in a post hoc analysis of the CHIPS trial [1080], the BP-lowering effect was beneficial for the primary outcome of the CHAP study, i.e. severe features, medically indicated preterm birth at less than 35 weeks gestation, placental abruption or fetal/neonatal death. The pregnancy-related composite outcome was reduced by 35% in the CHIPS and by 18%, in the CHAP trial; both studies indicating a reduction of severe preeclampsia. However, the CHIP trial observed a small-for-gestational age newborn outcome increase, albeit this was not found in the CHAP trial. This maintains alive the problem of the safety margin guaranteed by lower BP values. In light of the on-treatment BP values observed in CHIPS and CHAP trials (133/85 and 129/79 mmHg, respectively), we suggest that the threshold for BP-lowering treatment initiation or potentiation may be ≥ 140/90 mmHg and that in general, intensified BP-lowering should not be pursued because of the risk of fetal hypoperfusion. Labetalol and alpha-methyldopa are the first-choice drugs for BP control in women with preexisting hypertension [1071,1080,1081]. An alternative agent to use is extended-release nifedipine [1071]. The use of labetalol is controversial and not a choice in several countries in which it was removed from market 30 years ago, because of hepatotoxicity, which may also occur when used in pregnancy [1082].

16.1.6.2 Mild gestational hypertension

Although the CHIPS trial [1080] included a limited number of women with gestational hypertension, secondary analyses did not indicate a differential outcome effect between women with gestational and preexisting hypertension, both for primary and secondary outcomes. A treatment initiation at values ≥ 140/90 mmHg appears to be reasonable, while a DBP reduction to <80 mmHg is not recommended. The same drugs recommended for preexisting hypertension (see above) can be used in women with gestational hypertension.

16.1.6.3 Preeclampsia

All women with preeclampsia should be hospitalized and carefully monitored at first diagnosis. A diagnosis of preeclampsia at or after 37 weeks of gestation underscores the need for hypertension control and prompt delivery. Clinically stable women with preeclampsia before 37 weeks of gestation can be managed on an outpatient basis. However, despite optimal antihypertensive treatment, delivery is indicated even before 37 weeks, whenever hypertension remains severe. Delivery induction before 37 weeks is also recommended with (i) emerging maternal (neurological, hematological or CV) manifestations or (ii) a nonreassuring fetal status [1058].

The management of hypertension in preeclamptic women with mild or severe hypertension does not differ from women without preeclampsia (see above sections), although evidence is limited. Preeclampsia with severe features (severe hypertension with or without proteinuria, any hypertension grade with neurological, hematological, or cardiovascular complications, liver dysfunction or renal dysfunction) should be managed with MgSO4 infusion (and delivery) to prevent eclampsia. Infusion of MgSO4 for the 24 h postpartum seems reasonable for prevention purposes [1083], and MgSO4 remains the treatment of choice for eclamptic seizures. Hypertension control can be achieved by labetalol (unless contraindicated) alone or with the combination of labetalol, nifedipine extended-release and/or alpha-methyldopa.

16.1.6.4 Severe hypertension

In severe hypertension, hospitalization is mandatory to allow gradual BP reduction to <160/105 mmHg and exclude preeclampsia. Continuous cardiotocographic monitoring is also mandatory [1058]. The selection of antihypertensive drugs and the route of administration depends on (i) the initial diagnosis, (ii) the expected delivery time and (iii) the presence or absence of preeclampsia and also preferences and experience of attending physicians. A recent comprehensive network meta-analysis indicated that nifedipine could be recommended as a strategy for BP management in pregnant women with severe hypertension and that labetalol and hydralazine showed in fact limited efficacy [1084]. However, in cases of preeclampsia with severe features, persistent severe hypertension or recurrent severe hypertension despite orally administered agents, i.v. application of labetalol or urapidil should be used before, during and frequently after delivery. In case of preeclampsia without severe features or severe hypertension without preeclampsia, an effective and gradually escalated antihypertensive multiple drug regimen should be used to lower BP to target [1080], while before delivery, hydralazine should be avoided because of its association with more adverse perinatal effects than other drugs. Hydralazine should be reserved to cases of unavailability of labetalol or urapidil, failure to reduce BP, II or III degree AV block, severe HF, asthma, bradycardia or severe postpartum hypertension. Sodium nitroprusside should be used as the last resource because of an increased risk of fetal cyanide poisoning with prolonged administration. When preeclampsia is associated with pulmonary edema, the drug of choice is nitroglycerin, given as an i.v. infusion of 5 mg/ min and a gradual increase every 3– 5 min to a maximum dose of 100 mg/min [1085]. In a pregnant woman with severe hypertension living in a rural area away from a maternity hospital, 10 mg short-acting nifedipine may be administered orally and a second dose should be given after 1 h if severe hypertension persists. Sublingual short-acting nifedipine is contraindicated.

16.1.6.5 Preexisting secondary hypertension

Women with known preexisting hypertension should receive preconception counseling, including ruling out secondary hypertension causes. Renal Doppler ultrasounds should be performed in all women with hypertension planning a pregnancy. In women diagnosed with fibromuscular dysplasia (FMD) before pregnancy, further evaluation of other vascular beds should be performed to exclude any additional arterial damage [340]. Achievement of optimal BP control and, if indicated, renal artery revascularization are recommended before conception, because undiagnosed and untreated FMD can increase the risk of pregnancy-related complications [1086].

Women with primary aldosteronism should be discouraged from pregnancy because of the competitive antagonism of progesterone and aldosterone in excess at the level of the mineralocorticoid receptors [1087]. However, close laboratory work-up should be performed in women with known preconception hyperaldosteronism or clinical suspicion of the disease in early pregnancy. After the second trimester, eplerenone on top of the usual BP-lowering treatment may be considered for uncontrolled hypertension with or without hypokalemia [1088]. The drop in progesterone levels postpartum may increase BP and aggravate hypokalemia [1087,1088].

Pheochromocytoma in pregnancy is one of the most life-threatening conditions for the mother and fetus. Although extraordinarily rare, with a frequency of 0.002% of all pregnancies, this disease is notorious for its devastating consequences [1089]. As in nonpregnant patients, the signs and symptoms are quite variable and poorly specific, with hypertension being one of the most dominant signs. If undiagnosed, maternal and fetal mortality is around 50%. Early detection and proper treatment during pregnancy decrease maternal and fetal mortality to <5 and <15%, respectively. For the biochemical diagnosis, plasma or urinary metanephrines are the tests of choice, as they have the highest sensitivity and lowest falsely negative diagnostic value. For reliable localization, MRI is the most suitable technique with a sensitivity over 90%. When a pheochromocytoma is diagnosed during pregnancy, laparoscopic adrenalectomy should be performed after 10– 14 days of drug pretreatment as in nonpregnant patients (alpha-adrenoreceptor blockade combined with beta-adrenergic blockade some days later). If the pheochromocytoma is diagnosed in the third trimester, the patient should be managed until the fetus is viable, using the same drug regimen as for the surgical preparation. Caesarian section with tumor removal in the same session or at a later stage is preferred as vaginal delivery may be associated with higher mortality.

HDP are observed in almost half of pregnant women with CKD. Therefore, it is important to know the degree of CKD, level of eGFR or degree of proteinuria, before pregnancy rather than the underlying cause [1090]. Women without significant proteinuria, normal BP levels at early pregnancy and mild renal impairment usually experience an uneventful pregnancy course. By contrast, women with moderate or more advanced CKD are at increased risk of fetal complications as well as maternal complications and deterioration of the already impaired kidney function [1091]. A woman with a GFR of less than 40 ml/min per 1.73 m2 and proteinuria of more than 1 g/day should be considered at very high risk for pregnancy and kidney outcomes, including kidney replacement therapy.

16.1.7 Blood pressure during puerperium

Postpartum hypertension is common during the first week. Also, in women with a normotensive pregnancy, a BP elevation during the first-day postpartum is usually associated with (i) the use of vasoactive drugs to favor uterine contraction (oxytocin, methergine), (ii) blood transfusions, (iii) the physiological uterine ‘auto-transfusion phenomenon’ or (iv) an excessive fluid intake. In women with preeclampsia, a reduced diuresis during 12– 36 h postpartum is observed because of a delayed fluid redistribution associated with a greater colloid osmotic pressure drop compared to a normal pregnancy [1092]. In a small randomized placebo-controlled trial on women with hypertension in pregnancy, administration of furosemide 20 mg daily during the first 5 days postpartum prevented 1 woman out of 13 from developing postpartum hypertension [1093]. However, a wide use of furosemide postpartum needs confirmation from larger studies.

During puerperium, BP levels usually normalize within the first 6 weeks in women with gestational hypertension or preeclampsia. By contrast, women with preexisting hypertension or superimposed preeclampsia perpetuate elevated BP values beyond the 6 weeks of puerperium. A further unusual postpartum hypertension phenotype is the so-called ‘late postpartum hypertension’ phenotype, which appears 6 months after delivery and resolves several months later [1094]. The pathogenesis of this condition is unknown, but one possibility is that the return of postpartum menses increases BP through a spillover of the excess of progesterone and an activation of mineralocorticoid receptors. This is similar to the Geller syndrome, which exhibits an exacerbated hypertension in the third trimester of pregnancy [1094,1095].

All antihypertensive agents used during pregnancy may be used during puerperium to achieve BP control. However, the use of ACEis in the postpartum period should be reserved for women with cardiorenal comorbidities and is, thus, not recommended in healthy women with hypertensive disorders during puerperium. Methyldopa should be used with caution because of the risk of postpartum depression.

16.1.8 Postpartum hypertension and breastfeeding

Antihypertensive drugs taken by the nursing mother are excreted into breast milk, mostly in very low concentrations. Proper information on prescribable drugs in breastfeeding women is important [1057,1096]. Nifedipine and verapamil are considered compatible with breastfeeding. Although diuretics are not contraindicated, they may be associated with reduced milk production. Similarly, alpha-methyldopa is compatible with breastfeeding, although it is not a drug of first choice during puerperium because it increases the risk of postpartum depression. ACEis are compatible with breastfeeding and can be used in women with HDP and underlying CVD or CKD. ARBs are not currently recommended in lactating women because of a limited safety evidence [1096].

16.1.9 Risk of recurrence of hypertensive disorders in a subsequent pregnancy

Women experiencing hypertension in their first pregnancy are at increased risk of hypertension and hypertensive disorders in a subsequent pregnancy. The earlier the onset of hypertension in the first pregnancy, the higher the risk of recurrence in a subsequent pregnancy.

16.1.10 Long-term cardiovascular consequences of hypertensive disorders in pregnancy

Several registries [1097,1098] have demonstrated that pregnant women with hypertension and hypertensive disorders are at an increased risk for future CV events [1097,1098]. Compared to women with normotensive pregnancies, the risk shows a multifold increase in pregnant hypertensive women [33,1099], and the increase also involves the risk of developing future sustained hypertension[48,1100]. A meta-analysis of cohort studies showed that preeclampsia with more severe features was associated with a greater extent of future disease compared with preeclampsia with less severe features [49]. A genome-wide genetic association study using Mendelian randomization provided genetic evidence supporting an association between HDP and higher risk of CAD and stroke, which is only partially mediated by cardiometabolic factors. This study supports classification of HDPs as risk factors for CVD [1101]. In women who experienced HDP, lifestyle modifications are indicated to reduce the risk of complications in subsequent pregnancies as well as to reduce CV risk in general [33,1099]. Annual visits by the primary care physician, frequent BP measurements and CV risk assessment are recommended.

Hypertension management in pregnancy

figure43

16.2 Hypertensive urgencies and emergencies

The proper management of patients who come to the Emergency Department for BP elevation faces a number of difficulties. A major challenge for the physician is to identify and discriminate the patients at immediate risk of CV or kidney complications from those in whom the BP elevation does not carry any immediate risk for health. While the former patients require immediate attention and treatment, the latter are in no need of hospitalization or even treatment. FU and treatment of patients after discharge from the Emergency Department is still poorly defined and often inadequate [1102,1103].

16.2.1 Definitions of hypertensive urgencies and emergencies

Hypertension emergencies are conditions in which severe hypertension (grade 3) is associated with acute symptomatic HMOD. Hypertension emergencies can be life-threatening and require immediate intervention to lower BP, usually with intravenous (i.v.) therapy [1103]. The rate of the increase in BP may be at least as important as the absolute BP level in determining the clinical severity of the situation and the magnitude of organ injury [1103,1104]. Typical clinical presentations of a hypertension emergency are:

  • 1. Severe hypertension associated with conditions that need intensified BP management: acute stroke (hemorrhagic or ischemic/thromboembolic), aortic aneurysm or dissection, acute HF, acute coronary syndrome and kidney failure. These emergency conditions are compatible also with a relatively modest BP increase, which is sufficient to precipitate organ failure.
  • 2. Hypertension caused by phaeochromocytoma or exogenous sympathomimetics substances (e.g. substance abuse). Ingestion of sympathomimetic drugs such as meta-amphetamine or cocaine may precipitate acute and severe BP increases that may result in hypertension emergencies when there is evidence of acute HMOD.
  • 3. Severe forms of HDP, including preeclampsia/eclampsia with a HELLP syndrome (see Section 16.1) [1103].

Malignant hypertension with or without thrombotic microangiopathy or acute kidney failure is a hypertensive emergency characterized by small artery fibrinoid necrosis in the kidney, retina and brain. There might be also funduscopic changes (flame hemorrhages and papilledema), microangiopathy, disseminated intravascular coagulation, encephalopathy (15% of cases) or acute HF [1105]. Whether this traditional definition fully corresponds to the extent of the microvascular damage is still a matter of debate [1106]. The term ‘malignant’ reflects the very poor prognosis of this condition if untreated [1105,1107–1109]. The Birmingham, Bordeaux and Amsterdam malignant hypertension registries have shown that malignant hypertension is rising in Europe [1103,1105,1110]. The emergency symptoms depend on the organs affected and may include headache, visual disturbances, dizziness and other neurological deficits as well as chest pain and dyspnea. In patients with hypertensive encephalopathy, the presence of somnolence, lethargy, tonic clonic seizures and cortical blindness may precede loss of consciousness. Focal neurological lesions are rare and should raise the suspicion of acute stroke [4,1103,1104,1111,1112]. An acute stroke, especially intracerebral hemorrhage, is associated with severe hypertension and has often been included among the hypertensive emergencies. However, it is an emergency that requires specific strategies to govern the BP reduction, because of the complex effects that its magnitude, speed and relationship with the existing BP level can have on the ischemic brain area surrounding the core brain lesion (see Section 17.5). This condition should be handled by specialized (stroke) units.

The term ‘hypertension urgency’ has been used to describe severe hypertension in patients in whom there is no evidence of acute HMOD [1102,1108–1110]. The burden of hypertensive urgencies is not well defined mainly because of the different criteria used for the definition of this condition. Furthermore, the ambiguity of the term ‘hypertension urgency’ versus the so called ‘hypertensive crisis’ has influenced epidemiological data [1102,1108–1110].

For patients with a suspected hypertension emergency, the diagnostic work-up is shown in Table 23.

TABLE 23 - Diagnostic work-up of hypertension emergencies and urgencies
Common tests
Fundoscopy
ECG 12 leads
Hemoglobin, platelet count, fibrinogen, peripheral smear
Creatinine, eGFR, electrolytes, LDH, haptoglobin
UACR, urine microscopy for red blood cells, leucocytes and/or casts
Pregnancy test in women of child-bearing age
Specific tests
Troponin, (suspected HF and/or acute coronary syndrome) NT-proBNP
Chest X-ray or ultrasound (pulmonary congestion and fluid overload)
Echocardiography (heart failure, acute ischemia, aortic dissection)
CT angiography of thorax and/or abdomen in suspected aortic disease (aortic dissection)
CT or MRI brain (nervous system involvement)
Kidney ultrasound (renal impairment or suspected renal artery stenosis)
Urine drug collection (cocaine or methamphetamine use)

16.2.2 Hospital work-up, treatments and follow-up

Hypertensive emergencies, including BP 170/110 mmHg in a pregnant woman, should be hospitalized. Except for acute BP-lowering in stroke, there are no RCTs on the management of these conditions. It should first be established, which organs are affected to determine whether (i) they require any specific intervention other than BP-lowering and (ii) there is a precipitating cause for the acute rise in BP that might affect the treatment plan (e.g. pregnancy). Then a decision should be made on the timescale and magnitude of the BP-lowering as well as on the type of drug treatment that might be appropriate. Intravenous treatment with a drug that has a short half-life is ideal to allow careful titration of the BP response, keeping the patient in a close clinical area under continuous hemodynamic monitoring (Table 24). Rapid uncontrolled BP-lowering is not recommended [4] and, thus, low initial doses with cautious dose uptitration should be used. Oral therapy with ACEis, ARBs or BBs (at low initial doses and cautious upward titration) is sometimes effective in malignant hypertension because the RAS may be activated by the associated kidney ischemia [1105]. Recommended drugs are shown in Table 25, and further details can be found in dedicated publications [4,1103,1113].

TABLE 24 - Hypertensive emergencies requiring immediate BP-lowering with i.v. drug therapy
Clinical presentation Timing and BP target First-line treatment Alternative
Malignant hypertension with or without acute renal failure Several hours
Reduce MAP by 20–25%
Labetalola
Nicardipine
Nitroprusside
Urapidil
Hypertensive encephalopathy Immediately reduce MAP by 20–25% Labetalola
Nicardipine
Nitroprusside
Acute coronary event Immediately reduce SBP to <140 mmHg Nitroglycerine
Labetalola
Urapidil
Acute cardiogenic pulmonary edema Immediately reduce SBP to <140 mmHg Nitroprusside or nitroglycerine (with loop diuretic) Urapidil
(with loop diuretic)
Acute aortic dissection Immediately reduce SBP to <120 mmHg and heart rate to <60 bpm Esmolol AND nitroprusside or nitroglycerine or nicardipine Labetalola or metoprolol
Eclampsia and severe preeclampsia/HELLP Immediately reduce SBP to <160 mmHg and DBP to <105 mmHg Labetalola or nicardipine and magnesium sulphate Consider delivery
aNot available in several countries.

TABLE 25 - Drug types, dose and characteristics for treatment of hypertension emergencies
Drug Onset of action Duration of action Dose Contraindications Adverse effects
Esmolol 1 min 10–30 min 0.5–1 mg/kg i.v. bolus; 50–300 μg/kg/min i.v. infusion Second-degree or third-degree AV block, systolic heart failure, asthma, bradycardia Bradycardia
Metoprolol 1–2 min 5–8 h 2.5–5 mg i.v. bolus over 2 min; may repeat every 5 min to a maximum dose of 15 mg Second-degree or third-degree AV block, systolic heart failure, asthma, bradycardia Bradycardia
Labetalola 5–10 min 3–6 h 10–20 mg i.v. bolus in 1 min; incremental doses ≥20 mg may be administered i.v. at 10 min intervals (max 80 mg) or 1–3 mg/min i.v. infusion until goal BP is reached Second-degree or third-degree AV block; systolic heart failure, asthma, bradycardia Bronchoconstriction, fetal bradycardia
Fenoldopam 5–15 min 30–60 min 0.1–0.3 μg/kg/min i.v. infusion, increase every 15 min with 0.1 μg/kg/min increments until goal BP is reached Caution in glaucoma
Clevidipine 2 min 10 min 1–2 mg/h i.v. infusion, increase every 2 min with 2 mg/h until goal BP, then titrate by smaller increments every 5–10 min Headache, reflex tachycardia
Nicardipine 5–15 min 4–6 h 5–15 mg/h i.v. infusion, starting dose 5 mg/h, increase every 15–30 min with 2.5 mg until goal BP, maximum 15 mg/h Liver failure Headache, reflex tachycardia
Nitroglycerine 1–5 min 5–10 min 5–200 μg/min i.v. infusion, 5 μg/min increase every 5 min Headache, reflex tachycardia
Nitroprusside Immediate 1–3 min 0.3–0.5 μg/kg/min i.v. infusion, increase by 0.5 μg/kg/min every 5 min until goal BP (maximum dose 10 μg/kg/min) Liver/kidney failure (relative) Cyanide intoxication
Enalaprilat 5–15 min 4–6 h 0.62–1.25 mg i.v. bolus given over 5 min every 6 h History of angioedema
Urapidil 3–5 min 4–6 h 12.5–25 mg i.v. bolus;
5–40 mg/h as continuous infusion
Clonidine 30 min 4–6 h 0.2–0.5 μg/kg/min i.v. Sedation, rebound hypertension
Phentolamine 1–2 min 10–30 min 1–5 mg i.v. bolus or continuous i.v. infusion at a rate of 0.5–20 μg/kg/min Tachyarrhythmia, chest pain
aNot available in serveral countries.

Patients with hypertensive urgencies do not usually require hospitalization. However, they require BP reduction, which can be obtained by oral administration of antihypertensive drugs, aimed at lowering BP gradually over 24–48 h. Oral treatment may include reinstitution or intensification of previous treatment or starting new treatment. DHP-CCBs are suggested as first choice in an untreated patient as they have few or no contraindications and do not interfere with the diagnostic work-up for secondary hypertension. Sublingual, rapidly acting, administration of nifedipine should be avoided because the degree of BP decrease cannot be anticipated and may often be too fast and larger than desirable [4,1103]. Clinical surveillance in a short-stay observation unit is usually appropriate before discharge [1114]. As BP may remain elevated after Emergency Department discharge [1115], these patients need subsequent office and out-of-office BP measurements.

The survival of patients with hypertension emergencies has improved strikingly over past decades [1105], but these patients remain at high risk [4,1105,1108–1112,1116]. After discharge from hospital, when BP has reached a safe and stable level on oral therapy, an early FU and then frequent (at least monthly) medical visits and supervision by a hypertension specialist or a specialized hypertension center are recommended [4,1117,1118].

16.2.3 Blood pressure in the emergency department

BP measurements in the emergency department usually do not follow guidelines and may, thus, be inaccurate [4]. BP elevation is common, and hypertensive BP values in 48% of all patients referred to the Emergency Department during a 1-year period have been reported [1119]. However, the clinical significance of a BP elevation in the Emergency Department is not entirely clear, and available evidence is not univocal. In many patients, even a marked BP elevation is restored to normal when pain, distress and anxiety are relieved, thus originating from an alerting reaction [1120,1121]. However, some studies have not shown a close relationship of BP with pain, stress or anxiety in the Emergency Department. In addition, although in general BP values in the Emergency Department are not recommended for stratification of CV risk, an association of Emergency Department-measured BP and CV outcomes has been reported [1117]. In all patients, previous intake of some medications (e.g. NSAIDs) or other substances that may cause a BP elevation or oppose the BP reduction by treatment should be searched for. If BP is not severely elevated and the clinical picture does not suggest a hypertensive emergency, unattended BP measurements should be performed with the patient alone in a separate quiet room because unattended BP measurements are usually not accompanied by an alerting reaction [95]. BP measurements should be repeated at intervals for at least an hour. Evidence is available that in 30% of the patients, BP decreases to grade 2 or lower values in 30 min [1114].

16.3 Perioperative hypertension and its management

Because of its high prevalence and association with age, hypertension is an extremely common condition in patients undergoing surgery, in several cases with no awareness of the high BP status by the patient. A BP elevation at the time of a surgical intervention is not an innocent association because severe hypertension has been found to be accompanied by an increased incidence of surgery-related complications, including bleeding [1122]. Thus, at the presurgery visits, accurate office BP measurements and implementation of the other steps required to clinically characterize hypertension are mandatory. Although no study has defined the BP threshold on which to base the decision, it seems appropriate to suggest not to avoid or postpone surgery in untreated patients with grade 1 hypertension or in those in whom treatment has achieved controlled or almost controlled BP values. In untreated patients with grade 2 or 3 hypertension or in badly uncontrolled treated patients, surgery should be deferred until BP control is reached. Except for emergency surgical interventions, this should be the rule when SBP is 180 mmHg or DBP ≥ 110 mmHg [488,1123]. Decision about deferral may find help from out-of-office BP measurements (ABPM or HBPM) because a white-coat effect may be frequent in patients with scheduled surgery [1124]. Data are not entirely univocal, but it seems appropriate not to discontinue antihypertensive drugs during the days before surgery, although discontinuation has been advocated for fear of arrhythmias, volume depletion, electrolyte alterations, bradycardia and intra-operative and postoperative hypotension. Discontinuation should also be avoided in the case of BB treatment preceding noncardiac or cardiac surgery, as preexisting BB treatment has been reported to have a protective effect [1125], while abrupt discontinuation may lead to rebound heart rate and perhaps BP elevations [1126,1127]. If for any reason BBs are discontinued, transdermal clonidine may attenuate the rebound phenomenon. Continuation of antihypertensive treatment during the perioperative period may also minimize BP fluctuations, which in a RCT that enrolled patients undergoing abdominal surgery was found to be associated with surgery-related complications [1128].

Hypertension may occur during induction of anesthesia, the intubation maneuvers, the operation or in the early postanesthesia period. Induction of anesthesia generates pain and stress, and intubation evokes reflexes that may lead to cardiac and vascular sympathetic activation, increasing the odds of a rise in BP and heart rate as well as of excessive BP fluctuations and arrhythmias [1122]. BP increases may also be favored by i.v. application of vasopressor drugs during surgery. On the other hand, induction and maintenance of anesthesia may also cause vasodilatation, volume depletion and BP reduction, having blood loss as an adjunctive factor. Both intraoperative hypotension and hypertension may have an unfavorable effect on perioperative complications, also because of impairment of blood flow autoregulation, especially in old patients [1129]. No consensus is available on thresholds and targets to guide intraoperative BP management. As mentioned above, throughout the perioperative period, the BP management should focus on keeping BP within a safety range, avoiding large BP fluctuations and in general pursuing hemodynamic stability. During the postoperative period, hypertension may be present during the first 2 h and can usually disappear in a few hours after resumption of antihypertensive treatment. Oral BBs may be restarted immediately. Oral diuretics and RAS blockers can be restarted within 48 h after surgery, after assessing the BP level and fluid status, and according to the type of surgical intervention (cardiac versus noncardiac).

Perioperative hypertension and its management

figure44

17. HYPERTENSION IN ESTABLISHED CARDIOVASCULAR DISORDERS

17.1 Coronary artery disease

Hypertension is one of the major risk factors for CAD in its various clinical aspects (myocardial infarction, angina pectoris, acute coronary syndrome and chronic coronary syndromes), together with other risk factors including smoking, dyslipidemia and type-2 diabetes [531,1130]. There is a linear correlation between BP levels and the risk of CAD within a wide range of BP values, starting from 110 to 115 mmHg SBP and 70 to 75 mmHg DBP [35,1131]. Hypertension explains approximately 25% of the risk of myocardial infarction at a population level. The presence of CAD classifies a patient at a very high CV risk, even if SBP is below 140 mmHg or DBP below 90 mmHg. A large number of outcome-based RCTs and their meta-analyses have shown not only that BP-lowering treatment significantly reduces the risk of CAD (although the relative size of the reduction is less than that of stroke and HF [1132]) but also that the benefit extends to overall CV outcomes [1133]. In several outcome-based RCTs on antihypertensive treatment, CAD patients were a large fraction of the recruited patients and had an initial office SBP ≥ 140 mmHg or DBP ≥ 90 mmHg. Thus, antihypertensive treatment should be initiated at these BP levels [1134]. Whether lower BP thresholds for treatment should be considered has been a matter of debate for years because studies showing CV protection by BP reductions from baseline BP values <140/90 mmHg (usually in the high-normal BP range) have been mostly obtained in studies in which patients were already on background antihypertensive treatment [559]. However, there is evidence from a meta-analysis of RCTs that BP-lowering treatment in largely untreated patients with high-normal BP did not lower the risk of events if their CV risk was below the very high risk category but it reduced the incidence of stroke (and stroke plus CAD) if the risk was very high due to a history of CV events. The previous event was mainly a myocardial infarction [493], which makes CAD patients candidates for antihypertensive drug treatment when BP values are in the high-normal range. In this meta-analysis, the benefit was obtained with an SBP reduction of a few mmHg, which means that it was established at an average SBP target close to or less than 130 mmHg. This finds confirmation in a number of, although not in all, dedicated trials in CAD or in trials including CAD patients, which have shown that in these patients, outcomes further decrease if target BP values within the range between 120 and 129 mmHg or 70 and 79 mmHg are achieved [1135]. Thus, the present guidelines recommend these values as the target to pursue in CAD patients, based on RCTs performed in patients in whom a history of CAD was largely predominant [524,1136]. As already mentioned (see Sections 9 and 10), it should not be forgotten that CAD patients are almost invariably under treatment with BP-lowering drugs such as RAS blockers, BBs or CCBs, which are part of the CAD treatment strategy, independently from their BP values and for benefits regarded as unrelated to BP reductions [559,1136]. Thus, at a practical level, the question is not whether to give BP-lowering drugs to CAD patients with a high-normal BP, but to modulate their number and doses to reduce BP to the recommended target. Post hoc analyses of trials or prospective registries exclusively or largely based on CAD patients have not infrequently reported a J-shaped or U-shaped outcome incidence at achieved BP values <120/70 mmHg or even <130/80 mmHg [538,1133,1137–1140]. It has also been reported that, within the above SBP values, a J-curve can be observed for myocardial infarction but not for cerebrovascular events [1139], suggesting an impaired ability to maintain myocardial perfusion at reduced BP values in coronary disease [1141,1142]. In a post hoc analysis of the INVEST trial, the J-curve for myocardial infarction was more evident in patients without than in those with coronary revascularization [1139]. Given the limitations of the post hoc and in general the observational approach to treatment effects, this should not discourage clinicians from pursuing the recommended BP target. However, attention to symptoms and signs of possible coronary underperfusion can be appropriate because at lower BP targets, there may be a small safety margin for adequate coronary perfusion, and more so if, as in CAD, coronary vessels are anatomically abnormal. Furthermore, LVH is common in CAD patients, and the increased cardiac oxygen needs and impaired microcirculation associated with this condition may favor ischemia at perfusion pressures that are too low [534,537] (see Section 10.5).

Prevention of coronary artery disease (CAD) in hypertension

figure45

17.1.2 Treatment of hypertensive patients with coronary artery disease

BBs, DHP-CCBs and nondihydropyridine CCBs are the preferred drugs for the treatment of hypertension in symptomatic CAD patients with angina, with a DHP-CCB and a BB in combination if needed. In patients with a recent myocardial infarction, BBs also improve prognosis [1143–1146] and should be prescribed unless contraindicated. The duration of the BB-related benefits is uncertain. However, in the absence of specific inconveniences, there is no reason to stop BB therapy [1145,1147,1148]. In this context, it is important to mention that increased heart rate correlates linearly with CV events, and the benefit of heart-rate reduction as a treatment goal in CAD patients has been demonstrated by several drugs including BBs [598,1149]. Thus, targeting a lower heart rate to a value below 80 bpm and close to 70 bpm, seems a wise additional treatment goal in hypertensive patients with CAD. BBs or non-DHP-CCBs can be used for this purpose. ACEis have been shown to reduce CV outcomes in high CV risk patients including patients with CAD in RCTs [1150,1151], which supports their use in CAD as part of the antihypertensive combination therapy, while ARBs can substitute ACEis in patients with hypertension and CAD who are intolerant to ACEis (Fig. 15).

F15
FIGURE 15:
BP-lowering therapy in patients with hypertension and coronary artery disease.aTarget heart rate below 80 beats per minute, if BBs are contraindicated or not tolerated consider use of non-DHP CCB at any step instead of DHP-CCB.

Treatment of hypertension in coronary artery disease (CAD)

figure47

17.2 Heart failure

17.2.1 Prevention of heart failure in hypertension

Prevention of HF received less attention than prevention of stroke and myocardial infarction until a review of the placebo-controlled RCTs in hypertension [1152] showed a HF reduction of 50% or more by antihypertensive treatment. Considering that RCTs were analyzed as ‘intention to treat’ and not ‘per treatment’ protocol, this indicated that the true prevention of HF was probably greater because large fractions of patients in the RCTs stopped taking the drugs they were randomized to, or changed from placebo to active treatment [1152]. The mechanisms for prevention of HF were explained by prevention of CAD (in particular, myocardial infarction), prevention of LVH, regression of LVH during antihypertensive treatment [228,229] or prevention of arrhythmias, mostly AF with rapid ventricular frequency on top of CAD or LVH. Prevention of LVH development and its regression are important goals for antihypertensive treatment and have a profound impact on prevention of HF [193]. Treatment with all major antihypertensive drugs reduces LVH, although BBs and diuretics may be relatively less effective [193]. Several analyses from the LIFE study have shown that in patients with hypertension, LVH regression induced by antihypertensive treatment was accompanied by improvements in indices of both diastolic function and systolic performance [231]. Additionally, reduction in LVH was associated with lower rates of clinical endpoints including new-onset HF independently of BP reduction [228]. Although similar evidence regarding LVH regression in patients with HFpEF has not yet been collected, available data in hypertensive patients with LVH support also pursuing LV mass reduction in patients with HFpEF, using strategies that lower cardiac afterload, such as reduction of peripheral vascular resistance and central BP [193]. However, it has also been reported that even after effective LVH regression, CV risk remains higher than for control individuals [1153], emphasizing the importance of relying on LVH prevention rather than treatment.

A few RCTs have received special attention because of the HF findings. Their results, as well as those of large meta-analyses of RCTs [80,583], can be summarized as follows. One, the protective effect of BP-lowering treatment on HF extends to the very old population in which the size of the protection appears to be particularly marked. HF is markedly reduced by lowering SBP to <140 mmHg, but single RCTs and large meta-analyses have shown that, compared with a target SBP within the 130– 139 mmHg range, an SBP reduction within the 129–120 mmHg range is accompanied by a significant and sizeable incremental benefit for HF. Two, unfortunately, this is accompanied by a noticeable increase of side effects (especially an increase of adverse kidney effects) leading to treatment discontinuation. Three, in hypertensive patients, diuretics, BBs, ACEis, ARBs and CCBs all have all been shown to reduce HF compared with a control group, with the protection exerted by diuretics extending to Thiazides, chlortalidone and indapamide. This strongly suggests the major protective factor to be BP-lowering ‘per se’. Four, meta-analyses of RCTs have reported that for a similar BP reduction, diuretics have a greater and CCBs a lower protective effect compared to the remaining drugs. However, these findings should be interpreted with caution, because the greater protective effect of diuretics might have derived from the previously discussed results of the large ALLHAT trial [566,1154,1155] as well as by a possible masking effect of diuretics on HF symptoms and signs, at least in the early and milder HF phase. Ankle edema (especially if associated with obesity or shortness of breath) may have also favored a spurious diagnosis of HF with CCBs. The overall conclusion is that all major antihypertensive drug classes importantly prevent incident HF in patients with hypertension and that thus all can be used for this purpose. The BP threshold and target for treatment do not differ from those recommended for general CV prevention by antihypertensive treatment. Alpha-1 blockers (doxazosin and others) can be added to the major drug classes, especially in combination with a Thiazide/Thiazide-like diuretic and a BB to minimize fluid retention and reflex tachycardia, i.e. symptoms resembling or explained by HF. The adverse effect of doxazosin on incident HF reported by ALLHAT [566,1154,1155] has not been confirmed by its use in patients with difficult BP control in the ASCOT trial [600].

Prevention of heart failure in hypertension

figure48

17.2.2 Heart failure with reduced ejection fraction (HFrEF)

Hypertension is a major risk factor for HFrEF, against which antihypertensive treatment has a major protective effect. However, in patients with clinically manifest HFrEF, an elevated BP is an uncommon problem because the reduced cardiac output counterbalances or overrides the increase in systemic vascular resistance typical of hypertension, usually leading to a normal or reduced BP. According to outcome-based RCTs, four drug classes in combination therapy are recommended for treatment of HFrEF in current HF guidelines [1156,1157]: ACEis or ARNI, BBs, steroidal MRAs and SGLT2is (Fig. 16). With the exception of SGLT2is, all these drugs are also recommended for the treatment of hypertension (Fig. 14). ACEis and BBs are part of the basic antihypertensive treatment strategy, while MRAs are recommended in patients with true resistant hypertension (see Section 12). Additionally, diuretics are recommended to manage fluid balance and reduce congestion, and are another antihypertensive treatment cornerstone. Thiazide/Thiazide-like diuretics are preferable if fluid retention is not a major problem or there is sufficient kidney function. Treatment with a loop-diuretic (furosemide, torsemide) is given to patients with severe fluid retention, in particular if they have suffered from pulmonary edema or have advanced CKD (eGFR <30 ml/min/1.73 m2). In the rare cases in which the above drugs in combination therapy at maximum tolerated doses are not sufficient to lower an elevated BP to control values (<130/80 mmHg), DHP-CCBs may offer an additional option because they have been shown to be safe when added to existing treatments in HFrEF and are indicated when HFrEF is associated with CAD and angina (see Section 17.1). However, in the last few years, use of ARNI (a compound that combines an ARB with the endopeptidase neprylisin inhibitor sacubitril) has made an uncontrolled BP in HFrEF even more unlikely, because this drug has a BP-lowering effect as shown in hypertensive patients [605]. Moreover, in HFrEF, a further BP-lowering effect can also be provided by use of SGLT2is (see Section 18). The multiple options that are available for the treatment of hypertension in HFrEF should be modulated according to their coexistence with CAD, myocardial infarction, LVH, AF, diabetes, COPD or other conditions.

F16
FIGURE 16:
BP-lowering drugs in hypertension and heart failure. (a) Non-DHP CCB are not recommended in HFrEF and should not be combined with BB. (b) Use of Diuretics: Use T/TLDiuretic if eGFR >45 ml/min/1.73 m2. Consider transition to Loop Diuretic if eGFR is between 30 to 45 ml/min/1.73 m2. Use loop Diuretic if eGFR <30 ml/min/1.73 m2 or in patients with fluid retention/edema.

Treatment of hypertension in heart failure with reduced ejection fraction (HFrEF)

figure50

17.2.3 Heart failure with preserved ejection fraction (HFpEF)

HFpEF accounts for about half of all HF patients, and hypertension is by far the most frequent precursor of and comorbidity for this condition [193,1158]. This accounts, at least in part, for the high prevalence of LVH in HFpEF as well as for the appearance and progression of both diastolic and systolic dysfunction [193,1158]. Outcome-based RCT evidence is limited for HFpEF, and no trial with major antihypertensive agents has clearly documented that any specific antihypertensive drug class is associated with a reduction in mortality and hospitalization [193,1158]. However, (i) the close pathophysiological association of HFpEF with hypertension and cardiac HMOD and (ii) the marked reduction in the risk of any HF type with drug-based control of BP elevations [466] have led to the agreement that reduction of an elevated BP “per se” can be an appropriate therapeutic intervention in HFpEF. However, the SGLT2i class of drugs has recently been shown to importantly improve the primary outcomes in dedicated RCTs on HFpEF [1159,1160] in both diabetic and nondiabetic patients, and can, therefore, be used to treat this condition [1156]. Treatment with ARNI [762] or MRA (spironolactone preferred) [1161] can be also considered [193,1156]. Based on available evidence, we recommend to lower SBP to 130 mmHg in HFpEF patients. Replacement of RAS blockers by ARNI or addition of SGLT2is can be considered, the latter independently of the presence of diabetes (Fig. 17).

F17
FIGURE 17:
BP-lowering therapy in hypertension and HFpEF. (a) Use of Diuretics: Use T/TLDiuretic if eGFR >45 ml/min/1.73 m2. Consider transition to Loop Diuretic if eGFR is between 30 to 45 ml/min/1.73 m2. Use loop Diuretic if eGFR <30 ml/min/1.73 m2 or in patients with fluid retention/edema. (b) BB should be used as guideline directed medical therapy in respective indications or considered in several other conditions (Table 16) (c) Use SGLT2i according to approval.

Treatment of hypertension in heart failure with preserved ejection fraction (HFpEF)

figure52

17.2.4 Overall management of heart failure and classification

Optimal management of HF patients includes special care programs for both hospitalized patients and outpatients who can be managed in HF clinics, in which the problem of long-term drug adherence is also addressed [1162]. Patients with very low EF (<20%) may need to be stabilized in hospital as they frequently present with tachycardia as a compensatory mechanism to maintain cardiac output. Treatment can make use of BBs, which should be titrated carefully from initial small doses. Hospitalization is also indicated in patients known to be hypertensive who show a low BP in a severe decompensated HF condition. Under these circumstances, the hemodynamic hallmark is a very high systemic vascular resistance, and it may take several days or even up to 2 weeks to re-establish acceptable BP values. Finally, it should be mentioned that after decades of an HF classification based on reduced, intermediate or preserved ejection fraction, this practice is gradually declining because of the evidence that HFrEF and HFpEF are not characterized by selective systolic and diastolic dysfunction, respectively, but by the coexistence of the two abnormalities in either condition. This is especially evident in HFpEF, in which patients also have extensive systolic dysfunction [229] seen as a lack of longitudinal axis shortening in systole, poor mid-wall contractility or low stroke volume at echocardiography. Furthermore, the effects of virtually all drugs in HF are independent on the ejection fraction value [193,1158]. In this regard, the field seems to be moving towards a unified pathophysiologically based classification of HF, which might lead to a therapeutic unification as well.

17.3 Hypertension and atrial fibrillation

AF is by far the most common arrhythmia in adults, and both its persistent and paroxysmal (more or less frequent AF episodes of variable duration) phenotypes are associated not only with increased risk of stroke but also with an increase of overall CV morbidity and mortality, including the risk of developing HF [1163,1164]. It is estimated that worldwide approximately 2– 4% of adults suffer from AF, with a progressive increase with age and a prevalence up to 10% in older people and 20% in octogenarians [1163–1165]. The prevalence of AF will probably increase over the next decades because of the aging population [1166] as well as the expected increase in the prevalence of the risk factors for this condition [1163–1165]. Hypertension is the most common risk factor for the onset of AF because of (i) its large prevalence in the population and (ii) its role in the determination of cardiac alterations that favor AF such as LVH, LA enlargement and structural LA wall changes [1167]. Even a high-normal BP in apparently healthy people predicts AF [1168,1169]. Hypertension and AF are the two most important risk factors for ischemic thromboembolic and hemorrhagic stroke [4,1163–1165]. An elevated BP increases the risk of ischemic stroke, intracranial hemorrhage, major CV events and all-cause death in both oral anticoagulant-naive [1170] and oral anticoagulant-treated patients with AF [1171]. In patients eligible for anticoagulation treatment (because of increased stroke risk) a meta-analysis of RCTs (71 683 patients) indicated that hypertension was clearly the most prevalent comorbidity (88% of the patients) in the older population (mean age 71.5 years) [1172].

17.3.1 Blood pressure measurement in atrial fibrillation

Accurate measurement of BP is important and challenging in patients with AF who are usually older and more frequently hypertensive than people without AF [1173]. However, BP measurement in the presence of AF (and to a lesser degree of other major arrhythmias) can be problematic because of marked variations in ventricular filling time, ventricular contractility and stroke volume, all of which contribute to a marked increase of beat-to-beat BP variability [1173]. The previous 2018 ESC/ ESH guidelines [4] recommended multiple BP measurements by auscultation in AF patients to account for the varying BP values. No less than three readings should be used to define the office BP representative of a given visit or circumstance. However, use of automated oscillatory methods should not be excluded because a recent meta-analysis of validation studies that have made use of different methodologies has shown that in AF patients, oscillatory methods satisfactorily measure SBP and only modestly overestimated (2.1 mmHg) DBP [1173]. This is clinically relevant because AF patients are usually older (i.e. when SBP is prognostically more important than DBP) and more frequently affected by ISH. In addition, despite the methodological limitations, both auscultation and automated oscillatory BP measurements have prognostic value in AF patients and can thus be used for office measurements, while only the automated oscillatory method can be used for home and ambulatory BP assessment [127,1174]. Recently, specific algorithms for accurately detecting AF during BP measurements [105,1174–1178] have been included in automated oscillatory BP monitors. This may increase the potential for an extended AF detection because a large fraction of AF episodes is asymptomatic [1163,1164].

17.3.2 Detection of atrial fibrillation

Because the lifetime risk for development of AF in adults is high (one of three Europeans >50 years of age) [1165], detection of AF is of fundamental importance. Detection of AF is based on pulse palpation, ECG and 24 h Holter monitoring. Holter monitoring can be prolonged to 48 h or longer to increase the chance of detecting asymptomatic or misinterpreted episodes. When assessing ambulatory BP, use of automated BP devices with an algorithm that detects arrhythmias may be considered [1176]. In a meta-analysis that included six clinical trials (2332 patients), the diagnostic accuracy of AF detection with automated BP measurement [1176] showed a sensitivity of 0.98 and a specificity of 0.92. Wearable and cuff-less devices allowing the detection of AF are evolving, and their use may allow more extensive and earlier detection of AF in the future [1179].

17.3.3 Prevention and treatment of atrial fibrillation in hypertension

Prevention and treatment strategies for AF do not substantially differ in patients with and without hypertension. BBs combine their BP-lowering effect with heart rate reduction in patients with hypertension and rapid AF, thereby helping to achieve heart rate control, which is central in AF management [1180]. Digoxin may be added to a BB but often has a limited effect. Heart rate control can also be achieved with a non-DHP-CCB (diltiazem or verapamil), which should generally not be used in combination with BBs [1180]. An exception are hypertensive patients with severe palpitations and rapid AF, who may be candidates for cautious use of a BB in combination with a non-DHP-CCB. With this approach, the addition of amiodarone with its potentially serious adverse effects during long-term treatment can be avoided. Control of heart rate below 110 bpm is advisable for all AF patients targeting a resting heart rate below 80 bpm based on ECG analysis, while the optimal heart rate target remains to be documented [1180].

Although several small studies suggest a moderate decrease of AF recurrence rate by BBs, the effect of these drugs on clinically relevant outcomes (stroke, systemic embolism or HF) remains to be established [1181]. Nevertheless, in the absence of contraindications or side effects, BBs may have been preferentially used in hypertensive patients with AF. All major antihypertensive drug classes favor LVH regression (via reduction of the afterload), which is therapeutically appropriate in AF, because LVH predisposes to development and relapse of AF. ACEis, ARBs and CCBs are more effective on LVH regression than BBs and diuretics, with encouraging results for ACEis and ARBs in preventing AF in patients with LV dysfunction, LVH or alterations of the anatomical structure of the LA [1182–1185]. ARBs are also more effective than CCBs in preventing AF in patients with high-risk hypertension [1186], but no reduction of the AF burden with the use of ARBs in patients without structural heart disease has been reported [1187]. Furthermore, ARBs did not avoid relapse of AF after electroconversion in patients mostly without hypertension [1188]. Use of MRAs may decrease new-onset AF in patients with HF and preserved [1186] or reduced ejection fraction [1189]. Recent data have shown that the use of SGLT2is is associated with a significant decrease in the risk of incident AF in patients with or without diabetes [1190,1191]. However, there is not yet evidence that SGLT2is prevent stroke in these patients. When compared with DPP-4 inhibitors or GLP-1 RA, the risk of incident AF was significantly lower with SGLT2is [1192].

Prevention of atrial fibrillation in hypertension

figure53

Treatment of hypertension in atrial fibrillation

figure54

17.3.4 Oral anticoagulation and BP control

In AF, the administration of anticoagulants has become mandatory if the risk of stroke is high and greater than the risk of major bleeding associated with these drugs [1163,1164]. This can be assessed by scores that weight and balance the respective favoring factors [1163,1164]. However, treatment of hypertension and achievement of BP control has not lost its importance [1193] for two main reasons. The first reason is that multiple studies and meta-analyses have shown that in patients with AF, an elevated office BP or a diagnosis of hypertension is accompanied by a marked increase in the risk of stroke, systemic embolism, intracranial hemorrhage and CV morbidity and mortality. The risk includes different ethnicities and extends to very old people [127,1194]. The outcome incidence increased significantly also in patients with a home SBP ≥ 145 mmHg compared with a home SBP <125 mmHg (approximately equivalent to <130 mmHg office SBP), incidentally supporting the usefulness of HBPM in this clinical circumstance [127]. The second reason is that the BP level is a most important factor favoring major bleeding, including intracranial hemorrhage [1195]. This was documented early by a post hoc analysis of the PROGRESS study on patients with a history of cerebrovascular events and warfarin treatment, who showed a progressively lower incidence of intracranial hemorrhage, as BP was reduced by antihypertensive treatment to an SBP value of <130 mmHg [1196]. In addition, data from RCTs comparing warfarin with direct oral anticoagulants in AF showed a significant increase in the risk of ischemic or hemorrhagic stroke if SBP values were >140 mmHg [1197,1198]. This evidence has confirmed the inclusion of hypertension, as defined by an SBP >160 mmHg, as an important modifiable risk factor for bleeding in the HAS-BLED score [1199], although based on the available evidence mentioned above a rapid reduction of SBP values to <140 mmHg before initiation of anticoagulant treatment appears to be a more adequate safety measure. According to the present guidelines, BP needs close monitoring in patients with AF who are under anticoagulant treatment, and uncontrolled hypertension should be avoided by appropriate BP-lowering therapy. The SBP target should be <140 mmHg but values <130 mmHg should be cautiously pursued, if treatment is well tolerated, because many patients with AF have LHA, which may require cardiac perfusion pressures higher than in patients without LVH (see Section 10.5). There should be no attempt to lower SBP below <120 mmHg because the cautionary reasons mentioned in Section 10 apply also to patients with AF, in whom there is some evidence that BP levels and outcomes may be linked by a ‘J-shaped’ curve [1170,1200,1201]. HBPM may help evaluation of antihypertensive treatment effectiveness over time. Oral anticoagulants in patients with an elevation of SBP >160 mmHg, may be halted until BP control is improved or achieved. Most first-line BP-lowering drugs are safe in patients undergoing oral anticoagulation without a significant risk of clinically relevant drug interactions. Non-DHP-CCBs (verapamil and diltiazem) are an exception [1202], because they are moderate inhibitors of the cytochrome P 450 isoenzyme 3A4 and P-glycoprotein and may, therefore, increase the plasma concentration of oral anticoagulants and thus the bleeding risk (Fig. 18) [1203,1204].

F18
FIGURE 18:
BP-lowering therapy in hypertension and atrial fibrillation. (a) Target heart rate below 80 beats per minute, if BBs are contraindicated or not tolerated consider use of non-DHP CCB at any step instead of DHP-CCB.

Management of patients with hypertension and AF during oral anticoagulation

figure56

17.4 Valvular heart disease

The most common valve disorders of the heart are aortic stenosis (AS), aortic regurgitation (aortic insufficiency) and mitral regurgitation (mitral insufficiency). Detection is usually obtained by auscultation of murmur and subsequent echocardiography. In patients with progressive AS severity, ultimate treatment is open valve surgery or catheter-based valve replacement or repair. At a later stage, cardiac valve disease may be a component of the HF syndrome, and patients should receive HF treatment. Hypertension is common in patients with valvular heart diseases and particularly in patients with AS.

17.4.1 Aortic stenosis

AS is a degeneration of the aortic valve with fibrotic tissue and calcification. A bicuspid valve predisposes to this development. The pathophysiological mechanism is different from atherosclerosis with plaque formation in the intima of large arteries because intensive lipid-lowering treatment does not prevent worsening of aortic stenosis [1205]. At least two-thirds of patients with AS have hypertension, which plays a pathophysiological role. There is no major RCT on the role of antihypertensive medication to prevent worsening from moderate-to-severe AS. Thus, treatment of hypertension in patients with AS depends on learning from small mechanistic studies, clinical experience and data from observational studies [1206–1208]. Severe AS is associated with syncope, arrhythmias, chest pain because of myocardial ischemia (related to LVH or CAD) or HF. Most clinicians are somewhat cautious with antihypertensive treatment in patients with severe AS, who have suffered a syncope in order to avoid a sudden BP fall and recurrent syncope. An interesting, albeit still unresolved, question is whether LVH develops because of longstanding hypertension or in response to the aortic valve disease itself. At any rate, before and after surgery or catheter-based intervention, antihypertensive treatment in patients with AS should follow the general treatment algorithm, including target BP values, the choice of drugs and the specific indications for CAD, HF or arrhythmias [1206–1208]. Postoperative echocardiographic FU should be implemented until normalization of the cardiac geometry.

17.4.2. Aortic regurgitation

Hypertension with aortic regurgitation (insufficiency) should be treated according to the general treatment guidelines until the valve disease needs surgery or catheter-based treatment, which usually takes place when a dilated left ventricle is detected by echocardiography. An RCT performed three decades ago showed that treatment with nifedipine could delay the time point for surgery [1209], and thus treatment with CCBs is common in aortic regurgitation with the purpose of reducing the afterload by systemic vasodilatation. However, other vasodilators (e.g. RAS blockers) can serve the same purpose. In people with aortic regurgitation, the heart itself generates a high SBP because of a higher stroke volume.

17.4.3 Mitral regurgitation

Numerous cardiac diseases may be associated with damage of the mitral valve and mitral regurgitation (insufficiency). Longstanding or severe hypertension may cause mild or even moderate-to-severe mitral regurgitation through dilatation of the left ventricle and thus stretching of the chorda tendineae, separation of the two mitral leaflets in systole and leakage of blood into the left atrium. This appears to be a sequence of events that is not rare, because in a British population study [1210], each 20 mmHg increment of SBP was associated with a 26% higher risk of mitral regurgitation. Antihypertensive treatment and BP control can be cardioprotective when mitral regurgitation is diagnosed according to the general recommendations for the treatment of hypertension. Patients should be followed by echocardiography for the regression of the LV dilatation and reversal of mitral regurgitation. The mitral valve is usually intact, and valve repair is not indicated if antihypertensive or HF medications reverses the regurgitation. Valve repair is indicated if rupture of the chordae contributes to a severe mitral valve leakage or the chordae are damaged and the leaflets do not clog up the leakage.

Hypertension and cardiac valve disorders

figure57

17.5 Cerebrovascular disease and cognition

17.5.1 Management of elevated BP in acute stroke

17.5.1.1 Acute hemorrhagic stroke

In acute intracerebral hemorrhage, an increase of BP is common and associated with a greater risk of hematoma expansion, death and reduced chance of neurological recovery [1211,1212]. Management of the BP increase is different according to whether the BP-lowering intervention starts <6 or >6 h after onset of symptoms. Two earlier RCTs suggested that in patients treated <6 h after onset of symptoms lowering SBP to >140 mmHg from much higher initial values reduced disability and death whereas SBP reductions to <140 mmHg led to no benefit, and in one of the two RCTs, increased adverse kidney events [1213,1214]. However, a continuous association between the degree of the first 24 h BP reduction and the improvement of functional neurological status has been unveiled by an analysis of 3809 patients in whom treatment started on average 3.6 h after the onset of symptoms [1215]. Furthermore, in a meta-analysis of five RCTs, patients with small-to-moderate hematoma volumes at admission showed an improvement of the functional neurological status with a BP reduction to <140/90 mmHg [1216]. Thus, more rigorous BP targets may now be recommended for early BP-lowering treatment in patients with acute intracranial hemorrhage. Nevertheless, a caveat against excessive BP-lowering effects remains because a pooled analysis of the patients enrolled in the INTERACT2 and the ATTACH-II trials showed that patients with an SBP decrease greater than 60 mmHg had a worse prognosis compared with those with lesser SBP reduction [1217].

In patients with acute intracranial hemorrhage and an SBP <220 mmHg, current evidence for initiation of BP-lowering treatment >6 h after onset of symptoms is not entirely univocal. Recently, an individual patient data meta-analysis of 16 studies (6221 participants) has shown that a moderate SBP reduction (12.1 mmHg achieved over several hours) reduced the hematoma expansion, although with no clear effect on the chance of clinical recovery. Therefore, slow and moderate BP reductions are preferable over intensive BP targets [1216,1218]. Fewer data are available in patients with acute intracerebral hemorrhage and an SBP ≥ 220 mmHg. A meta-analysis [1219] and secondary outcome data from one RCT [1213] suggest an improvement of functional recovery with an SBP reduction to <180 mmHg. However, in the ATTACH-II trial, patients with SBP ≥ 220 mmHg showed a higher rate of clinical deterioration by an SBP reduction to <140/90 mmHg [1220]. Thus, in patients with acute intracerebral hemorrhage and an SBP ≥ 220 mmHg, a cautious BP reduction to SBP <180 mmHg, possibly via i.v. drug therapy, seems the wisest therapeutic option (Fig. 19).

F19
FIGURE 19:
BP management in acute stroke. (a) Avoid absolute reductions of SBP >60 mmHg from initial SBP.
17.5.1.2 Acute ischemic stroke

The beneficial effects of BP reduction are even less clear in acute ischemic stroke. In most patients, initial BP values are high or very high and show a spontaneous progressive reduction during the first 48–72 h after stroke [1221]. Elevated BP levels during this time frame are associated with a worse clinical and/or neurological outcome but this association cannot translate into the decision to pharmacologically reduce BP because most RCTs and meta-analyses have not reported reduction of death or dependency by early BP reduction after acute ischemic stroke [1216,1222–1228]. A further difficulty is that RCT evidence has been obtained with different measures of the benefit in patients differing for the type of stroke (lacunar, large vessel occlusion, cardio-embolic), the age and clinical features (e.g. a history of previous hypertension or HF), the magnitude of spontaneous BP changes during the first few hours after stroke, the infarct size and intracranial pressure, which may have concealed benefit or harm of BP-lowering interventions in some patient categories [1229]. A pragmatic recommendation can be to consider a cautious and slow BP reduction (15% over the 24 h after stroke) in patients with markedly elevated SBP or DBP values (i.e. 220/120 mmHg) [1225,1227,1230], and to abstain from BP-lowering treatments when BP is <220/120 mmHg during the 72 h after stroke, because under these circumstances, no benefit of a BP reduction has been consistently reported [1227,1231].

Patients with an acute ischemic stroke who have or will receive reperfusion interventions such as intravenous thrombolysis or mechanical thrombectomy are a special category, because observational studies have reported that these interventions increase the risk of intracerebral hemorrhage if BP is markedly elevated [1232,1233]. In these patients, BP should be lowered to <180/105 mmHg for at least the first 24 h after thrombolysis or thrombectomy [1234]. BP should be maintained stable at the lower BP values because a meta-analysis of seven studies (5874 patients) has reported a 20% increase of intracranial hemorrhage (and 12% increase of worse neurological outcome) for any 10 mmHg BP increase [1235]. On the other hand, no benefit has been found in these patients with an SBP reduction to <130 mmHg [1236,1237] (Fig. 19). For stable patients, who remain hypertensive (≥ 140/90 mmHg) more than 3 days after an acute ischemic stroke, initiation or reintroduction of BP-lowering medications should be considered [1238].

Management of blood pressure in acute stroke

figure59

17.5.2 Management of elevated BP in patients with previous stroke or transient ischemic attack

Placebo-controlled RCTs of antihypertensive treatment in clinically stable hypertensive patients (BP ≥ 140/90 mmHg) with a previous stroke or TIA [1239,1240] have shown that BP-lowering reduces the risk of recurrent stroke as well as CV events. Thus, initiation or resumption of BP-lowering therapy several days after stroke, when the clinical conditions are stabilized, or immediately after TIA, is recommended for untreated or previously treated patients with hypertension [1241]. No placebo-controlled trial has explored whether in patients with a previous stroke, antihypertensive treatment reduces stroke recurrency and CV events also when BP is in the high-normal range [1242] or lower. The optimal BP targets to prevent recurrent stroke are also uncertain, but a consistent finding of several trials and meta-analyses has been that within the 120–140 mmHg SBP range, the lower the achieved SBP, the lesser the risk of stroke recurrence [466,1242–1249]. It must be emphasized that these results are mainly applicable to individuals with an average age below 70 years, and that, when dealing with the secondary prevention of stroke, the target BP to aim at should be decided based on the functional status, frailty, cognition and associated conditions of the patient. The first and main goal should be to reduce BP to <140/80 mmHg, and then, whenever possible and under clinical control, achieve BP below 130/80 mmHg, if tolerated [1250]. SBP values <120 mmHg should be avoided.

Prevention of stroke has been observed in large RCTs using different drug regimens. However, RCTs comparing different treatment regimens [625,1251] and meta-analyses [466,583,584,1252] suggest that BBs are less effective for stroke prevention than the other major classes of antihypertensive agents, although also showing a sizeable protection against stroke in BP-lowering placebo-controlled trials [584,625,1251]. The factors involved in the lower cerebrovascular protection of BBs are not clear because there is no evidence that BBs exert a damaging effect on the brain or impair cerebral blood flow autoregulation. In a large meta-analysis of RCTs, the risk of stroke did not differ significantly between BBs and RAS blockers or diuretics, but it was greater when compared with CCBs [625,1251], raising the possibility of an origin from a slightly greater BP-lowering effect of CCBs to which stroke incidence might be especially sensitive [625,1251,1253,1254]. At any rate, mindful of the fact that the most common recurrent event after stroke is a further stroke rather than myocardial infarction [1255–1257], antihypertensive treatment for secondary stroke prevention should not consider BBs as the preferred drugs. Under these circumstances, BBs can be used in combination treatment, considering their specific indications and comorbidities (Table 16).

17.5.3 Management of patients with cognitive dysfunction and dementia

In the last 25 years, the incidence of dementia has sizably increased, mainly because of the increase in population aging. Dementia is more frequent in women than in men and is the fifth most common cause of death in the world [1258]. Several epidemiological and clinical studies have shown that hypertension in midlife predicts cognitive decline and both Alzheimer's disease and vascular dementia in older patients [1258,1259]. Furthermore, long-term cumulative BP is independently associated with subsequent cognitive decline and incident dementia among cognitively healthy adults [312,1260]. The pathophysiology of cognitive function in hypertension is related to remodeling of cerebral small vessels, leading to subclinical cerebral white matter lesions, microbleeds and lacunar infarcts [1261]. Arterial stiffness of large arteries and flow pulsatility contribute to cerebral small vessel disease, which results into blood flow reduction in specific brain regions related to cognition such as the basal ganglia and hippocampus [308,1262–1264]. Therefore, in hypertensive patients, routine clinical assessment should include attention to possible cognitive impairment, at least in those aged 65 years and older [300]. Evidence on the beneficial effects of BP-lowering on cognitive decline has been conflicting for years. However, a recent meta-analysis [1265] of five RCTs (28 008 patients) used multilevel logistic regression of pooled individual participant data to evaluate the treatment effect on incident dementia. During a median FU of 4.3 years, antihypertensive treatment reduced the risk of incident dementia by 13% with a mean SBP/DBP lowering of 10/4 mmHg. In addition, several studies have shown that strict BP control, i.e. SBP <130 mmHg, reduces the progression of cerebral white matter lesions and the decrease in global cognitive performance [471,1266,1267].

The question if some antihypertensive drugs or strategies are better than others in preventing cognitive decline and dementia is still under debate. Several observational studies and data from international registries suggested that ARBs, DHP-CCBs and Thiazide/Thiazide-like diuretics may be better than ACEis, non-DHP-CCBs and BBs in reducing the progression of cognitive decline and the incidence of dementia [1268,1269]. This suggestion seems to be supported by a very recent post hoc analysis of two RCTs, the PreDIVA trial [1270] and the SPRINT-MIND trial [1271]. Both trials have shown that treatment with ARBs, DHP-CCBs and Thiazide/Thiazide-like diuretics had lower rates of incident cognitive impairment (-24%) compared to ACEis, non-DHP and BBs. Further prospective controlled trials designed to confirm these observations are warranted. Current evidence supports the recommendation to implement antihypertensive treatment and pursue strict BP control in late-mid and later life to lower the risk of cognitive decline and dementia.

17.6 Vascular disease

17.6.1 Lower extremity arterial disease

LEAD is often a manifestation of more widespread atherosclerosis and especially of atherosclerotic renal artery disease. Indeed, an association has been found between LEAD and an increased risk for multiple adverse outcomes including CAD, HF, aortic aneurysm and CKD [1272], which means that patients with LEAD are at high or very high CV risk [1273]. LEAD is associated with BP levels and, in a large primary care registry from the UK, patients with a 20 mmHg higher than usual SBP exhibited a 63% greater risk of LEAD [1272]. The strength of the association declines with increasing age [1274]. So far, no RCT of different BP targets has been designed to specifically examine the effect of BP-lowering treatment on LEAD-related events as well as the BP targets to aim at and the antihypertensive drugs to use. Nevertheless, BP control by antihypertensive treatment can be recommended as an important part of the CV risk reduction strategy in LEAD patients with hypertension [485] to increase the current low rate of BP control that characterizes LEAD in clinical routine [1275]. Available evidence and extrapolation from the ALLHAT study suggests that major BP-lowering drug classes including diuretics, CCBs and RAS blockers prevent LEAD events with equal efficacy [1276,1277]. Therapeutic use also includes BBs (not primarily tested in ALLHAT) because these drugs have not been shown to worsen the symptoms of intermittent claudication in two meta-analyses [1278,1279]. In an RCT with the vasodilating beta-1 selective BB nebivolol and the nonvasodilating beta-1 selective BB metoprolol in patients with stable claudication [1280], both treatments were well tolerated, and there were no differences in quality of life, ABI and claudication distance. Thus, BBs remain one of the treatment options in hypertensive patients with LEAD, also considering the frequent association of this disease with CAD. When critical limb ischemia is present, BP reduction should be instituted slowly to reduce the risk of worsening ischemia. In the ALLHAT study, an on-treatment SBP <120 mmHg (and >160 mmHg) was associated with a higher hazard of a LEAD events (LEAD-related hospitalization, procedures, medical treatment or death), in comparison with an SBP between 120 and 129 mmHg. This was the case also for a low DBP (<60 mmHg) [1276]. Furthermore, in a large trial (EUCLID) [1281], an SBP >125 mmHg was associated with an increase in major CV and LEAD events, whereas SBP to ≤ 125 mmHg was associated with an increase of major CV, but not of LEAD events, and DBP was not associated with LEAD events at all. In patients with LEAD, antihypertensive treatment should include lifestyle changes, and especially smoking cessation. Lipid-lowering drugs and antithrombotic therapy [1282], are also frequently needed [1273].

17.6.2 Aortic dilatation, aneurysm and dissection

Hypertension can be associated with modest aortic root dilatation [1283]. When more extensive aortic root dilatation is present or the dilatation extends beyond the aortic root, an additional cause for aortopathy should be sought. All hypertensive patients with aortic dilatation, whether associated with Marfan syndrome, bicuspid aortic valve disease or not, should have their BP controlled [1283]. In patients with Marfan syndrome, prophylactic use of ARBs [1284] or BBs [1285] was associated with a reduction of either the progression of the aortic dilatation or the occurrence of complications as compared with no treatment. Both drugs were similarly effective [1286], at variance from a neutral effect of a losartan-based treatment in another study [1287]. In a recent individual patient data meta-analysis [1288], ARBs reduced the rate of increase of the aortic root diameter by about one-half, and the effect was similar with BBs. The authors suggested that combination therapy with both ARBs and BBs from the time of diagnosis would provide even greater reductions in the rate of aortic enlargement than either treatment alone, which, if maintained over a number of years, would be expected to lead to a delay in the need for aortic surgery. Antihypertensive drug treatment aimed at a 24 h SBP <130 mmHg has been recommended in the past and still seems a reasonable treatment goal [1289], although not firmly established by RCTs. There is no evidence on the efficacy of antihypertensive treatment in aortic disease of other etiologies.

Hypertension is an important risk factor for aortic dissection [1290], and under these circumstances, BP-lowering drug treatment must be implemented immediately. BP should be reduced at least to <130/80 mmHg [1283], but lower values, e.g. around 110 mmHg SBP [1289], should be pursued in the acute setting and possibly also, if tolerated, on a chronic basis. In a large retrospective cohort study from Taiwan [1291], including almost 7000 patients with aortic dissection, use of BBs, ACEis or ARBs after hospital discharge was associated with long-term mortality reduction.

Hypertension is also a risk factor for abdominal aortic aneurysm, and some studies suggest that special attention should be attributed to the DBP elevation [1274]. A meta-analysis including more than 15 000 patients under surveillance for abdominal aortic aneurysm [1292] has reported that BP values had no effect on aneurysm growth rate, but higher mean BP or pulse pressure were associated with greater aneurysm rupture rate. Data on the effect of antihypertensive treatment are not univocal. Antihypertensive drugs including ACEis, ARBs, CCBs and BBs in addition to statins and antiplatelet drugs have been associated with a lower 5-year all-cause mortality in patients with abdominal aortic aneurysm [1293]. A meta-analysis that included 10 RCTs (2045 patients) with asymptomatic abdominal aortic aneurysm has recently shown that BP-lowering medications did not reduce the aneurysm growth rate or related clinical events [1294]. Due to limitations in sample size and event rates, a small protective effect could not be ruled out. Although five cohort studies have raised the possibility of a benefit of BBs on abdominal aortic aneurysm growth rate, this has not been confirmed by three RCTs using BBs [1295]. Similarly, although a case– control study suggested a beneficial effect of ACEis in reducing the risk of abdominal aortic rupture [1296], these drugs were found to have only a borderline effect in another observational study [1297]. Nevertheless, it seems wise to reduce a chronically elevated BP in people with abdominal aortic aneurysm, with no drug preference, and complying with the general recommended target BP values for antihypertensive treatment.

18. HYPERTENSION AND DIABETES MELLITUS

18.1. Epidemiology and risk classification

Hypertension is common in type 1 and much more in type 2 diabetes. Type 1 diabetes develops mostly in children, adolescents and individuals usually below 30 years and shows a prevalence of hypertension greater than that of nondiabetic individuals or the age-matched general population. Prevalence of hypertension varies in different studies, and in an old large Danish population study has been reported to be about 15% of type 1 diabetic patients [1298,1299]. Similar rates have been found in a study on more than 3000 patients from 16 European countries [1300]. The cause of hypertension is regarded as being diabetic nephropathy (usually documentable by increased urinary protein excretion), although some patients (frequently with a strong family history of hypertension) may develop a BP elevation in absence of manifest kidney disease [1299]. Hypertension is associated with its typical HMOD and outcome complications and antihypertensive treatment is needed. Treatment should be implemented according to the large body of evidence mainly collected for type 2 diabetic patients, on BP threshold for treatment (≥ 140/90 mmHg), BP target (<130/80 mmHg) and drug treatment strategies, i.e. use of drug combinations of the five major drug classes, including a RAS blocker [1301]. Hypertension is common in type 2 diabetes, with a prevalence that, after a few years’ duration of the disease, involves the majority of the diabetic population. It is estimated that this association will grow in the future due to the progressive aging of the world population and the unfavorable influence of modern lifestyle. The prevalence of type 2 diabetes is steeply increasing in high-income countries, and the increase is even more pronounced in low-income countries [37]. However, the association between type 2 diabetes and hypertension is also due to common causal links and bidirectional interactive influences, such as the stimulating effect of insulin on the SNS and the increase of insulin resistance and serum insulin levels by sympathetic activation [1302,1303]. This can explain why a quantitative relationship has been reported between the prevalence of type 2 diabetes and BP values in the population [1304]. It has also been recognized that, in diabetes, hypertension usually exhibits characteristics that differ from those of nondiabetic hypertensive patients. The most frequent characteristics are a greater SBP elevation, a wider pulse pressure, a higher BP variability, a nondipping pattern, salt-sensitivity, a trend to hyperkalemia and orthostatic hypotension, particularly as the duration of the diabetes increases [1305]. Furthermore, in patients with type 2 diabetes, the prevalence of MH is much higher than in the general population [1306]. Finally, type 2 diabetes is associated with a higher rate of resistant hypertension and is recognized to be one of the most important factors that can make achievement of BP control difficult [538]. The presence of diabetes mellitus in patients with hypertension has an important influence on CV risk “per se”, regardless of the concomitance of HMOD, CVD or CKD. Only diabetic patients with well controlled, short-standing duration of the disease (less than 10 years) with no evidence of HMOD and no additional CV risk factors are categorized as being at moderate risk [33]. Otherwise patients with diabetes are considered to be at high CV risk or even at very-high risk in the presence of established CVD or advanced CKD. Consequently, hypertensive patients with diabetes are candidates for immediate initiation of antihypertensive drug treatment together with lifestyle interventions.

18.2 Benefits of BP-lowering

Overwhelming evidence supports the benefits of BP reduction in people with hypertension and type 2 diabetes to reduce macrovascular events and mortality, as well as to prevent microvascular complications, such as nephropathy and retinopathy [1307–1309]. However, the protective effect of BP-lowering treatment is not clear for all diabetes-related microvascular complications, such as the potential protective effects of BP reduction on diabetic-related dysautonomia. The recommended lifestyle interventions (see Section 7) that lower BP are very important in type 2 diabetic patients, with particular emphasis on interventions targeting overweight and obesity that improve the blood glucose and the dysmetabolic profile. Pharmacological treatment should be started when SBP is ≥ 140 mmHg or DBP is ≥ 90 mmHg, to achieve, if well tolerated, a goal of <130/80 mmHg, which has been found to offer incremental protection compared to higher BP values, particularly against stroke in meta-analyses of RCTs [1308,1309]. Support for an SBP target <130 mmHg is provided also by the ACCORD trial, which found that in type 2 diabetic patients, on-treatment SBP values of about 122 mmHg were associated with a clear reduction in the risk of stroke compared to on-treatment SBP values between 130 and 139 mmHg [1310]. In ACCORD, the intense SBP reduction was not accompanied by a reduction of combined CV events and all-cause mortality, but it was later recognized that this finding was probably due to the confounding effect of the factorial design of the trial [1310]. When the detrimental effect of the concomitant intense blood glucose reduction was accounted for, the benefit of intense BP reduction extended to the composite CV outcomes [1142]. Although recent studies advocate a SBP goal <120 mmHg [1310,1311], other observations do not support these lowest BP targets [466,559,1308,1312]. The present guidelines recommend antihypertensive treatment of type 2 diabetic patients to reach an SBP <130 mmHg but not <120 mmHg. This recommendation extends to a DBP <80 mmHg but not below 70 mmHg [1311]. Interestingly, in the ONTARGET trial, an increased risk of events was observed at DBP values slightly <70 mmHg [538], but in another study, the impact of intensive BP-lowering was found to be independent of the baseline DBP value, suggesting that low baseline DBP should not be an impediment to pursuing optimal SBP targets[1313]. If a target <130/80 mmHg cannot be obtained or is not tolerated, maintaining BP within the 130– 139/80– 89 mmHg range guarantees a sizeable degree of protection compared to BP values 140/90 mmHg [183,1314].

18.3 Antihypertensive drug treatment

The generally recommended strategy for antihypertensive drug treatment (see Section 11), i.e. starting with dual combination therapy and using drug combinations in the majority of the patients, is even more necessary for diabetic patients, considering the difficulties of BP control in diabetes and the importance to achieve BP targets in patients with high CV risk [184]. However, treatment intensity should be implemented considering the importance to minimize inconveniences such as orthostatic hypotension, particularly in old patients and long-lasting diabetes, and possibly dysautonomia. All major antihypertensive drug classes have been shown to reduce CV outcomes in type 2 diabetes [1315]. Treatment should include a RAS blocker, because outcome-based RCTs indicate that RAS blockers prevent appearance and progression of kidney complications of diabetes more effectively than other major antihypertensive drugs, as measured by the reduced incidence of new-onset microalbuminuria, the reduction of protein excretion in proteinuric patients, the attenuated decline of GFR in diabetic and nondiabetic nephropathy and the prevention of ESKD [1316].

Newer antidiabetic agents, i.e. SGLT2is and GLP-1 RA have been shown to reduce macrovascular and microvascular complications in type 2 diabetes [1317]. In addition to glucose control, SGLT2is have demonstrated a marked protective effect against HF (see Section 17.2) and kidney outcomes (see Section 19) [1318], while GLP-1 RA have demonstrated a reduction of CV events and a sizeable weight loss [1319–1321]. Use of these drugs has been recommended by the ESC/EASD guidelines as first step treatment in diabetic patients with a previous CV event, HMOD or multiple risk factors [1322]. An important additional aspect of the action of these drugs is that they can reduce office BP and ABPM by several mmHg [1319,1323,1324] even when diabetic patients are under multiple treatment with antihypertensive drugs [183]. Both classes of drugs may thus help to improve BP control, which is especially difficult in diabetes [184]. Although protective effects of SGLT2is against HF and kidney outcomes have been clearly documented, a less consolidated degree of cardiorenal protection has been reported for GLP-1 RA in diabetes [758]. The new non-steroidal MRA finerenone has also been shown to provide cardiac and renal protection in patients with diabetic nephropathy (see section 19.1.4). In this context, finerenone induced small but significant reductions in BP [582].

Treatment strategies in diabetes

figure60

19. HYPERTENSION AND THE KIDNEY

Hypertension is a strong independent risk factor for development of CKD and progression of CKD to ESKD [1325,1326]. Hypertensive kidney disease “per se” is the most common known cause of ESKD, after diabetic kidney disease [1327,1328]. The diagnosis of hypertension-induced kidney dysfunction is based on two pillars: (i) evaluation of the level of kidney function through estimation of glomerular filtration rate (eGFR), calculated by the 2009 Chronic Kidney Disease-Epidemiology Collaboration (CKD-EPI) formula [290] and (ii) detection of kidney damage by use of the urinary albumin: creatinine ratio (UACR), measured from a spot urine sample (preferably morning urine). eGFR and UACR are both independent and additive predictors of increased risk of CKD progression and increased CV risk [1329–1331]. To highlight its associated risk, albuminuria is currently divided into: (a) normal/mildly increased, UACR <30 mg/g (formerly termed normoalbuminuria); (b) moderately increased, UACR 30– 300 mg/g (formerly termed microalbuminuria); and (c) severely increased, UACR >300 mg/g, (formerly termed macroalbuminuria). CKD is diagnosed in any individual with an eGFR <60 ml/min/1.73 m2 at any level of albuminuria or a UACR >30 mg/g at any level of eGFR persisting for more than 3 months. Serum creatinine, eGFR and UACR should be documented in all patients at the initial evaluation for hypertension if CKD is diagnosed, and repeated at least annually thereafter. A negative urinary dipstick test does not rule out albuminuria, because it cannot detect UACR levels in the lower range [296]. However, urinary dipstick can offer information on other signs of kidney injury (i.e. microscopic hematuria, active urine sediment) and should be performed at least at the initial evaluation. The value of kidney imaging, including kidney ultrasound with Doppler measurements, is discussed in Section 5.5.3.

19.1 Treatment of hypertension in CKD

High BP is by far the most common modifiable factor for CKD progression [1332]. True resistant hypertension, elevated night-time BP and MH are common in patients with CKD, and associate with a lower eGFR, higher levels of albuminuria, and HMOD [1325,1333–1336]. In patients with CKD, several of the mechanisms that originate from the kidney and promote hypertension are exaggerated. These include increased sodium sensitivity, sodium/water retention, activation of the RAS and SNS, impaired endothelium-mediated vasodilatation and others [1326,1337]. Complications unique to CKD, such as secondary hyperparathyroidism and increased calcium-phosphate products leading to arterial stiffness, increased prevalence of OSA, and use of specific drugs such erythropoietin, glucocorticoids or calcineurin inhibitors may also be involved in the BP elevation in later CKD stages [1338,1339].

19.1.1 Treatment BP targets

Available evidence suggests that BP reduction with any type of the major antihypertensive drug classes can offer similar protection against major CV events (stroke, myocardial infarction or CV death) and all-cause death in individuals with CKD [1340]. However, for more than a decade, there has been considerable debate on the best (most protective) BP targets in patients with CKD, including CKD with comorbid diabetes [1341,1342]. Old observational data suggested an association between BP and the risk for ESKD, starting from an SBP level of >120 mmHg [1325]. More recent data from China obtained in CKD patients without antihypertensive therapy followed prospectively for 5 years indicated that a BP >130/90 mmHg was associated with a significantly increased risk of CV and kidney outcomes [1343]. However, trial evidence on the most protective target BP in response to treatment of CKD, was missing. Current evidence originates mainly from two trials in nondiabetic CKD that randomized patients to different levels or ranges of mean BP and examined kidney outcomes. In the MDRD study, the projected eGFR decline within 3 years, and the risk of ESKD and death were not significantly different between groups of low and usual BP target [1344]. However, analyses by baseline proteinuria showed that, among patients with proteinuria >1 g/day, those in the low BP target group had a decrease in protein excretion and a slower GFR decline over time compared to patients in the usual BP target group [1345]. Similarly, in the AASK study, no difference in outcomes between different BP target groups was observed in the overall population [1346]; however, lower BP was associated with better kidney outcomes in the small subset of patients with proteinuria >1 g/day [1347]. Further analyses of MDRD and AASK combined the randomized trial periods with subsequent observational FU phases. In the MDRD long-term analysis, the lower BP target group was associated with reduced risk for ESKD or the composite of ESKD and death, again mainly driven by a beneficial effect in patients with baseline proteinuria >1 g/day. In the AASK long-term FU analysis, no difference in the risk of doubling of serum creatinine, ESKD or death was noted between the treatment groups. However, for patients with a UPCR >0.22 g/g (which roughly equals to a proteinuria of 0.25–0.3 g/day), the lower BP target carried a beneficial effect [1348]. A recent analysis combining these trials (1907 patients and a median FU of 14.9 years), showed that the lower target BP was associated with significant reductions in the risks of ESKD and mortality in the total population, and this effect was mainly driven by patients with UPCR >0.44 g/g [1347]. Thus, sustainability of BP reduction and extent of proteinuria are major determinants of nephroprotection in patients with nondiabetic CKD.

The results of the SPRINT study have little relevance to the question discussed above. SPRINT randomized 9361 hypertensive patients at increased CV risk to intensive (SBP <120 mmHg) or standard treatment (SBP <140 mmHg) [97]. About 28% had CKD with eGFR 20–60 ml/min/1.73 m2, but very few patients had proteinuria because individuals with proteinuria >1 g/day or >1 g/g were excluded. Diabetes mellitus, i.e. the most common cause of ESKD, or prior stroke were also exclusion criteria. In the overall trial, the primary composite outcome (CV events as well as CV and total mortality) was significantly lower in the intensive-treatment group, but kidney outcomes did not differ between the two target groups. A sub-analysis of the SPRINT CKD subpopulation [1349] showed no difference between groups for the primary outcome or the prespecified kidney outcome but a lower total mortality rate in the intensive BP arm patients. All the above results must be interpreted with caution, as the SPRINT trial was not designed or powered to study kidney outcomes, which were very few (15 versus 16 in the two groups) in the trial. As a result, a recommendation to target SBP to <120 mmHg in patients with CKD cannot be made.

No direct evidence is available to answer the question of the optimal target BP in patients with CKD and type 2 diabetes mellitus. Older studies, including UKPDS38 [1350] and the sub-analysis of participants with diabetes in the HOT trial [1351], offered insight on the DBP target, as they randomized patients to different on-treatment DBP levels. The ACCORD-BP trial randomized high-risk patients with type 2 diabetes to a target SBP <120 or <140 mmHg [1352]. Apart from showing no difference in the primary outcome (possibly because of interactions with other arms of the factorial design) (see Section 18.2) [1341], ACCORD-BP excluded individuals with serum creatinine >1.5 mg/dl, thus offering very little insight on the optimal BP target in patients with CKD and diabetes. A post hoc analysis of the RENAAL study showed that a baseline SBP of 140–159 mmHg increased the risk of ESKD or death by 38% compared to SBP <130 mmHg [1353]. A post hoc analysis from the IDNT showed that SBP >149 mmHg was associated with a 2.2-fold increase in the risk of doubling SCr or ESKD compared with SBP <134 mmHg. Moreover, progressive lowering of SBP down to 120 mmHg improved kidney and patient survival, while below 120 mmHg, all-cause mortality increased [1354]. Finally, although limited by the heterogeneity of the included studies [1341], a recent meta-analysis of studies in patients with CKD stage 3–5 has reported a mortality benefit by a SBP reduction of 16 mmHg and an absolute SBP of 132 mmHg with a nonsignificant benefit at achieved SBP values of <125 mmHg [1355]. In a more recent pooled analysis of four RCTs (AASK, ACCORD, MDRD and SPRINT), all-cause mortality showed a tendency to a reduction with intensive treatments (BP <130 mmHg), but this finding was not statistically significant (hazard ratio 0.87, 95% CI 0.69–1.08, P= 0.21). However, after excluding patients with higher GFR and those undergoing intensive glycemic control, lowering BP to <130 mmHg decreased all-cause mortality (hazard ratio 0.79, 95% CI 0.63–1.00, P = 0.048) when compared with a standard target of <140 mm Hg [1356].

Taking these largely indirect findings together and considering that, at least after development of proteinuria, progression of kidney injury tends to follow the same course in different situations, it may be suggested that (i) the BP target for proteinuric nondiabetic CKD applies to patients with proteinuric diabetic kidney disease as well and (ii) for both patient categories, a target SBP of <130 mmHg and DBP <80 mmHg, if well tolerated, can be associated with protection against CKD progression in individuals with an albuminuria >30 mg/g. A similar target may be associated with a reduction in mortality in most patients with CKD. Particularly in patients with advanced CKD in stage 4 and 5, careful monitoring of eGFR is recommended as a further functional, but reversible, decline of GFR may occur on a lower BP.

An office SBP of <120 mmHg and DBP <70 mmHg cannot be recommended because of the absence of relevant evidence. However, these guidelines acknowledge that these recommendations have a number of limitations: (i) none of the trials comparing different BP targets included patients with diabetes and CKD, thus current evidence cannot be readily extrapolated to this subpopulation; (ii) MDRD and AASK trials randomized participants to different mean BP levels, which cannot be readily extrapolated to SBP and DBP values; (iii) MDRD and AASK trials recruited patient populations of a relatively young age (mean age 51.7 and 54.6 years, respectively), and thus, their findings cannot be readily extrapolated to older patients with CKD and (iv) even for the long-term observational analyses, the benefits associated with lower BP targets were mainly apparent in individuals with proteinuria.

19.1.2 Antihypertensive drug treatment

People with CKD should receive lifestyle interventions, as reported in Section 7, with special attention to sodium restriction, as low-sodium diet reduces protein excretion in proteinuric CKD [1357]. Achieving the recommended BP targets in CKD usually requires combination therapy, which should consist of a RAS blocker with a CCB or a Thiazide/Thiazide-like diuretic, if eGFR levels are ≥ 45 ml/min/1.73 m2 (CKD stages 3a). While in patients with an eGFR below 30 ml/min/1.73m2, Thiazide/Thiazide-like diuretics should be generally replaced by loop diuretics, the transition from treatment with a Thiazide/Thiazide-like to a loop diuretic should be individualized in patients with eGFR values between 30 and 45 ml/min/ 1.73m2. Clinical trials in people with diabetic [1358–1361] and nondiabetic CKD [1346,1362,1363] have established an ACEi or an ARB as the first treatment choice in hypertensive CKD patients, especially in those with moderate or severe albuminuria, where these agents were found to reduce proteinuria, the rate of GFR decline, and the risk of doubling of serum creatinine or progressing to ESKD. ACEis or ARBs should be given at the maximum tolerated doses to achieve optimal nephroprotection while dual combination of RAS blockers should be avoided because two outcome trials were prematurely terminated as a consequence of the increased risk of adverse events with the ACEi/ARB combination therapy [561,1364]. In normoalbuminuric individuals with hypertension, ACEis or ARBs are able to delay the progression to severely increased albuminuria compared with placebo [1365], but no evidence exists on whether these drugs lead to better preservation of kidney function compared with other major antihypertensive drug classes in the normoalbuminuric population [1366] (Fig. 20). Indeed, in a recent open-label trial in which patients with an eGFR were randomly assigned to either discontinuation or continuation of therapy with RAS inhibitors, discontinuation was not associated with a significant between-group difference in the long-term eGFR decline [1367].

F20
FIGURE 20:
BP-lowering therapy in patients with hypertension and chronic kidney disease. (a) Transition from T/TLDiuretic to Loop Diuretic should be individualized in patients with eGFR < 45 ml/min/1.73 m2. (b) Cautious start with low-dose. (c) Check for dose adjustment according to renal impairment for drugs with relevant renal excretion rate. (d) When SBP is ≥140mmHg or DBP is ≥90 mmHg provided that: maximum recommended and tolerated doses of a three-drug combination comprising a RAS blocker (either an ACEi or an ARB), a CCB and a T/TLDiuretic were used, adequate BP control has been confirmed by ABPM or by HBPM if ABPM is not feasible, various causes of pseudo-resistant hypertension (especially poor medication adherence) and secondary hypertension have been excluded (see Section 12). (e) Caution if eGFR <45 ml/min/1.73 m2 or serum potassium >4.5 mmol/l. (f) Should be used at any step as guideline directed medical therapy in respective indications or considered in several other conditions (Table 16). (g) SGLT2is and Finerenone should be used according to their approval for CKD treatment.

19.1.3 Special therapeutic challenges

CKD poses several additional difficult treatment problems. Firstly, the vasodilating effect of ACEis or ARBs on the efferent arteriole reduces intraglomerular pressure. This is frequently followed by eGFR reductions, on average 10–15% in the first weeks of treatment. A similar effect can be seen with large BP reductions by any antihypertensive agent. Thus, repeated monitoring of eGFR (as well as serum electrolytes, see the following) within the first 4– 8 weeks of treatment (depending on baseline kidney function) is important when treatment is initiated. Clinicians should not be alarmed by this early GFR reduction, but if the decline in GFR continues or becomes more severe (>30%), the RAS blocker should be stopped, and the patient should be investigated for the presence of renovascular disease. Secondly, use of RAS blockers in CKD patients further increases the risk of hyperkalemia [1368]. Incident hyperkalemia is associated with increased mortality [1369] and is the most frequent reason for dose reduction or discontinuation of ACEi or ARB administration in CKD patients [1370,1371]. However, reducing the dose or discontinuing RAS blockers has been associated with increased risk of CV events in large surveys [1372–1374] and should be avoided.

Novel potassium binders (patiromer and sodium zirconium cyclosilicate) were shown to normalize elevated serum potassium and chronically maintain normal serum potassium levels in CKD patients treated with ACEis, ARBs or spironolactone, with a good tolerability [1375,1376]. These agents can be used to maintain serum potassium <5.5 mmol/l in patients with CKD [1377,1378]. Most patients with CKD will not achieve target BP control with ACEi or ARB monotherapy, and a DHP-CCB or a diuretic should almost always be included in the treatment regimen, most often both drugs [1379,1380]. DHP-CCBs may increase proteinuria when used in the absence of a RAS blocker in patients with proteinuric CKD [1361,1381]. Nevertheless, in the general hypertensive population, where the majority of patients has no or a mild albuminuria, DHB-CCBs have similar effects on kidney outcomes as RAS blockers or diuretics [1366]. In addition, in a study of hypertensive patients in whom 19% had moderate and only 5% severe albuminuria at baseline, a combination of a RAS blocker with a DHP-CCB was superior in reducing kidney outcomes compared with an RAS blocker in combination with a Thiazide [326]. Diuretics are particularly useful in CKD patients, because these patients are most often sodium-sensitive (especially if older, diabetic or obese) and have a high prevalence of true resistant hypertension. Furthermore, diuretics can effectively reduce proteinuria when added to RAS blockers in proteinuric CKD [1357]. When GFR falls below 45 ml/min/1.73 m2, Thiazide diuretics become less effective, because they cannot reach their tubular site of action because of competition for tubular secretion with other substances that accumulate in CKD [1382]. This is the case also for Thiazide-like diuretics, although a recent RCT that included 160 patients with CKD stage 4 reported a 10.5 mmHg 24 h SBP reduction in patients randomized to chlortalidone (mean dose 23 mg daily) [573]. In general, in patients with CKD stage 3b, eGFR 30–44 ml/min/1.73 m2, diuretic therapy should be modified and the dosing individualized, while in patients with CKD stage 4, eGFR <30 ml/min/1.73 m2, Thiazide should be substituted with a loop diuretic. Within this class, torasemide might be preferred to furosemide because of its longer half-life, which allows a less frequent dosing scheme and a better adherence to treatment [1326]. Finally, triple antihypertensive drug therapy may not control BP in a number of CKD patients. In these cases, BBs and alpha-1 blockers can offer important help, as sympathetic activity is increased in CKD [1383,1384]. However, their effects in CKD have not been tested in outcome kidney trials. Non-DHP-CCBs (if used with RAS blockers) were associated with reductions in proteinuria and decline of kidney function in proteinuric CKD [1385,1386], but when added to an RAS blocker in normoalbuminuric hypertensive patients, they did not seem to offer additional nephroprotection [1387]. In the PATHWAY-2, spironolactone was shown to be particularly effective for BP-lowering when added to the standard triple-drug therapy, but patients with eGFR <45 ml/min/1.73 m2 or potassium above 4.5 ≥ mmol/l were excluded in this trial [580] (see Section 12.4). In the AMBER trial that used spironolactone with addition of placebo or patiromer in patients with treatment-resistant hypertension and eGFR values between 25 and 45 ml/min/1.73 m2, BP was effectively reduced in both groups, but the rates of hyperkalemia (potassium 5.5 mmol/l) were at about 60 and 35%, respectively, at 12 weeks [740]. Thus, use of spironolactone as a fourth antihypertensive agent in patients with CKD Stage 3b or higher is generally not advisable, unless in special circumstances, such as at low potassium levels, or when BP control is not achieved with addition of other agents. Use of novel potassium binders is advisable to maintain serum potassium below 5.5 mmol/l (Fig. 20).

19.1.4 Use of additional drugs for cardiovascular- and nephroprotection in CKD

In addition to achieving BP control with the agents described above, progression of CKD and risk of CV events and mortality can be reduced in CKD patients by two novel drug classes that also have some BP-lowering effects, although they are not approved as antihypertensive agents. SGLT2is were first introduced as oral hypoglycemic agents that reduce plasma glucose levels by inhibiting renal glucose reabsorption through the SGLT-2 transporters, located in the proximal convoluted tubule. Early clinical studies in patients with type 2 diabetes suggested that these agents can reduce office SBP/DBP by 3– 5/1– 2 mmHg in hypertensive patients [1388]. This has been observed also for ABPM values [1389], and larger reductions have been described in patients with CKD Stage 4 (about 7 mmHg SBP) [1390]. The main antihypertensive mechanism is likely to be a mild natriuretic/diuretic effect (from inhibition of proximal sodium reabsorption) and osmotic diuresis [1388], although BP reductions have also been reported in patients under diuretic treatment [1391]. These agents were also shown to reduce urine albumin excretion by 25– 40%, depending on the baseline albuminuric levels [1392] and can reduce plasma uric acid, which is also important in CKD patients [1318].

CV outcome trials with SGLT2i in patients with type 2 diabetes have included large proportions of patients with CKD, showing impressive and homogeneous reductions (around 40%) of composite kidney endpoints [544,1319,1393]. The protective renal effects have been shown in three large trials on diabetic and nondiabetic CKD, in which SGLT2i were used on top of standard therapy, including an ACEi or ARB at maximum tolerated doses [1394]. The CREDENCE trial (4401 patients with type 2 diabetes and CKD with severely increased albuminuria [1320]); the DAPA-CKD study (4304 patients with either diabetic or nondiabetic CKD and an UACR ≥ 200 mg/g) [1395]); and the EMPA-KIDNEY trial (6609 patients with either diabetic or nondiabetic CKD, an eGFR between 20 and <45 ml/min/1.73 m2 and any level of albuminuria or 45 to <90 ml/ min/1.73 m2 with a UACR of ≥ 200 mg/g)[1395]. All three trials were prematurely terminated due to significant reductions of the composite and individual kidney outcomes (doubling of SCr and ESKD) compared with placebo. In EMPA-KIDNEY, the reduction was evident across the whole range of eGFR and most striking in patients with severely increased albuminuria. The rate of eGFR loss was lower with empagliflozin in all UACR subgroups [1395]. At least in patients with diabetes, these benefits have been ascribed to direct kidney effects including a reduction in intraglomerular pressure and antifibrotic effects that are independent from glucose lowering [1396,1397]. A mild eGFR drop was present during the first weeks of treatment, and it was managed as in the case of RAS blockers. In CREDENCE and DAPA-CKD, SGLT2i were also able to reduce the risk of some CV events, and in DAPA-CKD, the risk of mortality was reduced [1398], a benefit that was not previously evident with RAS blockade or any other drug treatment in the CKD population [1398–1401]. Several previous studies have evaluated the addition of a steroidal MRA (spironolactone or eplerenone) on top of an ACEi or an ARB in patients with proteinuric diabetic CKD and showed significant reductions in urine albumin or protein excretion [1402–1404], independently of the BP-lowering effect. This antiprotenuric effect is ascribed to the inhibition of several deleterious genomic and nongenomic effects of aldosterone breakthrough, including kidney tissue inflammation and fibrosis mediated through mineralocorticoid receptor axis overactivation [1405]. However, in clinical practice, use of steroidal MRA in CKD is restricted by absence of evidence from outcome kidney trials and the increased risk of hyperkalemia [1406]. Finerenone is a novel, nonsteroidal MRA with a different duration of action and tissue distribution than steroidal MRAs, that inhibits binding of different coregulatory molecules to mineralocorticoid receptors. This allows reduction of inflammatory and fibrotic processes, with less interference with the mineralocorticoid-mediated actions in the distal tubule, e.g. with less BP reduction and less potassium increase than with steroidal MRAs [742,1405]. The BP reduction associated with finerenone appears to be less than that associated with spironolactone, and does not seem to substantially contribute to the organ-protective effects of the drug [582]. Recently, evidence of the dose-dependent reductions of albuminuria by finerenone [1407] has been shown in two outcome-based RCTs. In the FIDELIO-DKD trial (5734 participants with type-2 diabetes, CKD and moderate or severe albuminuria), finerenone on top of ACEi or ARB treatment was associated with significant reductions in the risk of kidney failure, eGFR decline (>40%), renal death and CV outcomes [545,546]. The SBP/DBP difference was 2.7/1.0 mmHg in favor of finerenone, consistently across all different baseline BP groups [582]. Hyperkalemia leading to discontinuation of the trial regimen was 2.3% with finerenone and 0.9% with placebo, and no fatal hyperkalemia related adverse events were reported [546]. However, this finding comes from an RCT, and whether similar figures will be obtained in a real-world setting remains to be demonstrated. In the FIGARO-DKD trial (including 7437 participants) with characteristics similar to the FIDELIO-DKD trial [1408], finerenone was associated with a 13% significant reduction in the risk of CV death, nonfatal myocardial infarction, nonfatal stroke or hospitalization for HF (the primary outcome), with consistent beneficial effects on kidney outcomes and similar tolerability profile [547]. In an on-treatment analysis combining the patient population of both trials, finerenone reduced mortality by 18% compared with placebo [1409]. Other nonsteroidal MRAs (esaxerenone and apararenone) have also shown to significantly reduce albuminuria in CKD patients in phase 2 clinical trials [1410], but they have not yet been tested in pivotal kidney outcome studies. In view of the above evidence, it is recommended to use SGLT2is or finerenone in patients with CKD in addition to lifestyle interventions and antihypertensive drug therapy. Use of an SGLT2is should be considered either in patients with diabetic and in patients with nondiabetic CKD with a moderate or severe increase of albuminuria, while use of finerenone is recommended in patients with diabetic nephropathy and moderate or severe albuminuria. The order of addition of an SGLT2is or finerenone has not been tested in clinical trials and can be based on the individual patient characteristics, including the need for improvement of glycemic control, potassium levels or persistent albuminuria (Fig. 20).

19.2 Renovascular disease

Renovascular hypertension represents one of the most common forms of secondary hypertension, with a reported prevalence of around 2– 5% of all hypertensive individuals and up to 30% of patients with secondary hypertension [1411,1412]. Traditionally, atherosclerotic renal vascular disease and FMD are regarded to account for 90 and 10% of these cases, respectively [1413,1414]. However, the actual prevalence of renovascular hypertension may be considerably higher in selected cohorts of patients, such as those with resistant hypertension [1415]. Atherosclerotic renal vascular disease has been reported in 6.8% of individuals over 65 years of age, 10–12% of patients with ESKD [1416], 15– 30% of patients with CAD and up to 50% of patients with HF [1417,1418]. Renal FMD varies between <1 and 6% of individuals in observational studies [1419]. The overall prognosis of atherosclerotic renal vascular disease is unfavorable because the CV event rate after atherosclerotic renal vascular disease diagnosis can exceed that of the general population by a factor of 3– 6 [1416]. In addition, CKD patients with atherosclerotic renal vascular disease not receiving dialysis have a mortality rate 1.5 times higher than patients with other causes of CKD, the mortality of those on dialysis being three times higher [1420]. Within patients with atherosclerotic renal vascular disease, prognosis varies considerably depending on the underlying clinical status and comorbid conditions. Patients presenting with flash pulmonary edema have a two-to-three-fold higher risk of CV events and death, compared with patients with low-risk phenotypes (those without flash pulmonary edema, refractory hypertension, or rapid loss of kidney function) [1421].

While revascularization with balloon angioplasty without stenting is the treatment of choice for patients with FMD and hemodynamically significant renal artery stenosis [340], the optimal treatment of atherosclerotic renal vascular disease has been matter of considerable debate for several years [1422,1423]. After some early years of little evidence-based support of revascularization, a few large RCTs attempted to test the effects of standard medical therapy plus percutaneous transluminal renal angioplasty compared with medical therapy alone in patients with atherosclerotic renal vascular disease, and showed no significant differences in BP levels, adverse CV or kidney outcomes between the two groups [1424–1426]. However, these trials met with severe criticisms because of methodological limitations, such as nonstandardized inclusion criteria, poor assessment of the stenosis severity, enrolment delays, protocol revisions during the trial, high crossover rates and low event rates. Most importantly, these trials included mainly patients with mild/asymptomatic atherosclerotic renal vascular disease, mild hypertension or advanced CKD and excluded patients in whom the clinical presentation was highly suggestive of critical renal artery stenosis, i.e. those with flash pulmonary edema, refractory hypertension, or rapidly declining kidney function [1414]. Several observational studies in patients with well documented severe atherosclerotic renal vascular disease (≥70% stenosis) and high-risk clinical profiles document significant benefits of revascularization in terms of BP control, preservation of kidney function and reductions in the risk of CV events and death [1421,1427]. Thus, the current consensus is to offer revascularization on top of medical therapy in patients with documented secondary hypertension because of atherosclerotic renal vascular disease or high-risk clinical profiles and documented high-grade stenosis (≥ 70%) [1418,1422,1423]. Medical therapy alone could be used for individuals with asymptomatic atherosclerotic renal vascular disease with <70% stenosis, patients with mild or moderate hypertension that is easily controlled with antihypertensive drugs and low-grade stenosis, or patients with nonviable kidney parenchyma, where revascularization has little to offer. In the medically treated patients, if treatment initiation with an ACEi or an ARB results in eGFR reduction of 30%, careful re-evaluation of the patient is warranted.

19.3 Hypertension in patients with kidney transplantation

Kidney transplantation is considered the optimal choice for kidney replacement therapy in patients with ESKD because of improved survival and quality of life compared with dialysis. This survival benefit has been attributed to kidney function improvement and delayed progression of CVD [1428]. However, the residual CV risk remains significantly higher in kidney transplant recipients than in the general population, and CVD remains the leading cause of death in these patients during the 10 years posttransplant [1429]. Kidney transplantation ‘per se’ is associated with significant BP reductions (24 h BP -8/ -5 mmHg) in the short-term and mid-term posttransplant periods as well as with reduction in the use of antihypertensive agents [1430,1431]. ABPM values in transplanted patients are significantly lower than those of matched hemodialysis patients and similar to patients with CKD and matched kidney function [1432]. Despite these improvements, hypertension represents the most prevalent posttransplantation comorbidity, with ABPM studies estimating the presence of hypertension in >95% of the patients [1433]. Elevated BP is associated with kidney function decline, organ damage, CV events and reduced graft and patient survival [1405,1434,1435]. Several studies showed that misclassification of hypertension status by office BP is commonly encountered in these patients [1436], mostly because of a particularly high proportion of MH (20–40%) associated with a frequently impaired dipping status (around 50%) [1437] and high rates of nocturnal hypertension (up to 70–80%) [1438,1439]. Ambulatory BP is a much stronger predictor of kidney function decline and organ damage than office BP in kidney transplanted patients [1435], and thus use of ABPM is recommended for the management of hypertension in this condition.

The pathogenesis of hypertension in kidney transplanted patients is multifactorial, involving traditional risk factors, factors related to CKD (most commonly impaired sodium handling and activation of RAS and SNS) and factors related to transplantation and its treatment [1440]. Among major immunosuppressive classes, purine pathway inhibitors (mycophenolate mofetil or azathioprine), and mammalian target of rapamycin (mTOR) inhibitors (everolimus or sirolimus) do not affect BP control [1440,1441]. Glucocorticoid treatment is known to increase BP, and avoidance or withdrawal protocols in kidney transplanted patients are associated with a better BP profile [1442,1443]. Calcineurin inhibitors (cyclosporine or tacrolimus) are also associated with BP elevations, through increased sodium reabsorption via the Thiazide-sensitive sodium chloride co-transporter in the distal convoluted tubule, the increase of vasoconstrictive substances leading to increase of total peripheral resistance and vasoconstriction of afferent arteriole at the preglomerular site [1440,1441], and, for cyclosporine, a marked activation of sympathetic nerve firing [1444]. The effects of tacrolimus on BP appear less pronounced compared to cyclosporine.

BP targets for hypertension management in kidney transplanted patients are extrapolated from data in CKD populations, as there are no specific RCTs that have tested different BP targets on major clinical endpoints. Yet, a target SBP of <130 mmHg is considered as a reasonable target for kidney transplant patients. Lifestyle modifications should be adopted on the basis of general recommendations for CKD, and drug combinations between major antihypertensive agents should be employed in most patients. Special benefits by ACEis/ARBs are not clearly established, because available studies provide conflicting results [1405,1441], although, in a recent meta-analysis, the risk of graft loss was reduced by 38% with ACEis/ARBs (9 studies, 1246 participants) without, however, any significant effect on nonfatal CV outcomes or death, and with an increased incidence of hyperkalemia [1445]. In kidney transplanted patients, DHP-CCBs have been consistently associated with benefits such as improved graft survival and minimization of the vasoconstrictive effects of calcineurin inhibitors at the preglomerular site. In the aforementioned meta-analysis, CCBs reduced the risk for graft loss by 42% (16 studies, 1327 participants), while in head-to-head comparisons with ACEis/ARBs, CCBs significantly increased GFR by 11.11 ml/min [1445]. Thus, CCBs can be preferentially used in the early posttransplantation period. Thiazide/Thiazide-like diuretics are also effective and useful in patients with kidney transplantation, because they block the cyclosporine-mediated sodium retention. At present, no data are available on the effect of antihypertensive drugs on long-term graft loss. Transplant renal artery stenosis is not uncommon in patients with kidney transplantation, and it should be effectively sought for in cases of uncontrolled or abrupt onset hypertension [1440]; PTRA has high success rates in these patients [1446].

Treatment strategies in patients with kidney disease

figure62
figure63

20. HYPERTENSION AND OTHER SELECTED COMORBIDITIES

20.1 Obesity

Obesity and arterial hypertension commonly occur in the same patients and often have type 2 diabetes has a third associated condition. Hypertensive obese patients may require more antihypertensive medications to have their BP controlled than nonobese individuals and are more likely to exhibit treatment-resistant hypertension [1447–1449], and metabolic side effects of antihypertensive medications may be particularly relevant in this population. There is a paucity of hypertension trials specifically dedicated to hypertensive, obese patients and data on combination of antihypertensive treatments and body weight management are also scarce. Nevertheless, based on available trials, observational studies and trial subgroup analyses, the following considerations and recommendations can be made.

20.1.1 Antihypertensive pharmacotherapy in obesity

When treating patients with obesity and hypertension, a sensible goal is to attain BP reduction without worsening obesity and associated metabolic risks. Although there is no evidence that BP targets differ between patients with and without obesity, achieving the target BP value is more difficult in the presence of obesity. For example, a retrospective analysis of the ALLHAT trial stratified patients according to BMI in normal weight (BMI <25 kg/m2), overweight (BMI 25– 29.9 kg/m2) and obese (BMI >30 kg/m2) groups [1450]. After 5 years of treatment, two-thirds of the patients achieved BP control, regardless of BMI, but patients with obesity required a larger number of medications [1450]. In patients with obesity and hypertension, it may be prudent to initiate antihypertensive therapy with ACEis, ARBs or CCBs, because none of these drug classes worsens insulin sensitivity or adiposity. RAS inhibitors may ameliorate glucose metabolism, although they do not necessarily prevent the greater risk of obese patients to develop type 2 diabetes mellitus [1451,1452], because of the close and common association of obesity with insulin resistance, a precursor of type 2 diabetes [1453]. Thiazide/Thiazide-like diuretics, particularly at higher doses, may worsen glucose and lipid metabolism and increase the risk of type 2 diabetes, while BBs promote weight gain, have dyslipidemic effects, and may also increase the risk of type 2 diabetes, the diabetogenic influence being greater for the BB plus diuretic combination. These disadvantages are attenuated or absent with vasodilating BBs [1454–1456]. Potential metabolic side effects of these drugs in hypertensive dyslipidemic patients have to be weighed against their efficacy in lowering BP, because in many obese patients, adequate BP control can only be obtained by combination therapy, and diuretics are especially effective BP-lowering agents [1450], possibly because obesity is associated with abnormal salt handling and volume expansion [1457,1458]. The latter may be explained in part by reduced natriuretic peptide actions in obesity [21]. Furthermore, in patients with obesity, hypertension may be associated with comorbidities such as postmyocardial infarction or HFrEF, for which BBs are a compelling indication. A potential advantage of BBs is that they may attenuate the CV influence of the SNS, which is activated in patients with obesity, and more so in patients with obesity and hypertension [1459,1460]. In this context, a subgroup analysis of the ACCOMPLISH trial has shown that the risk of CV outcomes did not differ between obese patients treated with an ACEi either combined with a CCB or a diuretic [1461] supporting the use of diuretic-based combinations in this clinical condition, in which the efficacy of BP-lowering may outbalance the potential metabolic disadvantages of this drug class. Other drug classes with antihypertensive effects have been tested in smaller trials in patients with obesity, including renin inhibitors and sacubitril/valsartan [1462,1463].

20.1.2 Role of nonpharmacological weight loss intervention

Strong evidence is available that in addition to improving metabolic risk factors, weight loss can have a beneficial effect on BP control [1464]. In individuals with high-normal BP randomized to a nutritional intervention, 2.7 kg body weight loss was associated with an incidence of hypertension of 8.8% compared with 19.2% in the control group [1465]. In the TOHP Phase I study [1466], men and women aged 30– 54 years who lost 3.9 kg weight showed a SBP/DBP decrease of 2.9/2.3 mmHg over 18 months, with a stronger reduction in severely obese patients. In TOHP Phase II, individuals who lost at least 4.5 kg after 6 months and maintained their weight reduction over 30 months had the greatest BP decrease and a 65% reduction in the risk of developing hypertension [1467]. In the TAIM trial, on overweight and obese mild hypertensive patients, body weight and DBP decreased by 4.7 and 8.8 mmHg, respectively, after 6 months under hypocaloric diet compared with control groups under antihypertensive drug treatment, a low-sodium/high-potassium diet, a usual diet or placebo [1468]. In the TONE trial, moderate weight loss decreased the need for antihypertensive therapy by approximately 30% in older hypertensive patients on a single antihypertensive drug [1469]. The BP-lowering effects of weight reduction achieved by dietary approaches are also supported by meta-analyses [1470], and evidence is available that loss of body weight reduces sympathetic activity [367], while reversing abnormalities in natriuretic peptide regulation [1471,1472]. Thus, lifestyle interventions aimed at reducing body weight are recommended in patients with obesity and hypertension through low caloric diets and increased physical activity. Involvement in the treatment plan of dieticians may be helpful. A problem is, however, that even when concerned teams are made available and therapeutic plans are implemented, relatively few patients with obesity are able to sustain weight loss in the long-term.

20.1.3 Role of weight loss medications

Very few currently available weight loss medications have been specifically tested in obese hypertensive patients, having BP reduction as the primary goal. Newer weight loss drugs that have been evaluated for CV protection in large-scale trials on overweight and obese type 2 diabetic patients have not been shown to improve CV outcomes in nondiabetic patients with overweight or obesity. Therefore, the role of such medications in managing BP in obese hypertensive patients is unclear. A recent meta-analysis has shown that not all medications that reduce body weight improve BP control in patients with hypertension [372], although in one study combining orlistat with a hypocaloric diet was more effective than diet alone to achieve BP control in patients with obesity and hypertension [1473]. Furthermore, in other studies, low-dose topiramate/ phentermine, which is not approved in Europe, reduced body weight and BP in obese hypertensive patients [1474,1475]. For other drugs, such as liraglutide or lorcaserin, BP responses have been modest or absent in the case of naltrexone/ bupropion [372,1476–1478]. Once weekly treatment with the GLP-1 RA semaglutide decreased body weight more compared with daily liraglutide administration and associated with a greater decrease in SBP and DBP as well [1479]. Seventy-two-week treatment with once weekly tirzepatide, a dual glucose-dependent insulinotropic polypeptide (GIP) GLP-1 RA lowered body weight in a dose-dependent fashion, with 20.9% body weight reduction at the highest 15 mg dose of the drug compared with 3.1% weight reduction under placebo [1323]. Compared with the placebo group, SBP in the pooled tirzepatide group decreased by 6.2 mmHg [1323]. It is important to mention that some antidiabetic agents favor weight loss [1480]. Metformin treatment is accompanied by a modest weight loss [1480,1481]. A consistent albeit moderate reduction in body weight has been also observed with SGLT2i treatment. Overall, weight loss medications should not be primarily prescribed for the management of hypertension in patients with obesity. However, when prescribed for other reasons, BP reduction can be an added benefit depending on the drug class. Indeed, particularly GLP-1 RA and dual GIP/GLP-1 RA are important novel drugs that are approved or designated for approval for the treatment of patients with a BMI of at least 27 kg/m2 but less than 30 kg/m2 (overweight) and hypertension in addition to their approval in type 2 diabetes or obesity.

20.1.4 Role of bariatric surgery

Bariatric surgery comprises various procedures such as sleeve gastrectomy, Roux-en-Y gastric bypass or biliopancreatic diversion with the common goal of reducing body weight. Procedures differ in terms of weight reduction, adverse effects and weight-independent effects on gut hormones. The nonrandomized Swedish SOS study showed that weight reduction through bariatric surgery improves CV morbidity and mortality in men and in women with a BMI of at least 34 or 38 kg/m2, respectively [1482]. A recent meta-analysis comprising 269 818 patients submitted to bariatric surgery and 1 270 086 control patients confirmed these findings [1483]. Bariatric surgery is also followed by a sustained reduction of metabolic and other risk factors, which improve overall CV risk profile [1483–1485]. A long-lasting reduction of sympathetic nerve activity [1486] and increased natriuretic peptide availability [1487,1488] may contribute to the beneficial effect on BP and risk. It has been remarked that long-term BP reduction following bariatric surgery is relatively modest for the large reduction in body weight [1483]. Along the same critical line, an analysis of SOS suggested that weight loss through bariatric surgery decreased the risk of hypertension only in the first few years following the procedure [1489]. However, the above-quoted meta-analysis has shown a substantial reduction in the incidence of hypertension in the bariatric surgery group [1483]. Furthermore, a more recent randomized trial in treated hypertensive patients with a BMI of 30– 39.9 kg/m2 has reported that compared with medical therapy, the Roux-en-Y gastric bypass surgery group required fewer antihypertensive medications to maintain BP below 140/90 mmHg for up to 3 years [1490] following surgery. In addition, after 1 year, no patients in the surgery group and 14.9% patients in the medical group exhibited treatment-resistant hypertension [760]. Thus, patients with obesity and hypertension who undergo bariatric surgery often appear to experience improvements in BP control, require fewer antihypertensive medications and have less risk to develop hypertension over time, although doubts still exist on whether the risk reduction persists indefinitely [1489]. Considering the risks associated with surgery and the limited amount of data, bariatric surgery should not be considered primarily for the management of hypertension. However, improved BP control appears to be an added benefit in patients with obesity submitted to bariatric surgery.

Hypertension management in obesity

figure64

20.2 Obstructive sleep apnea

Sleep disorders such as a reduction of sleep to <6 h per night [1491] are included in the list of risk factors for development of hypertension and increased CV risk. The most widely studied sleep disorder is OSA, which can be found in a considerable number of patients with difficult-to-treat or resistant hypertension [726], which favors this condition via multiple and complex pathophysiological mechanisms [1492]: hypoxia during night-time episodes of hypopnea and apnea, sympathetic overdrive, systemic vasoconstriction, oxidative stress and systemic inflammation [1493]. Hypertension mediated by OSA is often associated with MH, higher BP values during the night or nocturnal hypertension and a nondipping status [1494]. To reduce BP in patients with OSA, all major classes of antihypertensive drugs can be used. BP reduction has been reported also with the use of MRAs [1495]. Continuous positive airway pressure (CPAP) application has been shown to induce small reductions (about 3 mmHg for) in office and 24 h SBP, the 24 h BP reduction including day-time and night-time BP values [1496]. Patients that may benefit more from CPAP treatment have been reported to be younger than 60 years, with an uncontrolled BP before treatment initiation and with severe oxygen desaturation at baseline [1496]. The BP-lowering effect of CPAP can be seen also in patients with resistant hypertension and it has been found to be almost double for the nightime than to the daytime BP [1497], due to reduced sleep segmentation and improved intermittent hypoxia [1498]. Reduction of arterial stiffness, decrease of high-sensitive C-reactive protein, plasma cortisol and noradrenaline levels have also been associated with use of CPAP [753,1499,1500].

20.3 Asthma

Hypertension and asthma are common diseases frequently encountered together in the same patient. Epidemiological studies indicate an increased prevalence of hypertension in asthmatic patients compared with patients without asthma [1501]. It has also been reported that hypertension is associated with augmented asthma severity, reduced lung function and reduced forced expiratory volume in 1 s (FEV1) as a marker of CV mortality, independent of the smoking history [1502]. The hypothesis has been advanced that the bidirectional relationship between hypertension and asthma can at least in part reflect underlying low-grade systemic inflammation, and an inverse relationship between C-reactive protein levels and decreased FEV1 has indeed been reported [1501]. The pathophysiological interplay between hypertension and asthma may also result from the concomitance of other comorbid conditions. A meta-analysis of prospective studies showed that a condition closely related to hypertension such as obesity is an important risk factor for asthma and asthma-related morbidity [1503,1504]. Obese patients with asthma experience more frequently severe exacerbations of the disease and a reduced response to asthma medications, possibly via an increased production of pro-inflammatory cytokines and systemic inflammation [1503,1504]. The OSA syndrome is another asthma-related factor associated with hypertension and systemic inflammation. OSA was found to be more prevalent in asthma patients as well as an independent risk factor for poor asthma control [1501,1503–1505]. A combination of genetic factors, age, emotional stress, diet and lifestyle characteristics also contribute to the hypertensive asthmatic phenotype and may predispose patients with asthma to hypertension [1501,1506,1507].

Patients with hypertension and asthma constitute a subgroup of patients in whom treatment of either condition is more difficult and the risk of developing CV events is elevated [4,1501,1508]. Lifestyle modifications appropriate to both conditions should be implemented, smoking cessation being of obvious primary importance. For the drug treatment of hypertension, CCBs appear to be particularly suitable, as they may favor bronchial smooth muscle relaxation. Among RAS blockers, ARBs should be preferred because of the risk of developing cough during treatment with ACEis, which may be particularly disturbing in asthmatic patients. It is recommended to avoid BBs for antihypertensive treatment in patients with asthma because the safety margin of these drugs is smaller than in chronic COPD, where BBs are safe [1509], although this is a controversial issue [1510]. Nevertheless, a cautious use of BBs in patients with compelling indications to balance individual benefits and risks appears reasonable. Finally, treatment of asthma with beta-adrenergic agonists and corticosteroids may induce adverse CV effects by increasing heart rate and BP [1501,1508].

20.4 Obstructive pulmonary disease

Hypertension is the most common comorbidity in patients with COPD and both comorbidities are independently associated with increased risk of CV events [1511]. Patients with COPD are characterized by the presence of chronic systemic inflammation, which plays a role also in the pathophysiology of comorbidities that are frequently associated with COPD [1511], including hypertension. In hypertension, COPD also promotes HMOD such as endothelial dysfunction, atherosclerosis and cardiorenal damage [1511]. A common CV risk factor that is a major cause of COPD, i.e. smoking, also contributes to BP elevation and CV risk in hypertension [1512,1513]. Management of hypertension in patients with COPD should consider the effects of antihypertensive drug classes on impaired respiratory function, adverse pulmonary outcomes, including not only COPD exacerbations but also overall CV outcome and mortality [4,1511]. Interactions between antihypertensive drugs and agents used for treating COPD such as bronchodilators and corticosteroids should also be taken into account, because of their possible pressor and tachycardic effects [1511]. In this regard, the general recommended strategy for antihypertensive drug treatment applies also to COPD. This applies also to BBs, the use of which is upgraded compared with previous guidelines [4], and treatment recommendations in general. While in the past, BBs were contraindicated for COPD patients, because of their bronchoconstrictive effects, this view has now changed. A systematic review and meta-analysis of 49 studies involving more than 670 000 patients indicated that treatment of patients with COPD and CVD with both beta-1 selective and nonselective BBs significantly lowered heart rate (about 8 bpm) and reduced all-cause mortality as compared with COPD patients with no BB treatment. Additionally, use of beta-1 selective BBs, but not of nonselective ones, reduced COPD exacerbations [1514]. Thus, in COPD patients, hypertension and CVD should be treated, if tolerated with more beta-1 selective BBs, to reduce both mortality and COPD exacerbations (Table 16).

20.5 Gout and uric acid

The prevalence of hypertension in patients with gout is twice as high compared with patients without gout (36 versus 17%), while hyperuricemia (with or without gout) can be found in more than one out of four hypertensive patients [1515,1516]. Furthermore, hypertension is independently predicted by greater serum uric acid levels [1517,1518] and their predictive role can be seen also in childhood and adolescence, while being almost completely lost in the older population as well as in patients with secondary hypertension and normal kidney function [1519]. Gout is clearly associated with an increased risk of CV events (including myocardial infarction and stroke) [1520] through mechanisms that are not extraneous to the pathophysiology of hypertension and its associated CV and kidney complications, because inflammation and oxidative stress are common drivers of both diseases [1521]. Prevention of gout by lowering serum uric acid with allopurinol or xanthine-oxidase inhibitors has been reported to be associated with a small BP reduction [1522–1524] and achievement of serum uric acid level <6.0 mg/dl are recommended by recent guidelines on gout treatment [1525]. Drugs used for the treatment of gout flares, i.e. colchicine, NSAIDs and corticosteroids, may negatively impact on BP values and control in hypertensive patients, which means that under these circumstances, both office and out-of office BP monitoring should be intensified. The recently reported preventive role of colchicine against atherosclerotic disease [1526], presumably via reduction of its contribution to inflammation and atherosclerosis, requires confirmation. This is required also for the CV-protective effect of reducing serum uric acid by antiuricemic agents, which has been suggested by earlier studies [1522,1524] but recently denied by a RCT with allopurinol in patients with CAD [1527,1528].

Antihypertensive drugs have a well known differential effect on serum uric acid. Serum uric acid level increases with the use of Thiazide/Thiazide-like and loop diuretics [1529]. ACEis, ARBs, CCBs and BBs have no effect, although a reduction of kidney excretion of uric acid has been reported for BBs [1530]. Among ARBs, losartan has been shown to reduce serum urate levels through an uricosuric effect, with some favorable implications for CV outcome [1531]. Together with CCBs, losartan has also been shown to reduce the incidence of gout in hypertensive patients, regardless of the BP level, in a large nested case– control study [1530]. In line with other guidelines [1525], these guidelines suggest to prescribe diuretics with caution in patients with gout but not to avoid them if diuretics are needed to achieve BP control. In general, physicians should try to use lower doses of diuretics because the effect of these drugs on serum uric acid is dose-related.

20.6 Immune-mediated inflammatory diseases

Immune-mediated inflammatory diseases, including diseases such as rheumatoid arthritis, psoriasic arthritis or lupus erythematosus, are associated with an increased prevalence of hypertension that is often underdiagnosed and poorly controlled [1532–1534]. The increase in CV risk associated with immune-mediated inflammatory diseases is only partially related to traditional CV risk factors and possibly attributable to inflammatory changes in the vasculature [1533]. The increase of the overall CV risk in patients with immune-mediated inflammatory diseases can be estimated by available risk calculators [1525].

20.6.1 Rheumatoid arthritis

Rheumatoid arthritis is associated with an increased risk of CV disease [1535,1536]. According to some reports, the prevalence of hypertension in the unselected rheumatoid arthritis population is not greater compared with nonrheumatoid arthritis controls [1537]. However, this has not been confirmed by other studies that included hypertension as a baseline covariate, which have found an increased prevalence of hypertension or higher BP values in rheumatoid arthritis [1538]. Hypertension was reported to contribute to CV risk similarly in rheumatoid arthritis and in the general population [1539], although this is not entirely in line with the relationship between the duration of the disease and the progression of organ damage [1540]. Most of the currently used disease-modifying antirheumatic drugs do not seem to have substantial effects on BP. However, several other agents used in the symptomatic treatment of rheumatoid arthritis patients appear to raise BP [1541]. Paradigmatically, NSAIDs and glucocorticoids raise BP [1515] and may cause clinically significant hypertension or impairment of BP control in treated hypertensive patients. BP should be lowered in rheumatoid arthritis as in the general population, preferentially with CCBs and RAS inhibitors because of the evidence of an overactive RAS [1542] in this disease. Underlying diseases should be managed by reducing inflammation and by avoiding high doses of NSAIDs [1515].

20.6.2 Psoriatic arthritis

Psoriatic arthritis is associated with an increased cardiometabolic risk leading to an excess of CV disease [1543]. Psoriasis is a common, chronic inflammatory disease of the skin and joints that affects 2–4% of the general adult population. Recent studies suggest that psoriasis, particularly if severe, may be an independent risk factor for atherosclerosis, myocardial infarction and stroke [1544]. Hypertension is prevalent in patients with psoriatic arthritis (20– 25%) [1544] along with obesity and several metabolic abnormalities involving glucose and lipid profiles. In patients with psoriatic arthritis, BP should be lowered according to the general antihypertensive treatment recommendations, preferentially with RAS inhibitors and CCBs. BBs may trigger or worsen psoriasis and should be avoided if possible or carefully used in the presence of compelling indications [1545] (Table 16). Glucose control should be considered in patients with hypertension and psoriatic arthritis [1546].

20.6.3 Systemic lupus erythematosus

Patients with systemic lupus erythematosus have a higher burden of CV risk factors compared with the general population, and this is responsible for the high prevalence of premature CVD in the affected patients [1547]. CVD is an important cause of both morbidity and mortality in systemic lupus erythematosus resulting from a combination of factors with prothrombotic and/or atherogenic properties. In a recent meta-analysis, the relative risk of CVD was significantly elevated in patients with systemic lupus erythematosus compared with the general population [1548]. Studies differ as to the risk factors that are more frequently represented and prognostically relevant in the systemic lupus erythematosus population. A multivariate analysis identified age (relative risk 1.03), hypertension (1.71), smoking (1.48), diabetes (2.2) and dyslipidemia (2.18) as significant risk factors for CV events in this condition. This can help in the adoption of appropriate preventive measures, including BP control [1549]. Apart from small old uncontrolled studies [1550,1551], no trial targeting hypertension in systemic lupus erythematosus is available. The 10-year survival rate of systemic lupus erythematosus patients has significantly improved during the last half century, now standing at 92% [1552]. This improved prognosis is because of early recognition of milder cases and improvement in both specific and general medical care, e.g. treatment with immunosuppressive drugs, kidney replacement therapies and treatment of comorbidities, including hypertension.

20.7 Glaucoma in hypertension

Glaucoma is a leading cause of irreversible blindness affecting more than 60 million people worldwide [1553]. Primary open-angle glaucoma (POAG) is the most common type of glaucoma, accounting for more than 90% of all glaucoma cases [1554]. Normal tension glaucoma is a common form of POAG with no elevation of the intraocular pressure. Development and progression of glaucoma is influenced by complex interactions between arterial BP, intraocular pressure and ocular perfusion pressure. The mean ocular perfusion pressure is the difference between the pressure in the central retinal artery and vein [1555]. The central retinal arterial perfusion pressure can be determined by the mean arterial pressure value and the intraocular pressure [1555]. High BP may lead to high intraocular pressure with increased production of aqueous humor attributed to high ciliary blood flow and capillary pressure as well as low aqueous outflow due to increased episcleral venous pressure [1556,1557]. In contrast, low-BP, either spontaneous or due to antihypertensive therapy, may decrease ocular perfusion pressure, but this may result in ischemic damage of the optic nerve [1557–1560]. Accordingly, in patients with normal tension glaucoma, deterioration of the visual field may still progress despite an intraocular pressure in the normal range.

A systematic review and meta-analysis assessed the association between BP and POAG as well as intraocular pressure [1556]. Both SBP and DBP were positively associated with intraocular pressure, with a 0.26 and 0.17 mmHg increment per 10 and 5 mmHg increase in SBP and DBP, respectively. Hypertension was significantly associated with POAG in both cross-sectional [1556] and longitudinal studies [1561]. Additionally, a large population-based study including 41 235 hypertensive patients with glaucoma from a National Danish Registry found that initiation of antihypertensive treatment postponed the onsite of glaucoma by 2 years [1562].

Zhao et al.[1556] investigated the association of BP as a continuous variable with POAG and observed a J-curve relationship, indicating both low and high BP as risk factors for glaucoma. In this regard, it is important to know that intraocular pressure also varies according to change of the body position. In the supine position, and thus during sleep, an increase in intraocular pressure is observed leading to a decrease in ocular perfusion pressure [1559]. Nocturnal hypotension as recorded by ABPM strongly correlated with glaucoma progression [1563]. Further evidence showed that the link between nocturnal BP and glaucomatous optic neuropathy was driven by an adverse association of the nerve damage with extreme dipping, i.e. an excessive nocturnal BP reduction phenotype [1564]. In addition to circadian BP variations and body position, antihypertensive treatment can also influence ocular perfusion pressure. A study examining the timing of antihypertensive drug administration showed that patients with glaucoma taking antihypertensive drugs in the evening had lower nocturnal BP, more pronounced dipping, lower nocturnal ocular perfusion pressure and greater visual field loss [1565]. Moreover, different antihypertensive drugs may have a different impact on intraocular pressure and glaucoma. Use of BBs has been associated with a lower intraocular pressure [1566,1567] as well as a reduced risk of POAG [1568]. The reduced risk of POAG is in agreement with the well established intraocular pressure reduction by BBs through blockade of sympathetic nerve endings in the ciliary epithelium. This causes a fall in aqueous humor production and makes topical therapy with BBs a mainstay in glaucoma management [1569]. In this regard, it is worthwhile mentioning that topical treatment of glaucoma with BBs may also induce systemic effects, including side effects such as bradycardia, due to absorption of the drugs [1570]. RAS inhibitors have been associated with higher intraocular pressures [1566] and increased risk of POAG [1568]. However, in the above-mentioned National Danish Registry, all antihypertensive drugs, except vasodilators (e.g. hydralazine), delayed the progression of glaucoma [1562].

Taken together, not only hypertension but also low BP values, an extreme nocturnal dipping status and high 24 h BP variability leading to frequent BP peaks and dips (especially dips) are all associated with an increased risk of POAG. Therefore, in hypertensive patients with glaucoma the BP targets recommended for the general hypertensive population should be pursued with caution, and use should be made of ABPM to avoid low nocturnal BP, extreme dipping and thus glaucoma progression.

Glaucoma and hypertension

figure65

20.8 Hypertension oncology

20.8.1 Hypertension and its association with cancer

The association between hypertension and cancer is bidirectional with overlapping risk factors (e.g. unhealthy diet, alcohol intake, physical inactivity, smoking, increased BMI) and pathophysiological mechanisms (e.g. immunoinflammation and oxidative stress) involved in both conditions [1571,1572]. Particularly for renal cell carcinoma (RCC), hypertension has been proposed as an independent risk factor, although the estimated risk ratios range between a 2.5-fold increase in patients with grade 2 hypertension [1573] to an overall smaller risk estimate of 1.12-fold only [1574], while confounding cannot be completely excluded [1572]. A Mendelian randomization study found DBP to be an independent risk factor for RCC, while the role of SBP is less clear [1575]. The association between hypertension and other types of cancer is even less clear and a direct causative association is unproven [1571,1572]. The same applies to the potential role of antihypertensive drugs for cancer development as discussed in Section 11.10.10.3.

Due to its high prevalence, preexisting hypertension is the most common comorbidity in patients with cancer, particularly in older patients, in parallel with the high prevalence of hypertension in the old age (see Section 3.3). Careful BP monitoring and hypertension management in cancer patients is recommended, because these patients are at risk of hypertension-related complications, including hypertension emergencies [1576] (see Section 16.2), and long-term CV risk increases in cancer survivors [1571,1572].

20.8.2 Hypertension induced by cancer treatments

Data on incident hypertension in cancer trials are difficult to analyze due to differences in the definitions of hypertension used. The Common Terminology Criteria for Adverse Events (CTCAE) in cancer therapy uses a BP grading system with four grades starting in the normotensive range according to the definition of these guidelines. Grade 1 is defined by a SBP of 120– 139 mmHg or a DBP of 80– 89 mmHg [1577].

20.8.2.1 Hypertension induced by VEGF inhibitors

Anticancer drugs and adjunctive therapy used in oncology can induce de novo hypertension or contribute to worsening preexisting hypertension. BP elevation during therapy with various cancer drugs have been known for a long time. However, this problem has recently come to special attention due to the development of inhibitors of the vascular endothelial growth factor (VEGF), a pathway that causes some BP increase in virtually every patient [1571,1572]. Drugs targeting the VEGF pathway are used for the treatment of various cancers (e.g. renal, hepatocellular, thyroid, gastrointestinal stromal cancer) [1578] (Table 26). VEGF inhibitors include i.v. administered monoclonal antibodies (VEGF-A monoclonal antibody, bevacizumab; VEGF-R2 monoclonal antibody, ramucir umab; VEGF-R1/R2 fused to Fc portion of IgG1, aflibercept), or oral small-molecule receptor tyrosine kinase inhibitors (RTKIs) targeting the VEGF-R1–33 and other receptors which mediate the downstream cellular signaling pathways essential for tumor cell survival (Table 26) [1571,1578,1579]. All types of VEGF inhibitors increase BP [1579–1583] via multiple mechanisms that have a similarity with the pathophysiology of preeclampsia, including activation of the ET-1 pathway, decrease in NO bioavailability, capillary rarefaction with reduction in microvascular flow, activation of the renal epithelial amiloride-sensitive Na channel (ENaC) and increased salt-sensitivity. For this reason, use of low-dose aspirin, which is recommended for the prevention of preeclampsia (see Section 16.1) [1068–1071], has been also proposed for cancer patients who develop complications in response to VEGF inhibitor treatment [1571]. The effects of VEGF inhibitors on BP are classified as on-target effects because they reflect also anticancer treatment efficacy [1571]. Hypertension is the most commonly reported adverse event during VEGF inhibitor therapy, with a dose-dependent increase of BP, which is usually reversible and can become manifest within days from the initiation of the anticancer therapy [1580,1584], highlighting the importance of regular BP surveillance, at best with HBPM [1580,1584]. Even though BP increases can be seen in virtually all patients treated with VEGF inhibitors, the incidence of hypertension is highly variable due to (i) the presence or absence of preexisting hypertension and control BP status; (ii) the variable between-drug potency to block the VEGF pathway and (iii) the large between-patient pharmacokinetic variability in patients treated with the same VEGF inhibitor [1579,1580,1585]. The highest incidence of hypertension was observed with the potent RTKI axitinib and the multikinase inhibitor lenvatinib, among whom 13 and 43%, respectively, developed severe hypertension [205,1585,1586]. The use of VEGF inhibitors is associated with an increased risk of incident and worsening HMOD and of other hypertension-related complications including LV dysfunction and HF, cardiac arrhythmias, kidney damage and both arterial and venous thrombotic events [205,1586]. Cardiac damage may lead to severe HF and cardiac death [1571,1587]. The risk of kidney complications includes acute kidney injury with thrombotic microangiopathy [1588] or glomerular damage leading to massive proteinuria [1571,1572,1589]. However, even though the relative risk of these adverse events is high compared to placebo and other anticancer drugs, the absolute risk increase of the aforementioned complications remains low during the time-frame of treatment with VEGF inhibitors.

T26
TABLE 26:
Hypertension induced by selected anticancer treatments
20.8.2.2 Hypertension induced by other anticancer drugs

A list of additional anticancer drugs that can also induce hypertension is presented in Table 26[1571,1572]. The list includes classic chemotherapeutic compounds such as cyclophosphamide and more recently developed therapies such as proteasome inhibitors (bortezomib, carfilzomib), poly ADP ribose polymerase (PARP) inhibitors (olaparib, niraparib) and Bruton's tyrosine kinase inhibitors (ibrutinib, acalabrutinib). The last class increases also the risk of AF and HF [205,1586]. The mechanism by which these therapies increase BP is most likely multifactorial and remains to be elucidated [1590]. It is still unclear whether immune check-point inhibitors (ICIs), which include monoclonal antibodies against the inhibitory programmed death-1 (PD-1) receptor or its ligand (PD-L1), and antibodies against the cytotoxic T lymphocyte antigen-4 (CTLA-4) receptor, increase the risk of hypertension. Possibly, they synergistically increase CV toxicity when combined with VEGF inhibitors [1590]. Abiraterone acetate is a CYP17A1 inhibitor that decreases androgen synthesis and is thus used in patients with progressive metastatic prostate cancer. This drug can also increase BP because of the inhibition of CYP17A1 upstream in the steroid synthesis at the level of adrenal glands, which leads to the synthesis of excessive amounts of 11-deoxycorticosterone via the counter-regulatory stimulation of ACTH release in response to a decrease in cortisol synthesis [1591]. The binding of 11-deoxycorticosterone to the mineralocorticoid receptor induces hypertension with hypokalemia, sodium and fluid retention and decreases the plasma levels of renin and aldosterone [1592] A low dose of prednisone is usually prescribed concomitantly with abiraterone to reduce the stimulation of ACTH release and may control the hypertension. If needed a MRA can be added [1592].

20.8.2.3 Hypertension induced by adjunctive therapies, radiotherapy or surgery

Adjunctive therapies with glucocorticoids, erythropoietin (EPO) to treat anemia and NSAIDs to treat pain and/or inflammation, can contribute to the development of hypertension or the worsening of hypertension in cancer patients. Radiotherapy to the neck and/or extensive neck dissection surgery can cause baroreflex failure (see Section 14.10) leading to extreme BP variability with dramatic hypertensive surges, hypotensive episodes, and orthostatic hypotension in some patients [907–909]. In addition, abdominal radiotherapy has been associated with renal artery stenosis, which occasionally can cause renovascular hypertension [1593]. Moreover, during long-term FU, an increased risk for CV events has been noted in patients who have undergone radiotherapy for various malignancies including lymphoma, breast cancer and neck-cancer, for which radiation injury of the vasculature has been proposed as an underlying mechanism [1594].

20.8.3 Management of hypertension in cancer patients

20.8.3.1 BP monitoring and general management before start of cancer treatment

Before starting cancer therapy, BP control in patients with preexisting hypertension should be confirmed by appropriate office BP measurement as recommended in Section 4 and also include the use of out-of office (ABPM or HBPM) measurements whenever possible (Fig. 9). The use of HBPM is particularly important and should be encouraged with education on proper use (see Section 4.7), because HBPM can play an important role for BP monitoring both during cancer treatment and during FU after cancer treatment. Furthermore, BP measurements should be also performed in patients without previously documented hypertension in order to exclude hypertension and to document the baseline BP values before patients are exposed to anticancer drugs that may induce hypertension (Table 26). Control of pain and anxiety before BP measurements requires special attention in cancer patients. Unfavorable changes of individual risk factor profiles (e.g. increased stress, incident depression, impairment of sleep, Table 2) and lifestyle factors (e.g. unhealthy diet, increased alcohol intake, Section 7) may have contributed to worsening of hypertension and BP control in cancer patients and should be addressed. Drug therapy should be intensified and adapted if needed (see below) according to the general recommended strategy (Fig. 12).

BP monitoring and risk assessment in cancer patients

figure66
20.8.3.2 General BP-lowering therapy and management during cancer therapy

Patients with active cancer were regularly excluded from outcome-based RCTs in hypertension. Conversely, in cancer trials, patients with uncontrolled hypertension or elevated BP are commonly excluded. Consequently, no evidence from outcome-based RCTs is available to guide overall management and drug therapy for hypertension in patients with cancer. Mindful of these limitations, these guidelines recommend that, by extrapolation, the definition of hypertension, BP thresholds and targets, lifestyle interventions and drug treatment strategies recommended for the general hypertensive population should be also applied to cancer patients. In severely ill cancer patients, treatment of hypertension should be individualized according to symptoms, comorbidities and polypharmacy in a shared decision-making process. Although there is no general consensus, these guidelines recommend that the threshold for withholding anticancer therapy should be a SBP ≥ 180 mmHg and/or a DBP ≥ 110 mmHg [1572,1577]. Thus, initiation of anticancer therapy should not be deferred in patients with uncontrolled BP unless they are symptomatic or present with grade 3 hypertension. In these patients, measures to control BP and symptoms should be initiated first, ideally by a team-based multidisciplinary approach, to allow initiation of anticancer therapy as early as possible.

Some considerations on drug selection and use specific for cancer patients during active cancer treatment need to be emphasized:

  • 1. Thiazide/Thiazide-like diuretics should be used only if needed for BP control and in patients with fluid retention. They should be used with caution, because they may (i) increase serum calcium concentration in patients with bone metastasis, (ii) further increase the risk of cardiac arrhythmias induced by some anticancer drugs because of prolongation of the QT interval by inducing hypokalemia, (iii) worsen SIADH-dependent hyponatremic states occurring in some cancer patients and (iv) worsen hypovolemic states or dehydration [205,1579,1595].
  • A preferential use of a DHP-CCB in combination with a RAS blocker may, therefore, apply in many cancer patients (Fig. 12).
  • 2. Use of non-DHP-CCBs, which are indicated for heart rate control in patients with contraindications or intolerance to BBs, is problematic and should be avoided in some cancer patients. This results from the evidence that non-DHP-CCBs (i) are moderate inhibitors of CYP3A4 or P-gp and interfere with the pharmacokinetics of some, but not all, anticancer drugs, e.g. oral RTKIs that are substrates of CYP3A4 or P-gp [1595,1596] and (ii) can worsen HF induced by cardiotoxic anticancer drugs due to their negative inotropic effects (see Section 11.2.2). Nevertheless, this does not exclude their cautious use in cancer patients “per se”, e.g. in hypertensive patients with tachycardia who cannot tolerate BBs and in patients treated with anticancer drugs that do not show relevant pharmacokinetic interactions via P-gp or CYP3A4.
20.8.3.3 Treatment of hypertension induced by VEGF inhibitors

There are no data from RCTs that can help the prevention and/or treatment of de-novo hypertension induced by anticancer therapies, e.g. in response to VEGF inhibitor treatment. Nevertheless, sodium restriction may be helpful as a maximum intake of 4 g sodium per day in combination with dietary counseling has recently been shown to attenuate the VEGF inhibitor-induced BP rise by almost 50% in a small (pilot) study (16 patients) [1597]. In a retrospective cohort study involving 343 cancer patients that were treated with oral VEGF inhibitors (sorafenib, sunitinib, pazopanib, regorafenib, lenvatinib or cabozantinib), about half of the included patients exhibited significant BP increases (20 mmHg in SBP or 10 mmHg in DBP). Normotension at baseline and treatment with pazopanib identified as significant risk factors for this significant BP rise [1598]. Treatment with a CCB or RAS blocker (ACEi or ARB) effectively reduced SBP (24.1 and 18.2 mmHg, respectively) and DBP (12.0 and 11.0 mmHg, respectively) [1598]. In patients treated with VEGF inhibitors any BP-lowering therapy administered during the on-treatment periods must be carefully monitored, e.g. by HBPM, and reduced or even stopped as needed during the off-treatment periods because of the excess risk of hypotension.

Management of hypertension in cancer patients

figure67

20.8.4 Follow-up and management of hypertension in cancer survivors

Long-term close FU of cancer patients after active cancer treatment is important. One reason is that, acute BP changes may occur in patients who were exposed to anticancer drugs that induce short-term and reversible BP increases, requiring backtitration or eventually discontinuation of any previous BP-lowering drugs to avoid hypotension. Another reason is that cancer patients with preexisting hypertension may have developed progression of previously existing HMOD or de-novo organ damages because of the cardiorenal toxicity of anticancer drugs. Thus, depending on the individual risk profile and phenotypes, FU should include monitoring of HMOD parameters (Fig. 6). Overall, BP monitoring during long-term FU, preferably by HBPM, is recommended, because long-term cancer survivors are at increased risk to develop hypertension, e.g. long-term survivors of cancer in childhood [1599], and CV events. Finally, patients who developed severe worsening of hypertension during anticancer therapy may have a secondary cause of hypertension as underlying disease and appropriate diagnosis to exclude or detect such forms may be considered in suspected cases (see Section 6).

20.9 COVID-19 and hypertension

The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), causing the coronavirus disease 2019 (COVID-19) pandemic, led to fundamental disruptions of many if not all aspects of people's lives. In the early acute setting [1600], but also during the first year [1601] of the pandemic, COVID-19 was associated with a substantial increase in mortality including fatal CV events such as stroke, ischemic heart disease, arrhythmias, HF, venous thromboembolism, cardiac arrest and acute kidney injury and failure. All these complications have a close association with hypertension. However, several more direct aspects of the relationship between COVID-19 and hypertension have emerged. Coexistence of hypertension and COVID-19 was not only one of the most common comorbidities in patients with COVID-19 [1602], but its coexistence with this infection was also associated with an increased risk of severe COVID-19 (defined as hospitalization or death) [1602,1603]. Hypertension was also reported to be an independent factor in a vulnerability score predicting severe or fatal COVID-19 developed from millions of people from Italy [1604]. In treated hypertensive patients, elevated SBP showed a dose– response relationship with severe COVID-19 [1603], even when age and CV comorbidities were taken into account [1605]. These are important observations, because the possible association between hypertension and COVID-19 in the initial phase was confounded by the high age of patients hospitalized with severe COVID-19 in parallel with the high prevalence of hypertension at that age [1605]. The most likely explanation supporting hypertension as a risk factor for the COVID-19 severity is that hypertension often causes subclinical HMOD in vital organs [217,276], which may reduce the body's defenses against severe infections. Another possibly additive explanation could be the detrimental role of several immunological dysregulations, which may be associated with hypertension [1606].

20.9.1 COVID-19 and RAS inhibitors

The findings that (i) SARS-CoV-2 uses the angiotensin-converting enzyme 2 (ACE2) receptor for cell entry, and (ii) in experimental settings, treatment with ACEis and ARBs may lead to an upregulation of ACE2 expression [1607], generated the hypothesis that treatment with ACEis or ARBs could increase the infectivity and severity of the clinical course of COVID-19. This caused concern in health authorities and scientific societies, which feared that, neglecting the observation that ACE2 was found to have protective effects against lung injury in animal models [1607], patients might stop using these drugs in diseases for which they have a life-saving role, with a consequent increase of fatal or nonfatal CV events. However, large observational studies already made available in the early pandemic phase [1608,1609] consistently showed that treatment with RAS blockers did not affect the risk of COVID-19 infection, severe illness or mortality [1610], a conclusion confirmed by later meta-analyses of available studies [1611]. Minimal effects of RAS blockers in either direction on the course of COVID-19 were later shown by another much larger meta-analysis, which, however, additionally emphasized the limitations and bias of conclusions based on observational studies [1612]. This confirmed ‘a posteriori’ the recommendation issued in the early pandemic phase by several learned societies, including the ESH [1607], that patients treated with RAS inhibitors for hypertension, HF, CAD or other conditions should in general not discontinue this treatment. Unfortunately, as shown by a survey in ESH Excellence Centers, a sizeable number of patients did discontinue treatment, with possible, although never calculated, detrimental effects [1613].

20.9.2 COVID-19 lockdown and hypertension management

In an attempt to limit the infectious spread, diagnostic and therapeutic procedures related to hypertension were dramatically reduced during the first year of the COVID-19 pandemic, with the largest reduction during the first lockdown [1614]. This was also true for patients with severe hypertension [1615] and involved to a striking degree face-to-face consultations and clinic BP measurements [1616], followed by a decrease in medication adherence [1617]. In response, remote clinical management programmes, involving not only physicians, but also nurses and pharmacists, were developed [1618]. Such programmes, based on automatically transmitted standardized HBPM and evidence-based hypertension management algorithms, achieved some implementation during the pandemic [1618]. In fact, expanding the role of virtual management for hypertension may be regarded as an important favorable consequence of the pandemic in the field of hypertension (see Section 21). With respect to practical antihypertensive treatment aspects, no consistent alterations in the choice of drugs and treatment strategies (including BP thresholds and targets for treatment) have been adopted in patients with COVID-19 because of lack of available evidence and appearing reasons to do so.

On top of the disruption in medical care for chronic diseases such as hypertension [1614], the COVID-19 pandemic caused potential changes in lifestyle factors and behaviors (physical activity, dietary patterns, alcohol consumption, smoking, emotional/psychologic stress, changes in sleep patterns and diurnal rhythms), as well as environmental changes (air pollution, environmental noise) likely to influence BP control and CV risk during and after the pandemic [17], in particular during its strict or less strict lockdown phases. A few changes such as reduced air pollution and noise exposure (due to less traffic) were potentially favorable, because air pollution and noise have been shown to have a pressor effect. However, most were probably detrimental. Indeed, a rise in office [1616,1619] and home [1620] BP, as well as an increase in the proportion of patients with uncontrolled hypertension [1620] has been observed during the pandemic, albeit not consistently in all studies [1621].

20.9.3 Vaccination against SARS-CoV-2 and hypertension

Vaccination against SARS-CoV-2 has clearly reduced the occurrence of myocardial infarction and stroke after COVID-19 [1622,1623]. These data should not be interpreted as to mean that the vaccine has a specific CV protective effect. The reduced incidence of CV events is rather attributable to the fact that the SARS-CoV-2 infection affects multiple organs and can by all accounts be considered a systemic rather than lung-restricted disease. No signal of a BP increase or consistent change has emerged from the RCTs evaluating the efficacy and safety of the vaccines [1624]. This seems to be in line with the results of pharmacovigilance databases and observational studies that the pooled estimated proportion of abnormal/increased BP after vaccination was only 3.2% [1625], but in the absence of randomized data, no reliable conclusion is possible. Short-term BP changes such as stressor-related BP responses and white-coat effects may have played some role.

20.9.4 Long COVID-19 and hypertension

Regarding the persistent CV symptoms several months after COVID-19, i.e. Long-COVID, the CV system is frequently affected [1605] and preexisting hypertension may be a modest risk factor [1626,1627]. Hypertension may be among the more common reasons for medical consultations after COVID-19. However, long-term FU studies, and more data, in general, are needed to shed more light on this important issue [1628,1629].

21. FOLLOW-UP

21.1 Importance of follow-up

The FU of hypertensive patients is crucial not only to ascertain achievement of BP control but also to support lifestyle modifications, assess drug adherence and medication side effects and to adapt therapy and check for development or changes in HMOD (Fig. 21). In this context, three main questions need to be addressed: (i) how often patients should be seen, (ii) what needs to be checked and (iii) who and which setting should be involved in the management of patients during FU, in addition or alternatively to office re-visits. Unfortunately, for each of these questions, available studies are scarce. Nevertheless, these guidelines acknowledge their importance for patients’ FU and address them in some detail below. The initiation of drug therapy may be evaluated monthly until BP control is achieved [97,1352]. However, three large observational studies and one small RCT suggest that FU visits with shorter intervals (every 2 weeks) result in earlier and more common BP control rates [1630–1633]. It seems obvious that during the treatment titration phase, the frequency of the visits should not be subjected to rigid rules but differ according to BP phenotypes and response to treatment. Other obvious parameters that affect the frequency of FU visits are severity of hypertension, presence and type of HMOD, CVD or CKD as well as other comorbidities. Return visits have been reported to be more than twice more common in patients with ≥ 3 comorbidities compared with patients with fewer or no comorbidities [1634]. A BP reduction is slower with monotherapy than with dual-drug combinations, with which a reduction in BP levels is expected within 1– 2 weeks, although a further smaller progressive BP decline may continue for few more weeks.

F21
FIGURE 21:
Follow-up of patients with hypertension.aCan be adapted according to the clinical status.

Once the BP target is reached, a visit interval of a few months seems reasonable. No difference was detected in BP control between FU visits at 3 and 6 months intervals in one study [1635], thus favoring a twice or four times a year visit frequency during the first year after treatment initiation for the majority of hypertensive patients. A 3-month time interval is in line with the results of a large observational study in 90 000 hypertensive patients that revealed that when the time to return visit exceeded 2.7 months, the risk of CV events was increased by 18% [646]. After the first year, two or even one FU visit per year seems a reasonable frequency, if medical conditions are stable, BP control appears to be consistent (based on information from HBPM whenever possible) and patients are not affected by major treatment-related problems. Patients should be made aware that office and out-of-office BP values tend to be a few mmHg lower during the summer than during the winter months [1636,1637], due to differences in indoor and outdoor temperature as well as to other factors [1636,1638–1640]. Usually, lower BP values in the summer do not pose specific problems to the patient but may require a supplementary medical visit. If lower BP values are associated with dizziness or fatigue, patients may be advised to modestly downtitrate drug treatment, especially diuretics.

Another useful bit of information is that BP increases under conditions of hypobaric hypoxia due to a number of factors, including sympathetic activation [1641]. Patients with grade 2 and 3 hypertension and poor BP control by treatment should avoid exposure to altitudes above 2500 m from sea level. Vacationing at lower altitudes does not appear to pose BP problems to hypertensive patients, although BP checks before and during exposure to altitude appears advisable. Possible impairment of blood oxygenation and other problems may extend the risk of hypobaric hypoxia to lower altitude levels in older patients or patients with respiratory or CV diseases [1641].

Obviously, during the FU visits, BP measurements according to the recommendations of the guidelines are of fundamental importance. BP measurements should be extended to ABPM whenever indicated and feasible, at intervals usually longer than those for office BP measurements. HBPM can be advised not only before visits but also on a more regular basis as a useful tool to verify consistency of BP control as well as to improve adherence to treatment. During visits, there should be collection of the most recent medical history, including questions addressing lifestyle interventions, tolerability of drug therapy and any changes of comorbidities and comedications, including intake of over-the-counter medicines. A new physical examination should usually be dispensed with, if the medical course before the re-visit was inconspicuous. Standard laboratory investigations, electrolytes (serum potassium) and kidney function are among the parameters that should be monitored most frequently together with serum cholesterol, the lipid profile, blood glucose and other parameters based on the individual drug treatment strategies and comorbidities. Re-assessment of HMOD should be carried out according to the considerations made in Section 5.5. HMOD assessment is important, as progression or regression of HMOD at FU visits has a major influence on management strategies and scheduling of further FU visits.

During long-term FU, visits may be also carried out by nonphysician healthcare professionals, such as qualified nurses or pharmacists. This approach has been adopted in some European and other countries, depending on the local organization of health resources and is supported by the results of RCTs, observational studies and meta-analyses, which point towards similar BP reductions, when treatment is handled by primary care physicians, nurses or pharmacists [1642–1645]. An important contribution in this direction may be provided by tele-health technologies in combination with patient empowerment. Further development of this approach can be expected to make an important contribution to FU management in hypertension in the future [139].

21.2 Adherence

21.2.1 Definitions

Adherence is defined as the extent to which a person's behavior, such as taking a medication, following a diet or executing lifestyle changes, corresponds with agreed prescriptions or oral recommendations from a healthcare provider. Persistence defines the length of time between the initiation of treatment and the last dose immediately preceding its discontinuation [479,1646]. To this end, adherence reflects a punctuated assessment of a patient's alignment with the prescribed treatment, whilst persistence is an illustration of adherence to the prescribed therapy over a longer period. Because of negative connotations, compliance is not the most appropriate term to use when referring to patients following of therapeutic recommendations from healthcare providers [1646].

21.2.2 Prevalence of nonadherence and associated burden

The estimates of the prevalence of nonadherence to antihypertensive treatment vary across studies depending on (i) the method used to assess adherence, (ii) the country/healthcare system where the research was undertaken and (iii) the clinical characteristics of hypertensive patients recruited. In general, nonadherence rates are higher in low-income and middle-income countries than in westernized societies [616] and in patients with difficult-to-manage hypertension (e.g. suboptimal BP control) compared with those from the general hypertensive population [657,669]. In a large sample of patients from the general population who were treated with antihypertensive drugs, only approximately one-fourth were adherent (a prescription coverage for ≥ 75% of the FU). On average, approximately one in three to four hypertensive patients are nonadherent to antihypertensive treatment, based on direct biochemical analysis of bodily fluids [612,657,658]. Nonpersistence is one of the most common cause of poor adherence in hypertension particularly among newly treated patients [479,1647]. In newly treated patients from Italy, about 36% did not renew the initial prescription of antihypertensive medications a second time [1648]. There is a strong correlation between the extent of nonadherence to antihypertensive treatment and the magnitude of BP elevation, based on both office BP measurements and 24 h ABPM [657]. Nonadherence is associated with several adverse CV outcomes, including LVH [1649], microalbuminuria [1650], myocardial infarction, stroke, HF [1651,1652], CKD [1653,1654], hospitalization rates, all-cause mortality [1655], reduced quality of life and overall increased healthcare costs [479]. CV hospitalization and death are also closely related to treatment discontinuation [478] and the relationship between adherence and clinical outcomes is apparent in both men and women [1008], younger and older patients as well as patients with different levels of comorbidities and mortality risk[479]. For this reason, the 2018 ESC/ ESH guidelines put a particularly strong emphasis on the detection and management of nonadherence to BP-lowering therapy [4], and this is mirrored in both the AHA/ACC guidelines [488] and the ISH guidelines [32]. Emphasis on adherence is recognized as one of the areas of convergence between European and American guidelines [1656].

21.2.3 Methods to detect nonadherence to antihypertensive treatment

The ideal method of detection of nonadherence should be to prove the ingestion and provide information on how frequently this occurs over time because adherence is a dynamic process, i.e. it can vary considerably from one period to another. Today, none of the available methods fulfils these two criteria. Relying on the physician's impressions derived from interviews is convenient but is associated with a significantly biased assessment leading to an overestimation of adherence. Adherence questionnaires do not fulfil all the validity and reliability test criteria [1657] and show generally very poor/no relationship with BP control and CV outcomes [1658]. Directly observed administration of BP-lowering medications (usually followed by BP monitoring) can be informative but is not practical in most healthcare settings (expensive, operationally challenging) and cannot take into account the variable nature of adherence over time. It can be also clinically hazardous (severe BP drops have been reported in patients following witnessed intake of medicines) [1659]. Checking prescription refill records is inexpensive and may provide information on patients’ persistence with their antihypertensive treatment, particularly in the context of healthcare systems in which all or almost all prescriptions are registered and centrally stored [477]. However, collecting a prescription is not equivalent to taking a medication, and thus this approach overestimates adherence. Electronic drug monitoring using sensors that register an act of opening a medication-dispensing container/ blister pack is generally very accurate and provides detailed information upon both timing and frequency of adherence as well as persistence on treatment [675]. However, the costs, the risk of intentional/unintentional dose removal from the container without ingestion and the unfeasibility of the method for numerous antihypertensive medications represent important limitations. Biochemical detection of antihypertensive medications and/or their metabolites in bodily fluids provides direct and objective confirmation of drug intake. Yet, this method is costly, does not account for day-to-day variability in adherence and is not immune to toothbrush adherence effect (increased adherence before visits) [479,1660]. Self-reported adherence is usually overestimated and should not be routinely used in clinical practice. However, an affirmative confirmation of nonadherence is generally informative. Digitally tagged pills are not currently available in clinical practice and have a very high cost.

21.2.4 Etiology of nonadherence to antihypertensive treatment

WHO multidimensional adherence model classifies barriers to nonadherence into several categories, including: (i) health system (communication with a healthcare provider, low satisfaction with pharmacy services or problems with drug reimbursements); (ii) therapy-related (side-effects, complexity of drug regime or interference with daily routine); (iii) disease-related (severity of symptoms, lack of symptoms or presence of comorbidities); (iv) patient-related (low self-efficacy or inaccurate beliefs about medications) and (v) socioeconomic status-related (poverty, lack of family support or unemployment) [1661]. The causes of nonadherence to antihypertensive treatment largely maps onto these above dimensions. Unintentional is more common than intentional nonadherence [479] and in some countries, access to healthcare and cost of medications are becoming an increasingly important contributor to nonadherence in hypertensive patients [139,616]. The generally asymptomatic nature of hypertension is likely to augment the risk of nonadherence when compared with CV conditions known to produce symptoms when treatments are not followed. For example, chronic HF is associated with much lower rates of nonadherence than hypertension [1662], and better adherence to antihypertensive treatment was reported in patients with a history of hospitalization because of CV or renal events [619]. No difference in adherence has been reported between brand name drugs and generics [1663], while the number of prescribed antihypertensive medications has been shown to act as a key determinant of nonadherence to BP-lowering treatment [669]. In contrast, class of antihypertensive treatment is not a consistent determinant of nonadherence to antihypertensive treatment, although persistence is higher with some antihypertensive drugs than with others [669]. Algorithms that use demographic and simple clinical data are not yet sensitive or specific enough to predict nonadherence to antihypertensive treatment, as drug adherence is a very dynamic and poorly predictable process. Overall, nonadherence to antihypertensive treatments has a multifactorial, complex background and there is a large variation in the causes of nonadherence to BP-lowering treatment among patients.

21.2.5 When and how to screen for nonadherence

Screening for nonadherence to antihypertensive treatment should be part of the routine assessment of effectiveness of BP-lowering drugs [4] and should be checked (i) at every clinical appointment, (ii) prior to escalation of antihypertensive treatment [32], (iii) prior to screening for secondary hypertension [32] and (iv) when true resistant hypertension is suspected [32,1664]. Screening for nonadherence to antihypertensive treatment should also be considered in patients who are taking two antihypertensive medications and have an inadequate BP response to this treatment [4,659]. Objective indirect or direct methods (review of pharmacy records, electronic monitoring devices, directly witnessed intake of medication and biochemical detection of medicine in urine) are generally preferred over subjective methods of diagnosing nonadherence to antihypertensive treatment [32]. In lower resource settings, where these tests are not available, confirmation of nonadherence (but not of ‘adherence’) by a patient can be informative. Lack of an expected response to antihypertensive treatment, such as no BP improvement with multiple antihypertensive drugs, elevated heart rate, despite treatment with BBs or non-DHP CCBs, increase the probability of nonadherence, in particular when combined with information on the most established risk factors of nonadherence such as polypharmacy or the occurrence of side-effects.

Screening for nonadherence to BP-lowering therapy improves the diagnostic approach and treatment decisions almost in all hypertensive patients. Indeed, in patients with suboptimal BP control (e.g. apparent resistant hypertension), a confirmation of nonadherence to antihypertensive treatment (i) substantiates a diagnosis of pseudoresistant hypertension, (ii) explains the cause of resistance to treatment, (iii) helps to avoid unnecessary treatment escalations and expensive additional investigations and (iv) reduces the healthcare costs [1665]. A confirmation of adherence to antihypertensive treatment in such patients justifies undertaking additional diagnostic tests (e.g. to exclude secondary hypertension) and suggests that a change in antihypertensive treatment may be necessary. In patients with satisfactory BP control, a confirmation of nonadherence to antihypertensive treatment allows a de-escalation of antihypertensive treatment while results consistent with adherence to therapy provide a positive reinforcement of the effectiveness of this treatment [479].

21.2.6 Management of nonadherence to antihypertensive treatment

Management of nonadherence to antihypertensive treatment should be tailored to the individual modifiable drivers of nonadherence in each patient. There is not a single universal strategy that could help to manage nonadherence in all hypertensive patients. To this end, a confirmation of nonadherence to antihypertensive treatment should be followed by a nonjudgmental discussion between a patient and a healthcare professional (as partners) [669,1660,1665,1666]. This should lead to understanding the barriers to adherence to antihypertensive treatment. The therapeutic strategies to consider, in combination or alone, are: (i) simplification of treatment (e.g. reduction in the number of antihypertensive medications or pills/tablets with a preference for long-acting and once-daily administration systems) [616,673], (ii) single-pill combinations [617,647,673], (iii) reminders and electronic monitoring [616], (iv) elimination of medications that cause side-effects or are unnecessary [479], (v) financial incentives [616] and (vi) addressing incorrect beliefs about hypertension and BP-lowering therapy [616], family and social support [616], repeated adherence testing [616,658] and matching therapy with daily routines [479]. Collaboration with other healthcare providers in the context of team-based care[1667,1668] and the development of virtual management of hypertension based on HBPM [62,1669].

Nonadherence to antihypertensive therapy

figure69

21.3 Clinical inertia

Clinical inertia in hypertension management can be divided into (i) diagnostic inertia, i.e. failure of the physician or another health-care provider to detect a BP elevation in the patient and (ii) therapeutic inertia, i.e. failure of the physician to start or modify drug treatment (drug change, dose change or drug addition) in the presence of BP values above the recommended target [1670]. There is little information on the contribution of diagnostic inertia to the fraction of the hypertensive population unaware of its high BP status. On the other hand, evidence is available that therapeutic inertia significantly contributes to the low rate of BP control that characterizes treated hypertension in real life [532,1671,1672]. In a Dutch study on 530 564 individuals followed by general practitioners, 10% (n 64 000) of the hypertensive population was found to have uncontrolled BP because of therapeutic inertia [1673]. Greater percentages have been reported in other studies, in which uncontrolled BP in the absence of appropriate drug titration has been directly determined or indirectly inferred from questionnaires [1674–1678]. The factors associated with therapeutic inertia have also been investigated and found to include the doctor's feeling that the measured BP does not reflect the patient's true BP, deferment of the decision to future visits, hope for an effect of lifestyle changes or fear of adverse effects [1679]. Therapeutic inertia has also been consistently associated with a limited SBP increase, older age and BP values moderately above target [1676,1678]. Finally, a very common type of inertia involves failure of the physician to use combinations of antihypertensive drugs in most patients. In a large study on the populations of Belgium and Luxemburg, about 50% of the patients were found to be on monotherapy after 8 years of treatment [1675]. Similarly, in a population of northern Italy, only 36% of patients starting antihypertensive drug monotherapy moved to drug combination after 3 years of treatment [608]. In the same population, long-term antihypertensive monotherapy has been recently found to extend to about two-thirds of the patients, largely independently of the demographic and clinical conditions and with few changes in three cohorts spaced over a 6-year period [614]. Prevalence of antihypertensive monotherapy is shared by several European countries [612]. This is an example of large-scale therapeutic inertia because BP control requires more than one antihypertensive drugs in the vast majority of patients [1680,1681]. It is important to know that therapeutic inertia is also reported in the setting of clinical trials in which, despite specific treatment protocols and regular verification of the planned BP targets, a number of patients do not complete the planned treatment titration and thus fail to reach BP control [609]. Importantly, the adverse effects of therapeutic inertia usually extend beyond an uncontrolled BP and include lack of control of associated risk factors.

Although only few studies have addressed the question of how to limit doctor's inertia, a few suggestions can be made. The most important step is obviously to improve doctors’ knowledge of the increased CV risk associated with hypertension, whatever the age, gender and protective effects of adequate BP reductions, a goal that needs dissemination of treatment guidelines. Patients’ knowledge and empowerment as well as home BP measurements are also important to provide appropriate and timely feed-backs to the doctor, with perhaps a further help of telemetric transmission of the home BP measurements [1682]. A CV prevention strategy combining three approaches has been shown to improve BP control and to reduce therapeutic inertia in a large group of hypertensive patients in the USA [1683]. More frequent visits and simpler uptitration treatment strategies can also be considered. Incentives based on the number of patients achieving BP control has also been proposed [1684]. It should be noted that in some cases, therapeutic inertia is caused by patients’ reluctance to increase the number and/or doses of the drugs or to the doctor's need to depart from the average guidelines recommendations in a number of individual patients. In this case, therapeutic inertia is only apparent [1685].

There are potential solutions to overcome clinical inertia in the management of hypertension. They can be classified as not only educational (i.e. improve knowledge), clinical (i.e. improve screening and detection of uncontrolled BP) and therapeutical but also organizational (improved FU). Most of the time they are directed to the physician but improving a patient's attitude can also contribute to fighting clinical inertia [479].

21.4 Patient empowerment

In order to follow a health-care plan and reach therapeutic targets, involving the patient is strongly recommended both initially and at every step of FU [931]. Factors such as poor health literacy or lack of education can directly affect the quality of care, thus patient information and sharing medical decisions are the first steps of a patient-centered approach. This empowerment process (i.e. giving the power to the patient to be active in managing his/her medical condition) is a key factor for success. These approaches are based on behavioral and motivational strategies. They offer a good chance to improve management of high BP by enhancing adherence to drugs or lifestyle modifications but also more generally to achieve a healthy lifestyle, as these approaches have been validated for the management of other CV risk factors (smoking cessation, weight loss, moderation in alcohol intake, increased physical activity and consumption of a healthy diet). Although high-quality evidence supports empowerment and cognitive interventions in reducing BP [1683,1686], the level of evidence and impact of shared medical decision on BP control is low [1687].

Discrepancy between physician and patient expectations can lead to behavioral changes affecting BP control. In this situation, interventions such as goal setting, provision of feedback, self-monitoring, FU, motivational interviewing, and promotion of self-sufficiency are effective, and effectiveness is better when the interventions are combined rather than promoted individually (lower level of evidence) [137,1688]. The use of telemedicine and mobile health technologies (see Section 21.6) when the patient is at home can be of help for not only promoting self-monitoring and self-sufficiency but also for improving health and well being. The physician– patient interaction needs to be framed within a team-based care approach involving the healthcare system, professionals and a multidisciplinary group of health providers (see Section 21.8), leading to a comprehensive patient-centered plan of care. These approaches need to be personalized and cultural, social and economic context variables need to be considered.

21.5 Follow-up of low-risk hypertensive patients and deprescribing

The therapeutic management of low-risk hypertensive patients is perceived to differ from intermediate/high-risk patients, not only because monotherapy can be more frequently used as first-line (although the present guidelines recommend initial two-drug SPCs to also be used in most of these patients) but also because it may seem rational to assume that the FU of low-risk patients should be less close, as BP levels are lower and no significant comorbidities may exist. However, even these patients might benefit from closer FU visits, and more so in the initial period of therapy, because they are at high risk of nonpersistence on treatment. In a retrospective study of almost 3000 patients younger than 40 years of age and with incident hypertension, it was found that during a 2-year FU period, BP control rates were higher with shorter visit intervals [1689]. In some low-risk hypertensive patients in whom treatment is accompanied by effective BP control for an extended period, it may be possible to reduce the number and/or dosage of drugs. This may particularly be the case if BP control is accompanied by healthy lifestyle changes such as weight loss, exercise habit and a low-fat and low-salt diet, which minimizes environmental pressor influences. A reduction of medications should be made gradually, and the patient's BP should be checked frequently because reappearance of hypertension is frequent and can occur at variable time intervals, i.e. not only within days or weeks but also possibly after many months, because of the slow re-development of structural arteriolar changes that favor a BP elevation [676]. Patients with a high CV risk, HMOD or previous accelerated hypertension should not have their treatment withdrawn. Patients with high-normal BP or WCH frequently have additional risk factors and may also exhibit asymptomatic HMOD with a prevalence that, for WCH, is intermediate between normotensive and sustained hypertensive individuals [1690–1694]. Thus, even when untreated, they should be scheduled for regular (at least annual) FU visits to measure office BP, as well as to check the CV risk profile. At annual visits, recommendations on lifestyle changes, which represent the most frequent recommended treatment in many of these patients, should be reinforced. In WCH patients, annual measurements should include not only office BP but also out-of office BP and assessment of HMOD, because in these patients, there is evidence of a greater risk of development of new HMOD and sustained hypertension [217].

21.6 Use of telemedicine and tele-health technologies

The advent of new technologies has allowed use of internet-based interactive digital interventions (tele-health) and health-related mobile applications that can also be installed on smartphones. This enables, at least in perspective, virtual care of hypertension [139]. The COVID-19 pandemic highlighted the role of remote management of chronic conditions and greatly contributed to the familiarization of both patients and physicians with these new technologies [17,139].

Interactive digital interventions include behavioral aids and promotional material for hypertension self-management. Several studies and meta-analyses suggest that these interventions are associated with better patient education, greater BP reduction and even reduced CV outcomes [1695–1697]. Mobile applications include the assessment of heart rate, thereby recognizing AF, sleep quality, physical activity and even cuff-less BP measurement [62,1698]. Transmission of HBPM, or even additional data obtained by physician are under investigation by many studies, including those promoted by ESH [1699,1700]. At present, it is premature to reach a conclusion on the benefits of these technologies, and the virtual management approach in general. Nevertheless, favorable data on home BP telemonitoring have been obtained. A meta-analysis of 46 RCTs in about 14 000 hypertensive patients revealed that home BP telemonitoring is associated with significant BP reduction and improved BP control [1701]. Similar results were reported by others [139,1702] including a study during the COVID-19 pandemic [1621].

21.7 Challenges of long-term follow-up

Hypertension requires lifelong therapy. Therefore, long-term FU is needed, and a proper FU organization is essential. Strategies for FU organization depend on the specialty of the treating physicians (e.g. primary or specialist care i.e. cardiologists, internists, nephrologists, etc.), the setting of the care (private office, hypertension clinic, hospital unit), and the resources available. In practice, the vast majority of hypertensive patients are taken care of by primary or family physicians and only a small percentage is seen and followed by specialists and, even more rarely, by hypertension centers. Hypertensive patients are also not candidates for hospitalization except for hypertension emergencies or when a hypertension-related clinical complication occurs. Ideally, complete electronic health records should be available for each patient and all information regarding the initial evaluation and lifelong management should be included, i.e. the records should incorporate demographic information, medical history, lifestyle habits, clinical findings, comorbidities, HMOD, concomitant medications, side effects of medications, laboratory results and hospital records. In several European countries, very little of this integrated information is made available for a number of reasons, including strict privacy rules that do not consent the sharing of clinical data. Nevertheless, physicians involved in long-term hypertension FU should build patients’ records that include crucial information, such as the trajectories of office and, if available, the out-of-office BP profile, the history of the treatment strategies and of their inconveniences, the CV risk factor profile, the diagnoses at discharge from hospital and the HMOD dynamic status.

21.8 Role of general physician, pharmacies and team-based care

Physicians in primary care, i.e. general and family physicians, play a pivotal role in hypertension management. However, primary care workload has increased markedly over the last decades and appears to be close to or to have reached the saturation point in many countries. A key component of this workload is the diagnosis and management of long-term chronic conditions, among which hypertension-related consultations play an important role. It is likely that the hypertension-related fraction of this workload will further increase because the prevalence of hypertension is increasing, and lower BP targets make antihypertensive treatment more complex and medical visits more frequently needed. In order to avoid this problem, it is of fundamental importance that medical visit time should not be shortened. Visit times are already too short and frequently even below the time needed to properly measure BP. Attention should rather be directed to healthcare plans that carefully quantify the ‘time needed to treat’ by primary or specialist care physicians, including the one necessary for management of hypertension [1703], and to provide the necessary personnel accordingly. Alternative models of hypertension care, in which other health professionals participate, may also be tested and implemented. In this context, favorable results have been obtained by adopting care models that are different from the current one. A substantial increase in the rate of BP control as assessed over a 3-year FU has been reported by a Chinese study in which the care model was based on cost elimination, home BP measurements and home BP visits [1704]. An improvement of BP control and CV risk has been reported by use of a healthcare model, which involved primary care physicians but was led by nonphysician health workers [1705]. The evaluation of hypertension care by pharmacists has shown that BP control can be improved [1667,1706] and that community pharmacies may offer proximity and accessibility to the majority of the population in many countries, including individuals who would otherwise not seek medical advice. Community pharmacies may also be suitable places to provide health education on hypertension management [1707], which is essential for long-term treatment of largely asymptomatic diseases. In Europe, there are about 160 000 community pharmacies, with an estimated 46 million citizens visiting a pharmacy every day [1708], and available evidence supports a favorable role for them in the detection and control of hypertension and CV risk factors [1709–1716]. Finally, hypertensive patients may be followed by healthcare teams that provide the expertise and cooperation of different medical specialists and other healthcare professionals. This may offer substantial benefits, provided that each team member has a clear and specific role. Team-based interventions can significantly reduce physicians’ workload and have been associated with significantly greater BP reductions and enhanced BP control rates compared with usual care [1668,1713]. Participation of nurses in healthcare teams with a role that goes far beyond BP measurements and includes instrumental examinations, nonpharmacological and pharmacological treatment delivery and explanations of the nature of the clinical condition and the rationale of the medical decisions (already successfully implemented in several European countries) may be of special importance.

21.9 Hypertension clinics

Outpatient Hypertension Units are available in most large hospitals and can offer a high-quality care to a considerable number of hypertensive patients, including patients with complicated hypertension, resistant hypertension or suspected secondary hypertension. The importance of these centers can be illustrated by the Hypertension Excellence Centers that have been established by ESH throughout Europe, fulfilling prespecified criteria regarding the ability to provide multidisciplinary services and offer high-quality inpatient and outpatient care [1717]. These centers are also contributing to clinical hypertension research, and thereby contributing to the advances of knowledge in the field [1614,1718].

21.10 Health risks at workplace

Industrialization and globalization have highlighted the role of occupational medicine reviving the previous concept of ‘blue collar’ and ‘white collar’ hypertension. Several occupational factors have been associated with hypertension and CVD and should be given attention by physicians responsible for hypertensive patients during FU. Long working hours have been associated with unhealthy lifestyles, obesity and physical inactivity [410,1719], i.e. factors that are implicated in BP elevation and CV events. Recent meta-analyses revealed that shift work is associated with increased risk of incident hypertension [1720], which is more evident in night workers compared with rotational shift workers [1721]. Occupational physical inactivity has also been associated with increased risk of incident hypertension [1719], and this is the case for sedentary occupations as well [1722]. Job-related stress is another important contributing factor to a BP elevation. In a recent report including 63 800 employees from the Dutch LifeLines Cohort Study, higher levels of job strain were associated with higher BP and increased risk of incident hypertension [1723]. The aforementioned occupational risk factors are usually interconnected and may act synergistically. Indeed, a recent epidemiological study revealed that the risk for hypertension was greater in workers with high job strain and physical inactivity[1724]. Recently, several meta-analyses have unveiled an increased risk of hypertension with high occupational noise exposure in line with a similar effect of environmental noise. The increased risk varies from marginal (8%) [1725] to substantial (155%) [16], highlighting the complexity of the association and the necessity of standardization in future studies. Both job-related stress and shift work have been associated with increased CV risk [1726]. The recent COVID-19 pandemic uncovered another occupational hazard, i.e. the unemployment fear that may also adversely impact on BP and CV events. The WHO emphasizes the workplace as a priority setting for promotion of health and wellbeing, including provision of a safe and healthy physical and psychosocial work environment. Evidence concerning health promotion in the workplace suggests that health promotion programs are effective when interventions address both individual and environmental influences [1727].

21.11 Patient organizations

Patient organizations are becoming increasingly important for chronic diseases. They are nonprofit organizations formed by patients or those who care for them. Although they were initially created mainly to provide patients with support and advice (e.g. Alcoholics Anonymous), their role is constantly expanding. An advocacy role aiming not only to reach public awareness but also to exert political pressure has been attained with large campaigns in mass and social media as well as with the inclusion of organization representatives on official advisory and decision-making boards. In some cases, patient organizations are even actively involved in research and clinical trials and operate to guide future research towards their needs. The structure and the size of patient organizations vary significantly between diseases and geographic regions. Large patient organizations for hypertension exist in a few European countries (e.g. Germany and France) but unlike for other diseases (HF, CKD or diabetes mellitus) they are absent or only nominally active in other countries. This has negative implications, because patient organizations may help patients to better share their experience with others and better cope with the multiple problems posed by a chronic disease. An example is the involvement of patients in programs that evaluate the role of RDN as a therapeutic intervention in hypertension [769]. A future role of patient organizations may also be their participation in the elaboration of guidelines to which they would offer a wider perspective.

Hypertension management during follow-up

figure70
figure71

22. GAPS IN EVIDENCE AND FUTURE OPPORTUNITIES

As mentioned in Section 1, the present guidelines have placed RCTs at the top of the evidence on which to base the recommendations on hypertension management. This acknowledges a fundamental advantage of this research approach, i.e. the identical composition or at least close similarity of the compared groups at baseline and thus the possibility to safely ascribe their FU differences to the intervention studied. However, RCTs have limitations, and their adoption over almost 60 years has not been ableto provide an answer to many important clinical problems posed bya BP elevation. For example, RCTs can last no more than a small fraction of the life expectancy of most hypertensive patients, which means that the current recommendation of life-long hypertension treatment is necessarily based on extrapolation from much shorter time data. Recommendations in very old (>85 years of age) and young people pose unsurmountable difficulties because RCTs, have never been done at these extreme ages. Whether BP-lowering interventions (drug-based or device-based) lead to patient protection in true resistant hypertension is unknown and this is the case also for BP reduction in common conditions such as MH and WCH. Even the popular use of ABPM and HBPM is not validated by any trial in which traditional office BP-guided treatment is compared with out-of-office BP-guided treatment or the two intervention strategies together are compared with one or the other intervention strategy alone. Furthermore, even when RCTs are available, the transferability of their results to clinical practice can be problematic because RCTs are conducted with a superior level of expertise and in an environment, that guarantees fewer errors, a much better treatment adherence [1728] and a lower therapeutic inertia than in real-life practice. The gaps between the two situations are particularly evident in the assessment of drug tolerability, usually considerably more optimistic when initially investigated in trials than when later addressed by real-word studies, in which context even previously unsuspected side effects may emerge. Hypertension management as well as management of diabetes, dyslipidemia and other chronic diseases can now gain important information from additional research approaches [477], which were previously downgraded due to a greater risk of bias due to confounding. Local, regional and even national registries, administrative or health utilization databases extended on regional or national level and extensive biobank data, most covering long periods of time are now available and suitable data sources for addressing problems unaddressed by trials. Originally available in the United States and run by public health organizations or medical insurance companies, these databases are now available in most European countries, where they can collect data from large proportions of or even the entire population, which in Europe, has the advantage of a greater residential stability than in the United States. Additional advantages over RCTs are that (i) compared with the relative homogeneity of trial data, these databases reflect the real-world patient heterogeneity, thereby offering better options for the development of precision or individualized medicine and (ii) their results can be obtained at reduced cost and much more quickly than in trials. Quick data collection is an especially important advantage, as experienced with the recent COVID-19 pandemic, during which collection of trial data was unfeasible and responses to important public health questions by trials unavailable. Future use of these approaches will be facilitated by statistical methods that allow equalization of compared groups at baseline or sophisticated testing for unmeasured confounders. As the temporal length of these bases is now considerable, future studies will more and more frequently be able to compare treatment data within individuals, with a substantial reduction of the confounding associated with comparisons of nonrandomized groups of patients [477]. Lastly, regional- and nationwide administrative and clinical databases may now be used as a foundation for register-based RCTs, combining many advantages related to recruitment, FU and generalizability from observational studies with the unbiased estimates derived from an experimental design [1729]. Several contributions of real-word research to present knowledge of hypertension epidemiology, diagnosis and treatment have been mentioned in the present guidelines and considered for the guideline recommendations. These contributions will grow in the future and will thus have to be taken into progressively greater consideration by physicians in their update of research progress and knowledge on hypertension. Although complex and somewhat controversial, this growth will include the analysis of big real-word datasets by the machine learning and artificial intelligence approaches [1730,1731]. Machine learning aims at processing complex databases automatically by methods that make use of highly sophisticated statistics in order to develop new diagnostic and treatment algorithms [1732]. Artificial intelligence takes the analysis further by including analytical processes specific of human-related decision steps. This approach extends to a large variety of human activities and in hypertension has produced promising results for an improvement of the ability to predict the risk of incident hypertension and future organ damage [1733,1734]. Promising results have also been obtained on the possibility to personalize antihypertensive treatment [1735,1736]. Improved prediction of future hypertension and HMOD development by these approaches would be of particular importance, because of the possibility to focus intensive preventive treatment on people at greater risk (Table 27).

Table 27 - Gaps in the evidence
Epidemiology and risk
- Association between BP levels in children and adolescents and the risk for clinical CV and kidney outcomes
- Trajectories of BP and hypertension phenotypes throughout life and their association with CV and kidney outcomes
- The optimal SBP and DBP level at different time points in life
- Predictive ability and therapeutic responsiveness of HMOD
- Incremental benefit of more advanced risk estimation (SCORE2 => HMOD => vascular imaging/polygenic risk scores)
- Incremental accuracy of risk estimation by use of short and long term BP variability
Diagnostic procedures
- Benefits of screening
- Optimal interval for reassessment of BP in nonhypertensive patients
- Does the incremental prognostic ability of ABPM and HBPM substantially improve diagnosis and treatment?
- Association of ABPM and HBPM with CV and kidney outcomes by serial ABPM and HBPM measurements
- Validity and application of cuffless BP measurement devices
- Optimal BP measurement methods and interpretation of BP values in AF
Treatment strategies
- Optimal time-point and BP level to initiate treatment in young patients
- Optimal and safe BP thresholds and targets in very old and frail patients
- Office vs out-of-office guided treatment on clinical outcomes
- BP thresholds and targets in low-to moderate risk individuals
- BP thresholds and targets in specific patient groups (LVH, ISH, CKD, people aged 80 years or older)
- BP thresholds and targets using ABPM and HBPM
- Treatment effect on clinical outcomes in MH and WCH
- Effect of nocturnal BP reduction by treatment on clinical outcomes
- Effect of lifestyle interventions of CV outcomes
- Strategies to implement lifestyle recommendations effectively
- Choice of first-line antihypertensive agent and sequence of titration from a population and individual level perspective
- Effectiveness and implementation strategies for individualized antihypertensive treatment
- Effect of device-based therapy (RDN) on CV and kidney outcomes
- Effect of drug treatment of true resistant hypertension on CV and kidney events
- Effects of down-titration and treatment withdrawal in different clinical settings
Follow-up
- Optimal timing and frequency of follow-up
- Optimal BP measurement modality (OBP, HBPM, ABPM) for follow-up
- The role of cuff-less devices for monitoring
- Effect of distance monitoring and digital alert systems on clinical outcomes
- Evaluation of, and interventions to improve, adherence

ACKNOWLEDGEMENTS

We wish to sincerely thank Medycyna Praktyczna Publishers, Krakow (Poland), and in particular Doctor Wiktoria Leśniak and Jacek Zieliński for their excellent support in the preparation of the graphical content of the Guidelines. We also express our sincere gratitude to Mandy Elgner (ESH coordinator) and Clara Sincich for their assistance and help during the preparation of the Guidelines.

Conflicts of interest

The conflict of interest declaration of Task Force members were compiled into one file that can be found on the ESH website: https://www.eshonline.org/guidelines/2023-guidelines/

The Task Force received its entire financial support exclusively from ESH.

REFERENCES

1. 2003 European Society of Hypertension-European Society of Cardiology guidelines for the management of arterial hypertension. J Hypertens 2003; 21:1011–1053.
2. Mancia G, De Backer G, Dominiczak A, Cifkova R, Fagard R, Germano G, et al. 2007 Guidelines for the Management of Arterial Hypertension: The Task Force for the Management of Arterial Hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). J Hypertens 2007; 25:1105–1187.
3. Mancia G, Fagard R, Narkiewicz K, Redon J, Zanchetti A, Bohm M, et al. 2013 ESH/ESC Guidelines for the management of arterial hypertension: the Task Force for the management of arterial hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). J Hypertens 2013; 31:1281–1357.
4. Williams B, Mancia G, Spiering W, Agabiti Rosei E, Azizi M, Burnier M, et al. 2018 ESC/ESH Guidelines for the management of arterial hypertension: The Task Force for the management of arterial hypertension of the European Society of Cardiology and the European Society of Hypertension: The Task Force for the management of arterial hypertension of the European Society of Cardiology and the European Society of Hypertension. J Hypertens 2018; 36:1953–2041.
5. Mancia G, Laurent S, Agabiti-Rosei E, Ambrosioni E, Burnier M, Caulfield MJ, et al. Reappraisal of European guidelines on hypertension management: a European Society of Hypertension Task Force document. J Hypertens 2009; 27:2121–2158.
6. Balshem H, Helfand M, Schünemann H, Oxman A, Kunz R, Brozek J, et al. GRADE guidelines: 3. Rating the quality of evidence. J Clin Epidemiol 2011; 64:401–406.
7. Guyatt GH, Oxman AD, Kunz R, Vist GE, Falck-Ytter Y, Schünemann HJ. What is “quality of evidence” and why is it important to clinicians? Bmj 2008; 336:995–998.
8. Brunström M, Thomopoulos C, Carlberg B, Kreutz R, Mancia G. Methodological Aspects of Meta-Analyses Assessing the Effect of Blood Pressure-Lowering Treatment on Clinical Outcomes. Hypertension (Dallas, Tex: 1979) 2022; 79:491–504.
9. Schünemann HJ, Oxman AD, Brozek J, Glasziou P, Jaeschke R, Vist GE, et al. Grading quality of evidence and strength of recommendations for diagnostic tests and strategies. Bmj 2008; 336:1106–1110.
10. Titze J, Luft FC. Speculations on salt and the genesis of arterial hypertension. Kidney Int 2017; 91:1324–1335.
11. Kelly TN, Sun X, He KY, Brown MR, Taliun SAG, Hellwege JN, et al. Insights From a Large-Scale Whole-Genome Sequencing Study of Systolic Blood Pressure, Diastolic Blood Pressure, and Hypertension. Hypertension (Dallas, Tex: 1979) 2022; 79:1656–1667.
12. Padmanabhan S, Dominiczak AF. Genomics of hypertension: the road to precision medicine. Nat Rev Cardiol 2021; 18:235–250.
13. Oparil S, Acelajado M, Bakris G, Berlowitz D, Cifkova R, Dominiczak A. Hypertension. Nat Rev Dis Primers 2018; 4:18014.
14. Mary S, Boder P, Padmanabhan S, McBride MW, Graham D, Delles C, et al. Role of Uromodulin in Salt-Sensitive Hypertension. Hypertension (Dallas, Tex: 1979) 2022; 79:2419–2429.
15. Shin S, Bai L, Oiamo TH, Burnett RT, Weichenthal S, Jerrett M, et al. Association between road traffic noise and incidence of diabetes mellitus and hypertension in Toronto, Canada: a population-based cohort study. J Am Heart Assoc 2020; 9:e013021.
16. Yang B-Y, Qian Z, Howard SW, Vaughn MG, Fan S-J, Liu K-K, et al. Global association between ambient air pollution and blood pressure: a systematic review and meta-analysis. Environ Pollut 2018; 235:576–588.
17. Kreutz R, Dobrowolski P, Prejbisz A, Algharably EAE, Bilo G, Creutzig F, et al. Lifestyle, psychological, socioeconomic and environmental factors and their impact on hypertension during the coronavirus disease 2019 pandemic. J Hypertens 2021; 39:1077–1089.
18. Kreutz R, Abd el-Hady Algharably E. Blood Pressure Control. In: Offermanns, S., Rosenthal, W., editors. Encyclopedia of Molecular Pharmacology. Cham: Springer International Publishing; 2021. pp. 317–322.
19. Mancia G, Grassi G. The autonomic nervous system and hypertension. Circ Res 2014; 114:1804–1814.
20. Harrison DG, Gongora MC. Oxidative stress and hypertension. Medical Clinics 2009; 93:621–635.
21. Jordan J, Birkenfeld AL, Melander O, Moro C. Natriuretic peptides in cardiovascular and metabolic crosstalk: implications for hypertension management. Hypertension (Dallas, Tex: 1979) 2018; 72:270–276.
22. Panza JA, Casino PR, Badar DM, Quyyumi AA. Effect of increased availability of endothelium-derived nitric oxide precursor on endothelium-dependent vascular relaxation in normal subjects and in patients with essential hypertension. Circulation 1993; 87:1475–1481.
23. Li J, Zhao F, Wang Y, Chen J, Tao J, Tian G, et al. Gut microbiota dysbiosis contributes to the development of hypertension. Microbiome 2017; 5:1–19.
24. Wilck N, Matus MG, Kearney SM, Olesen SW, Forslund K, Bartolomaeus H, et al. Salt-responsive gut commensal modulates TH17 axis and disease. Nature 2017; 551:585–589.
25. Norlander AE, Madhur MS, Harrison DG. The immunology of hypertension. J Exp Med 2018; 215:21–33.
26. Griendling KK, Camargo LL, Rios FJ, Alves-Lopes R, Montezano AC, Touyz RM. Oxidative Stress and Hypertension. Circ Res 2021; 128:993–1020.
27. Avery EG, Bartolomaeus H, Maifeld A, Marko L, Wiig H, Wilck N, et al. The Gut Microbiome in Hypertension: Recent Advances and Future Perspectives. Circ Res 2021; 128:934–950.
28. Landsberg L. Insulin-mediated sympathetic stimulation: role in the pathogenesis of obesity-related hypertension (or, how insulin affects blood pressure, and why). J Hypertens 2001; 19:523–528.
29. Lembo G, Napoli R, Capaldo B, Rendina V, Iaccarino G, Volpe M, et al. Abnormal sympathetic overactivity evoked by insulin in the skeletal muscle of patients with essential hypertension. J Clin Invest 1992; 90:24–29.
30. Page IH. Pathogenesis of arterial hypertension. J Am Med Assoc 1949; 140:451–458.
31. Harrison DG, Coffman TM, Wilcox CS. Pathophysiology of hypertension: the mosaic theory and beyond. Circ Res 2021; 128:847–863.
32. Unger T, Borghi C, Charchar F, Khan NA, Poulter NR, Prabhakaran D, et al. 2020 International Society of Hypertension Global Hypertension Practice Guidelines. Hypertension (Dallas, Tex: 1979) 2020; 75:1334–1357.
33. Visseren FLJ, Mach F, Smulders YM, Carballo D, Koskinas KC, Back M, et al. 2021 ESC Guidelines on cardiovascular disease prevention in clinical practice. Eur Heart J 2021; 42:3227–3337.
34. WHO. Guideline for the pharmacological treatment of hypertension in adults. World Health Organization; 2021 Licence: CC BY-NC-SA 30 IGO 2021.
35. Lewington S, Clarke R, Qizilbash N, Peto R, Collins R, Prospective Studies C. Age-specific relevance of usual blood pressure to vascular mortality: a meta-analysis of individual data for one million adults in 61 prospective studies. Lancet 2002; 360:1903–1913.
36. Kreutz R, Brunstrom M, Thomopoulos C, Carlberg B, Mancia G. Do recent meta-analyses truly prove that treatment with blood pressure-lowering drugs is beneficial at any blood pressure value, no matter how low? a critical review. J Hypertens 2022.
37. Worldwide trends in hypertension prevalence and progress in treatment and control from 1990 to 2019: a pooled analysis of 1201 population-representative studies with 104 million participants. Lancet 2021; 398:957–980.
38. Benetos A, Petrovic M, Strandberg T. Hypertension management in older and frail older patients. Circ Res 2019; 124:1045–1060.
39. Ji H, Kim A, Ebinger JE, Niiranen TJ, Claggett BL, Merz CNB, et al. Sex differences in blood pressure trajectories over the life course. JAMA Cardiol 2020; 5:255–262.
40. Franklin SS, Gustin IVW, Wong ND, Larson MG, Weber MA, Kannel WB, et al. Hemodynamic patterns of age-related changes in blood pressure: the Framingham Heart Study. Circulation 1997; 96:308–315.
41. Zhou B, Perel P, Mensah GA, Ezzati M. Global epidemiology, health burden and effective interventions for elevated blood pressure and hypertension. Nat Rev Cardiol 2021; 18:785–802.
42. Sundström J, Neovius M, Tynelius P, Rasmussen F. Association of blood pressure in late adolescence with subsequent mortality: cohort study of Swedish male conscripts. Bmj 2011; 342.
43. Son JS, Choi S, Kim K, Kim SM, Choi D, Lee G, et al. Association of Blood Pressure Classification in Korean Young Adults According to the 2017 American College of Cardiology/American Heart Association Guidelines With Subsequent Cardiovascular Disease Events. JAMA 2018; 320:1783–1792.
44. Domanski M, Mitchell G, Pfeffer M, Neaton JD, Norman J, Svendsen K, et al. Pulse pressure and cardiovascular disease–related mortality: follow-up study of the Multiple Risk Factor Intervention Trial (MRFIT). JAMA 2002; 287:2677–2683.
45. Franklin SS, Lopez VA, Wong ND, Mitchell GF, Larson MG, Vasan RS, et al. Single versus combined blood pressure components and risk for cardiovascular disease: the Framingham Heart Study. Circulation 2009; 119:243–250.
46. Lu S, Bao MY, Miao SM, Zhang X, Jia QQ, Jing SQ, et al. Prevalence of hypertension, diabetes, and dyslipidemia, and their additive effects on myocardial infarction and stroke: a cross-sectional study in Nanjing, China. Ann Transl Med 2019; 7:436.
47. Sehestedt T, Hansen TW, Li Y, Richart T, Boggia J, Kikuya M, et al. Are blood pressure and diabetes additive or synergistic risk factors? Outcome in 8494 subjects randomly recruited from 10 populations. Hypertens Res 2011; 34:714–721.
48. Boucheron P, Lailler G, Moutengou E, Regnault N, Gabet A, Deneux-Tharaux C, et al. Hypertensive disorders of pregnancy and onset of chronic hypertension in France: the nationwide CONCEPTION study. Eur Heart J 2022; 43:3352–3361.
49. Grandi SM, Filion KB, Yoon S, Ayele HT, Doyle CM, Hutcheon JA, et al. Cardiovascular Disease-Related Morbidity and Mortality in Women With a History of Pregnancy Complications. Circulation 2019; 139:1069–1079.
50. Grundy SM, Balady GJ, Criqui MH, Fletcher G, Greenland P, Hiratzka LF, et al. Primary prevention of coronary heart disease: guidance from Framingham: a statement for healthcare professionals from the AHA Task Force on Risk Reduction. American Heart Association. Circulation 1998; 97:1876–1887.
51. Haq IU, Ramsay LE, Yeo WW, Jackson PR, Wallis EJ. Is the Framingham risk function valid for northern European populations? A comparison of methods for estimating absolute coronary risk in high risk men. Heart 1999; 81:40–46.
52. Menotti A, Puddu PE, Lanti M. Comparison of the Framingham risk function-based coronary chart with risk function from an Italian population study. Eur Heart J 2000; 21:365–370.
53. Hense HW, Schulte H, Löwel H, Assmann G, Keil U. Framingham risk function overestimates risk of coronary heart disease in men and women from Germany--results from the MONICA Augsburg and the PROCAM cohorts. Eur Heart J 2003; 24:937–945.
54. Conroy RM, Pyörälä K, Fitzgerald AP, Sans S, Menotti A, De Backer G, et al. Estimation of ten-year risk of fatal cardiovascular disease in Europe: the SCORE project. Eur Heart J 2003; 24:987–1003.
55. SCORE2 risk prediction algorithms: new models to estimate 10-year risk of cardiovascular disease in Europe. Eur Heart J 2021; 42:2439–2454.
56. SCORE2-OP risk prediction algorithms: estimating incident cardiovascular event risk in older persons in four geographical risk regions. Eur Heart J 2021; 42:2455–2467.
57. Jaspers NEM, Blaha MJ, Matsushita K, van der Schouw YT, Wareham NJ, Khaw KT, et al. Prediction of individualized lifetime benefit from cholesterol lowering, blood pressure lowering, antithrombotic therapy, and smoking cessation in apparently healthy people. Eur Heart J 2020; 41:1190–1199.
58. Schmidt BM, Durao S, Toews I, Bavuma CM, Hohlfeld A, Nury E, et al. Screening strategies for hypertension. Cochrane Database Syst Rev 2020; 5:Cd013212.
59. Tanner L, Kenny R, Still M, Ling J, Pearson F, Thompson K, et al. NHS Health Check programme: a rapid review update. BMJ Open 2022; 12:e052832.
60. Krist AH, Davidson KW, Mangione CM, Cabana M, Caughey AB, Davis EM, et al. Screening for Hypertension in Adults: US Preventive Services Task Force Reaffirmation Recommendation Statement. JAMA 2021; 325:1650–1656.
61. Victor RG, Blyler CA, Li N, Lynch K, Moy NB, Rashid M, et al. Sustainability of Blood Pressure Reduction in Black Barbershops. Circulation 2019; 139:10–19.
62. Stergiou GS, Palatini P, Parati G, O’Brien E, Januszewicz A, Lurbe E, et al. 2021 European Society of Hypertension practice guidelines for office and out-of-office blood pressure measurement. J Hypertens 2021; 39:1293–1302.
63. Imholz BP, van Montfrans GA, Settels JJ, van der Hoeven GM, Karemaker JM, Wieling W. Continuous non-invasive blood pressure monitoring: reliability of Finapres device during the Valsalva manoeuvre. Cardiovasc Res 1988; 22:390–397.
64. Parati G, Casadei R, Groppelli A, Di Rienzo M, Mancia G. Comparison of finger and intra-arterial blood pressure monitoring at rest and during laboratory testing. Hypertension (Dallas, Tex: 1979) 1989; 13:647–655.
65. Imholz BP, Langewouters GJ, van Montfrans GA, Parati G, van Goudoever J, Wesseling KH, et al. Feasibility of ambulatory, continuous 24-hour finger arterial pressure recording. Hypertension (Dallas, Tex: 1979) 1993; 21:65–73.
66. O’Brien E, Fitzgerald D. The history of blood pressure measurement. J Hum Hypertens 1994; 8:73–84.
67. Zanchetti A, Mancia G. The centenary of blood pressure measurement:a tribute to Scipione Riva-Rocci. J Hypertens 1996; 14:1–12.
68. Stergiou GS, Mukkamala R, Avolio A, Kyriakoulis KG, Mieke S, Murray A, et al. Cuffless blood pressure measuring devices: review and statement by the European Society of Hypertension Working Group on Blood Pressure Monitoring and Cardiovascular Variability. J Hypertens 2022; 40:1449–1460.
69. Mancia G. Short- and long-term blood pressure variability: present and future. Hypertension (Dallas, Tex: 1979) 2012; 60:512–517.
70. Asayama K, Fujiwara T, Hoshide S, Ohkubo T, Kario K, Stergiou GS, et al. Nocturnal blood pressure measured by home devices: evidence and perspective for clinical application. J Hypertens 2019; 37:905–916.
71. Mukkamala R, Yavarimanesh M, Natarajan K, Hahn JO, Kyriakoulis KG, Avolio AP, et al. Evaluation of the Accuracy of Cuffless Blood Pressure Measurement Devices: Challenges and Proposals. Hypertension (Dallas, Tex: 1979) 2021; 78:1161–1167.
72. Stergiou GS, Alpert B, Mieke S, Asmar R, Atkins N, Eckert S, et al. A universal standard for the validation of blood pressure measuring devices: Association for the Advancement of Medical Instrumentation/European Society of Hypertension/International Organization for Standardization (AAMI/ESH/ISO) Collaboration Statement. J Hypertens 2018; 36:472–478.
73. Mukkamala R, Shroff SG, Landry C, Kyriakoulis KG, Avolio AP, Stergiou GS. The Microsoft Research Aurora Project: Important Findings on Cuffless Blood Pressure Measurement. Hypertension 2022; submitted.
74. Bradley CK, Shimbo D, Colburn DA, Pugliese DN, Padwal R, Sia SK, et al. Cuffless Blood Pressure Devices. Am J Hypertens 2022; 35:380–387.
75. Stergiou GS, Alpert BS, Mieke S, Wang J, O’Brien E. Validation protocols for blood pressure measuring devices in the 21st century. J Clin Hypertens (Greenwich) 2018; 20:1096–1099.
76. O’Brien E, Stergiou G, Palatini P, Asmar R, Ioannidis JP, Kollias A, et al. Validation protocols for blood pressure measuring devices: the impact of the European Society of Hypertension International Protocol and the development of a Universal Standard. Blood Press Monit 2019; 24:163–166.
77. Picone DS, Deshpande RA, Schultz MG, Fonseca R, Campbell NRC, Delles C, et al. Nonvalidated Home Blood Pressure Devices Dominate the Online Marketplace in Australia: Major Implications for Cardiovascular Risk Management. Hypertension (Dallas, Tex: 1979) 2020; 75:1593–1599.
78. Stergiou GS, O’Brien E, Myers M, Palatini P, Parati G, Kollias A, et al. STRIDE BP international initiative for accurate blood pressure measurement: Systematic review of published validation studies of blood pressure measuring devices. J Clin Hypertens (Greenwich) 2019; 21:1616–1622.
79. Turner MJ, Speechly C, Bignell N. Sphygmomanometer calibration--why, how and how often? Aust Fam Physician 2007; 36:834–838.
80. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure lowering on outcome incidence in hypertension. 1. Overview, meta-analyses, and meta-regression analyses of randomized trials. J Hypertens 2014; 32:2285–2295.
81. Stergiou GS, Kyriakoulis KG, Kollias A. Office blood pressure measurement types: Different methodology-Different clinical conclusions. J Clin Hypertens (Greenwich) 2018; 20:1683–1685.
82. Kallioinen N, Hill A, Horswill MS, Ward HE, Watson MO. Sources of inaccuracy in the measurement of adult patients’ resting blood pressure in clinical settings: a systematic review. J Hypertens 2017; 35:421–441.
83. Sprafka JM, Strickland D, Gomez-Marin O, Prineas RJ. The effect of cuff size on blood pressure measurement in adults. Epidemiology 1991; 2:214–217.
84. Palatini P, Asmar R, O’Brien E, Padwal R, Parati G, Sarkis J, et al. Recommendations for blood pressure measurement in large arms in research and clinical practice: position paper of the European society of hypertension working group on blood pressure monitoring and cardiovascular variability. J Hypertens 2020; 38:1244–1250.
85. Plumettaz C, Viswanathan B, Bovet P. Hypertension Prevalence Based on Blood Pressure Measurements on Two vs. One Visits: A Community-Based Screening Programme and a Narrative Review. Int J Environ Res Public Health 2020; 17:
86. Kronish IM, Edmondson D, Shimbo D, Shaffer JA, Krakoff LR, Schwartz JE. A Comparison of the Diagnostic Accuracy of Common Office Blood Pressure Measurement Protocols. Am J Hypertens 2018; 31:827–834.
87. Jose AP, Awasthi A, Kondal D, Kapoor M, Roy A, Prabhakaran D. Impact of repeated blood pressure measurement on blood pressure categorization in a population-based study from India. J Hum Hypertens 2019; 33:594–601.
88. Handler J, Zhao Y, Egan BM. Impact of the number of blood pressure measurements on blood pressure classification in US adults: NHANES 1999-2008. J Clin Hypertens (Greenwich) 2012; 14:751–759.
89. Sakhuja S, Jaeger BC, Akinyelure OP, Bress AP, Shimbo D, Schwartz JE, et al. Potential impact of systematic and random errors in blood pressure measurement on the prevalence of high office blood pressure in the United States. J Clin Hypertens (Greenwich) 2022; 24:263–270.
90. Clark CE, Warren FC, Boddy K, McDonagh STJ, Moore SF, Teresa Alzamora M, et al. Higher Arm Versus Lower Arm Systolic Blood Pressure and Cardiovascular Outcomes: a Meta-Analysis of Individual Participant Data From the INTERPRESS-IPD Collaboration. Hypertension (Dallas, Tex: 1979) 2022; 79:2328–2335.
91. Clark CE, Taylor RS, Shore AC, Ukoumunne OC, Campbell JL. Association of a difference in systolic blood pressure between arms with vascular disease and mortality: a systematic review and meta-analysis. Lancet 2012; 379:905–914.
92. Myers MG. A Short History of Automated Office Blood Pressure - 15 Years to SPRINT. J Clin Hypertens (Greenwich) 2016; 18:721–724.
93. Johnson KC, Whelton PK, Cushman WC, Cutler JA, Evans GW, Snyder JK, et al. Blood Pressure Measurement in SPRINT (Systolic Blood Pressure Intervention Trial). Hypertension (Dallas, Tex: 1979) 2018; 71:848–857.
94. Roerecke M, Kaczorowski J, Myers MG. Comparing Automated Office Blood Pressure Readings With Other Methods of Blood Pressure Measurement for Identifying Patients With Possible Hypertension: A Systematic Review and Meta-analysis. JAMA Intern Med 2019; 179:351–362.
95. Grassi G, Quarti-Trevano F, Seravalle G, Dell’Oro R, Vanoli J, Perseghin G, et al. Sympathetic Neural Mechanisms Underlying Attended and Unattended Blood Pressure Measurement. Hypertension (Dallas, Tex: 1979) 2021; 78:1126–1133.
96. Seo J, Lee CJ, Oh J, Lee SH, Kang SM, Park S. Large discrepancy between unobserved automated office blood pressure and ambulatory blood pressure in a high cardiovascular risk cohort. J Hypertens 2019; 37:42–49.
97. Wright JT Jr, Williamson JD, Whelton PK, Snyder JK, Sink KM, Rocco MV, et al. A Randomized Trial of Intensive versus Standard Blood-Pressure Control. N Engl J Med 2015; 373:2103–2116.
98. Myers MG, Kaczorowski J, Paterson JM, Dolovich L, Tu K. Thresholds for diagnosing hypertension based on automated office blood pressure measurements and cardiovascular risk. Hypertension (Dallas, Tex: 1979) 2015; 66:489–495.
99. Le VV, Mitiku T, Sungar G, Myers J, Froelicher V. The blood pressure response to dynamic exercise testing: a systematic review. Prog Cardiovasc Dis 2008; 51:135–160.
100. Schultz MG, La Gerche A, Sharman JE. Blood Pressure Response to Exercise and Cardiovascular Disease. Curr Hypertens Rep 2017; 19:89.
101. Niebauer J, Borjesson M, Carre F, Caselli S, Palatini P, Quattrini F, et al. Recommendations for participation in competitive sports of athletes with arterial hypertension: a position statement from the sports cardiology section of the European Association of Preventive Cardiology (EAPC). Eur Heart J 2018; 39:3664–3671.
102. Stergiou GS, Parati G, Asmar R, O’Brien E. European Society of Hypertension Working Group on Blood Pressure M. Requirements for professional office blood pressure monitors. J Hypertens 2012; 30:537–542.
103. Bonso E, Saladini F, Zanier A, Benetti E, Dorigatti F, Palatini P. Accuracy of a single rigid conical cuff with standard-size bladder coupled to an automatic oscillometric device over a wide range of arm circumferences. Hypertens Res 2010; 33:1186–1191.
104. Sharman JE, O’Brien E, Alpert B, Schutte AE, Delles C, Hecht Olsen M, et al. Lancet Commission on Hypertension group position statement on the global improvement of accuracy standards for devices that measure blood pressure. J Hypertens 2020; 38:21–29.
105. Stergiou GS, Kyriakoulis KG, Bountzona I, Menti A, Destounis A, Kalogeropoulos P, et al. Automated blood pressure measurement in atrial fibrillation: validation process modification and evaluation of a novel professional device which detects atrial fibrillation and adapts its blood pressure measurement algorithm. J Hypertens 2021; 39:614–620.
106. Lakhal K, Ehrmann S, Boulain T. Noninvasive BP Monitoring in the Critically Ill: Time to Abandon the Arterial Catheter? Chest 2018; 153:1023–1039.
107. Lehman LW, Saeed M, Talmor D, Mark R, Malhotra A. Methods of blood pressure measurement in the ICU. Crit Care Med 2013; 41:34–40.
108. Wax DB, Lin HM, Leibowitz AB. Invasive and concomitant noninvasive intraoperative blood pressure monitoring: observed differences in measurements and associated therapeutic interventions. Anesthesiology 2011; 115:973–978.
109. Picone DS, Schultz MG, Otahal P, Aakhus S, Al-Jumaily AM, Black JA, et al. Accuracy of Cuff-Measured Blood Pressure: Systematic Reviews and Meta-Analyses. J Am Coll Cardiol 2017; 70:572–586.
110. Meidert AS, Dolch ME, Muhlbauer K, Zwissler B, Klein M, Briegel J, et al. Oscillometric versus invasive blood pressure measurement in patients with shock: a prospective observational study in the emergency department. J Clin Monit Comput 2021; 35:387–393.
111. Sharman JE, Avolio AP, Baulmann J, Benetos A, Blacher J, Blizzard CL, et al. Validation of non-invasive central blood pressure devices: ARTERY Society task force consensus statement on protocol standardization. Eur Heart J 2017; 38:2805–2812.
112. McEniery CM, Cockcroft JR, Roman MJ, Franklin SS, Wilkinson IB. Central blood pressure: current evidence and clinical importance. Eur Heart J 2014; 35:1719–1725.
113. Boutouyrie P, Achouba A, Trunet P, Laurent S. Amlodipine-valsartan combination decreases central systolic blood pressure more effectively than the amlodipine-atenolol combination: the EXPLOR study. Hypertension (Dallas, Tex: 1979) 2010; 55:1314–1322.
114. Kollias A, Lagou S, Zeniodi ME, Boubouchairopoulou N, Stergiou GS. Association of Central Versus Brachial Blood Pressure With Target-Organ Damage: Systematic Review and Meta-Analysis. Hypertension (Dallas, Tex: 1979) 2016; 67:183–190.
115. Desbiens LC, Fortier C, Nadeau-Fredette AC, Madore F, Hametner B, Wassertheurer S, et al. Prediction of Cardiovascular Events by Pulse Waveform Parameters: Analysis of CARTaGENE. J Am Heart Assoc 2022; 11:e026603.
116. Li WF, Huang YQ, Feng YQ. Association between central haemodynamics and risk of all-cause mortality and cardiovascular disease: a systematic review and meta-analysis. J Hum Hypertens 2019; 33:531–541.
117. Liu W, Ye Y, Wang L, Gao C, Bai Y, Chu H, et al. Central versus ambulatory blood pressure for predicting mortality and cardiovascular events in hemodialysis patients: a multicenter cohort study. J Hypertens 2022; 40:180–188.
118. Rahman M, Hsu JY, Desai N, Hsu CY, Anderson AH, Appel LJ, et al. Central Blood Pressure and Cardiovascular Outcomes in Chronic Kidney Disease. Clin J Am Soc Nephrol 2018; 13:585–595.
119. Cheng YB, Thijs L, Aparicio LS, Huang QF, Wei FF, Yu YL, et al. Risk Stratification by Cross-Classification of Central and Brachial Systolic Blood Pressure. Hypertension (Dallas, Tex: 1979) 2022; 79:1101–1111.
120. Herbert A, Cruickshank JK, Laurent S, Boutouyrie P. Reference Values for Arterial Measurements C. Establishing reference values for central blood pressure and its amplification in a general healthy population and according to cardiovascular risk factors. Eur Heart J 2014; 35:3122–3133.
121. Palatini P, Rosei EA, Avolio A, Bilo G, Casiglia E, Ghiadoni L, et al. Isolated systolic hypertension in the young: a position paper endorsed by the European Society of Hypertension. J Hypertens 2018; 36:1222–1236.
122. Parati G, Stergiou GS, Bilo G, Kollias A, Pengo M, Ochoa JE, et al. Home blood pressure monitoring: methodology, clinical relevance and practical application: a 2021 position paper by the Working Group on Blood Pressure Monitoring and Cardiovascular Variability of the European Society of Hypertension. J Hypertens 2021; 39:1742–1767.
123. Stergiou GS, Baibas NM, Gantzarou AP, Skeva II, Kalkana CB, Roussias LG, et al. Reproducibility of home, ambulatory, and clinic blood pressure: implications for the design of trials for the assessment of antihypertensive drug efficacy. Am J Hypertens 2002; 15:101–104.
124. Guo QH, Cheng YB, Zhang DY, Wang Y, Huang QF, Sheng CS, et al. Comparison Between Home and Ambulatory Morning Blood Pressure and Morning Hypertension in Their Reproducibility and Associations With Vascular Injury. Hypertension (Dallas, Tex: 1979) 2019; 74:137–144.
125. Niiranen TJ, Asayama K, Thijs L, Johansson JK, Ohkubo T, Kikuya M, et al. Outcome-driven thresholds for home blood pressure measurement: international database of home blood pressure in relation to cardiovascular outcome. Hypertension (Dallas, Tex: 1979) 2013; 61:27–34.
126. Ward AM, Takahashi O, Stevens R, Heneghan C. Home measurement of blood pressure and cardiovascular disease: systematic review and meta-analysis of prospective studies. J Hypertens 2012; 30:449–456.
127. Kario K, Hasebe N, Okumura K, Yamashita T, Akao M, Atarashi H, et al. Home Blood Pressure Can Predict the Risk for Stroke/Bleeding Events in Elderly Patients With Nonvalvular Atrial Fibrillation From the ANAFIE Registry. Hypertension (Dallas, Tex: 1979) 2022.
128. Piper MA, Evans CV, Burda BU, Margolis KL, O’Connor E, Whitlock EP. Diagnostic and predictive accuracy of blood pressure screening methods with consideration of rescreening intervals: a systematic review for the U.S. Preventive Services Task Force. Ann Intern Med 2015; 162:192–204.
129. Mancia G, Facchetti R, Seravalle G, Cuspidi C, Corrao G, Grassi G. Adding Home and/or Ambulatory Blood Pressure to Office Blood Pressure for Cardiovascular Risk Prediction. Hypertension (Dallas, Tex: 1979) 2021; 77:640–649.
130. Sega R, Facchetti R, Bombelli M, Cesana G, Corrao G, Grassi G, et al. Prognostic value of ambulatory and home blood pressures compared with office blood pressure in the general population: follow-up results from the Pressioni Arteriose Monitorate e Loro Associazioni (PAMELA) study. Circulation 2005; 111:1777–1783.
131. Matsumoto A, Satoh M, Kikuya M, Ohkubo T, Hirano M, Inoue R, et al. Day-to-day variability in home blood pressure is associated with cognitive decline: the Ohasama study. Hypertension (Dallas, Tex: 1979) 2014; 63:1333–1338.
132. Kario K, Shimbo D, Hoshide S, Wang JG, Asayama K, Ohkubo T, et al. Emergence of Home Blood Pressure-Guided Management of Hypertension Based on Global Evidence. Hypertension (Dallas, Tex: 1979) 2019; HYPERTENSIONAHA11912630.
133. Stergiou GS, Asayama K, Thijs L, Kollias A, Niiranen T J, Hozawa A, et al. Prognosis of white-coat and masked hypertension: International Database of HOme blood pressure in relation to Cardiovascular Outcome. Hypertension (Dallas, Tex: 1979) 2014; 63:675–682.
134. Mancia G, Facchetti R, Bombelli M, Grassi G, Sega R. Long-term risk of mortality associated with selective and combined elevation in office, home, and ambulatory blood pressure. Hypertension (Dallas, Tex: 1979) 2006; 47:846–853.
135. Fletcher BR, Hartmann-Boyce J, Hinton L, McManus RJ. The Effect of Self-Monitoring of Blood Pressure on Medication Adherence and Lifestyle Factors: A Systematic Review and Meta-Analysis. Am J Hypertens 2015; 28:1209–1221.
136. McManus RJ, Mant J, Haque MS, Bray EP, Bryan S, Greenfield SM, et al. Effect of self-monitoring and medication self-titration on systolic blood pressure in hypertensive patients at high risk of cardiovascular disease: the TASMIN-SR randomized clinical trial. JAMA 2014; 312:799–808.
137. McManus RJ, Little P, Stuart B, Morton K, Raftery J, Kelly J, et al. Home and Online Management and Evaluation of Blood Pressure (HOME BP) using a digital intervention in poorly controlled hypertension: randomised controlled trial. BMJ 2021; 372:m4858.
138. Omboni S, McManus RJ, Bosworth HB, Chappell LC, Green BB, Kario K, et al. Evidence and Recommendations on the Use of Telemedicine for the Management of Arterial Hypertension: An International Expert Position Paper. Hypertension (Dallas, Tex: 1979) 2020; 76:1368–1383.
139. Khan NA, Stergiou GS, Omboni S, Kario K, Renna N, Chapman N, et al. Virtual management of hypertension: lessons from the COVID-19 pandemic-International Society of Hypertension position paper endorsed by the World Hypertension League and European Society of Hypertension. J Hypertens 2022; 40:1435–1448.
140. Ntineri A, Niiranen TJ, McManus RJ, Lindroos A, Jula A, Schwartz C, et al. Ambulatory versus home blood pressure monitoring: frequency and determinants of blood pressure difference and diagnostic disagreement. J Hypertens 2019; 37:1974–1981.
141. Parati G, Pomidossi G, Casadei R, Mancia G. Lack of alerting reactions to intermittent cuff inflations during noninvasive blood pressure monitoring. Hypertension (Dallas, Tex: 1979) 1985; 7:597–601.
142. Sega R, Cesana G, Milesi C, Grassi G, Zanchetti A, Mancia G. Ambulatory and home blood pressure normality in the elderly: data from the PAMELA population. Hypertension (Dallas, Tex: 1979) 1997; 30:1–6.
143. Mancia G, Sega R, Bravi C, De Vito G, Valagussa F, Cesana G, et al. Ambulatory blood pressure normality: results from the PAMELA study. J Hypertens 1995; 13:1377–1390.
144. Thomopoulos C. Target blood pressure in isolated systolic hypertension. A meta-analysis of randomized outcome trials. J Hypertens 2023; in press.
145. Wang JG, Staessen JA, Gong L, Liu L. Chinese trial on isolated systolic hypertension in the elderly. Systolic Hypertension in China (Syst-China) Collaborative Group. Arch Intern Med 2000; 160:211–220.
146. Mancia G, Parati G, Bilo G, Choi J, Kilama MO, Ruilope LM. Blood pressure control by the nifedipine GITS-telmisartan combination in patients at high cardiovascular risk: the TALENT study. J Hypertens 2011; 29:600–609.
147. Kyriakoulis KG, Ntineri A, Niiranen TJ, Lindroos A, Jula A, Schwartz C, et al. Home blood pressure monitoring schedule: optimal and minimum based on 2122 individual participants’ data. J Hypertens 2022; 40:1380–1387.
148. Hodgkinson JA, Stevens R, Grant S, Mant J, Bray EP, Hobbs FDR, et al. Schedules for Self-monitoring Blood Pressure: A Systematic Review. Am J Hypertens 2019; 32:350–364.
149. Barochiner J, Aparicio LS, Martínez R, Boggia J. Prognostic value of home blood pressure monitoring in patients under antihypertensive treatment. J Hum Hypertens 2022; 1–8.
150. Kollias A, Ntineri A, Stergiou GS. Association of night-time home blood pressure with night-time ambulatory blood pressure and target-organ damage: a systematic review and meta-analysis. J Hypertens 2017; 35:442–452.
151. Kario K, Kanegae H, Tomitani N, Okawara Y, Fujiwara T, Yano Y, et al. Nighttime blood pressure measured by home blood pressure monitoring as an independent predictor of cardiovascular events in General Practice. Hypertension (Dallas, Tex: 1979) 2019; 73:1240–1248.
152. Staplin N, de la Sierra A, Ruilope LM, Emberson JR, Vinyoles E, Gorostidi M, et al. Relationship between clinic and ambulatory blood pressure and mortality: an observational cohort study in 59 124 patients. Lancet 2023; 401:2041–2050.
153. Clement DL, De Buyzere ML, De Bacquer DA, de Leeuw PW, Duprez DA, Fagard RH, et al. Prognostic value of ambulatory blood-pressure recordings in patients with treated hypertension. N Engl J Med 2003; 348:2407–2415.
154. Mancia G. Evidence in favour of ambulatory blood pressure grows but gaps in knowledge remain. Lancet 2023; 401:2014–2015.
155. Redon J, Campos C, Narciso ML, Rodicio JL, Pascual JM, Ruilope LM. Prognostic value of ambulatory blood pressure monitoring in refractory hypertension: a prospective study. Hypertension (Dallas, Tex: 1979) 1998; 31:712–718.
156. Kario K, Pickering TG, Umeda Y, Hoshide S, Hoshide Y, Morinari M, et al. Morning surge in blood pressure as a predictor of silent and clinical cerebrovascular disease in elderly hypertensives: a prospective study. Circulation 2003; 107:1401–1406.
157. Mancia G, Bombelli M, Facchetti R, Madotto F, Corrao G, Trevano FQ, et al. Long-term prognostic value of blood pressure variability in the general population: results of the Pressioni Arteriose Monitorate e Loro Associazioni Study. Hypertension (Dallas, Tex: 1979) 2007; 49:1265–1270.
158. Li Y, Wang JG. Isolated nocturnal hypertension: a disease masked in the dark. Hypertension (Dallas, Tex: 1979) 2013; 61:278–283.
159. Mancia G, Verdecchia P. Clinical value of ambulatory blood pressure: evidence and limits. Circ Res 2015; 116:1034–1045.
160. Mancia G, Omboni S, Parati G, Ravogli A, Villani A, Zanchetti A. Lack of placebo effect on ambulatory blood pressure. Am J Hypertens 1995; 8:311–315.
161. O’Brien E, Parati G, Stergiou G, Asmar R, Beilin L, Bilo G, et al. European Society of Hypertension position paper on ambulatory blood pressure monitoring. J Hypertens 2013; 31:1731–1768.
162. Yang WY, Thijs L, Zhang ZY, Asayama K, Boggia J, Hansen TW, et al. Evidence-based proposal for the number of ambulatory readings required for assessing blood pressure level in research settings: an analysis of the IDACO database. Blood Press 2018; 27:341–350.
163. Zanchetti A, Mancia G. Longing for clinical excellence: a critical outlook into the NICE recommendations on hypertension management--is nice always good? J Hypertens 2012; 30:660–668.
164. di Rienzo M, Grassi G, Pedotti A, Mancia G. Continuous vs intermittent blood pressure measurements in estimating 24-hour average blood pressure. Hypertension (Dallas, Tex: 1979) 1983; 5:264–269.
165. Schutte AE, Kollias A, Stergiou GS. Blood pressure and its variability: classic and novel measurement techniques. Nat Rev Cardiol 2022; 19:643–654.
166. Stevens SL, Wood S, Koshiaris C, Law K, Glasziou P, Stevens RJ, et al. Blood pressure variability and cardiovascular disease: systematic review and meta-analysis. BMJ 2016; 354:i4098.
167. Bilo G, Dolan E, O’Brien E, Facchetti R, Soranna D, Zambon A, et al. The impact of systolic and diastolic blood pressure variability on mortality is age dependent: Data from the Dublin Outcome Study. Eur J Prev Cardiol 2020; 27:355–364.
168. Sega R, Corrao G, Bombelli M, Beltrame L, Facchetti R, Grassi G, et al. Blood pressure variability and organ damage in a general population: results from the PAMELA study (Pressioni Arteriose Monitorate E Loro Associazioni). Hypertension (Dallas, Tex: 1979) 2002; 39:710–714.
169. Kollias A, Stergiou GS, Dolan E, O’Brien E. Ambulatory arterial stiffness index: a systematic review and meta-analysis. Atherosclerosis 2012; 224:291–301.
170. Schillaci G, Parati G, Pirro M, Pucci G, Mannarino MR, Sperandini L, et al. Ambulatory arterial stiffness index is not a specific marker of reduced arterial compliance. Hypertension (Dallas, Tex: 1979) 2007; 49:986–991.
171. Kips JG, Vermeersch SJ, Reymond P, Boutouyrie P, Stergiopulos N, Laurent S, et al. Ambulatory arterial stiffness index does not accurately assess arterial stiffness. J Hypertens 2012; 30:574–580.
172. Mancia G, Facchetti R, Cuspidi C, Bombelli M, Corrao G, Grassi G. Limited reproducibility of MUCH and WUCH: evidence from the ELSA study. Eur Heart J 2020; 41:1565–1571.
173. Mancia G, Facchetti R, Bombelli M, Quarti-Trevano F, Cuspidi C, Grassi G. Short- and Long-Term Reproducibility of Nighttime Blood Pressure Phenotypes and Nocturnal Blood Pressure Reduction. Hypertension (Dallas, Tex: 1979) 2021; 77:1745–1755.
174. Cuspidi C, Meani S, Valerio C, Sala C, Fusi V, Masaidi M, et al. Reproducibility of dipping/nondipping pattern in untreated essential hypertensive patients: impact of sex and age. Blood Press Monit 2007; 12:101–106.
175. Mancia G, Bombelli M, Brambilla G, Facchetti R, Sega R, Toso E, et al. Long-term prognostic value of white coat hypertension: an insight from diagnostic use of both ambulatory and home blood pressure measurements. Hypertension (Dallas, Tex: 1979) 2013; 62:168–174.
176. Bevan AT, Honour AJ, Stott FH. Direct arterial pressure recording in unrestricted man. Clin Sci 1969; 36:329–344.
177. Mancia G, Ferrari A, Gregorini L, Parati G, Pomidossi G, Bertinieri G, et al. Blood pressure and heart rate variabilities in normotensive and hypertensive human beings. Circ Res 1983; 53:96–104.
178. Frattola A, Parati G, Cuspidi C, Albini F, Mancia G. Prognostic value of 24-hour blood pressure variability. J Hypertens 1993; 11:1133–1137.
179. Palatini P, Saladini F, Mos L, Fania C, Mazzer A, Cozzio S, et al. Short-term blood pressure variability outweighs average 24-h blood pressure in the prediction of cardiovascular events in hypertension of the young. J Hypertens 2019; 37:1419–1426.
180. Parati G, Bilo G, Kollias A, Pengo M, Ochoa JE, Castiglioni P, et al. Blood pressure variability: methodological aspects, clinical relevance and practical indications for management - a European Society of Hypertension position paper ∗. J Hypertens 2023; 41:527–544.
181. Muntner P, Levitan EB, Reynolds K, Mann DM, Tonelli M, Oparil S, et al. Within-visit variability of blood pressure and all-cause and cardiovascular mortality among US adults. J Clin Hypertens (Greenwich) 2012; 14:165–171.
182. Kikuya M, Ohkubo T, Metoki H, Asayama K, Hara A, Obara T, et al. Day-by-day variability of blood pressure and heart rate at home as a novel predictor of prognosis: the Ohasama study. Hypertension (Dallas, Tex: 1979) 2008; 52:1045–1050.
183. Mancia G, Kjeldsen SE, Zappe DH, Holzhauer B, Hua TA, Zanchetti A, et al. Cardiovascular outcomes at different on-treatment blood pressures in the hypertensive patients of the VALUE trial. Eur Heart J 2016; 37:955–964.
184. Mancia G, Schumacher H, Redon J, Verdecchia P, Schmieder R, Jennings G, et al. Blood pressure targets recommended by guidelines and incidence of cardiovascular and renal events in the Ongoing Telmisartan Alone and in Combination With Ramipril Global Endpoint Trial (ONTARGET). Circulation 2011; 124:1727–1736.
185. Mancia G, Messerli F, Bakris G, Zhou Q, Champion A, Pepine CJ. Blood pressure control and improved cardiovascular outcomes in the International Verapamil SR-Trandolapril Study. Hypertension (Dallas, Tex: 1979) 2007; 50:299–305.
186. Wang J, Shi X, Ma C, Zheng H, Xiao J, Bian H, et al. Visit-to-visit blood pressure variability is a risk factor for all-cause mortality and cardiovascular disease: a systematic review and meta-analysis. J Hypertens 2017; 35:10–17.
187. Hata J, Arima H, Rothwell PM, Woodward M, Zoungas S, Anderson C, et al. Effects of visit-to-visit variability in systolic blood pressure on macrovascular and microvascular complications in patients with type 2 diabetes mellitus: the ADVANCE trial. Circulation 2013; 128:1325–1334.
188. Rothwell PM, Howard SC, Dolan E, O’Brien E, Dobson JE, Dahlof B, et al. Prognostic significance of visit-to-visit variability, maximum systolic blood pressure, and episodic hypertension. Lancet 2010; 375:895–905.
189. Mancia G, Schumacher H, Bohm M, Redon J, Schmieder RE, Verdecchia P, et al. Relative and Combined Prognostic Importance of On-Treatment Mean and Visit-to-Visit Blood Pressure Variability in ONTARGET and TRANSCEND Patients. Hypertension (Dallas, Tex: 1979) 2017; 70:938–948.
190. Wang N, Harris K, Hamet P, Harrap S, Mancia G, Poulter N, et al. Cumulative Systolic Blood Pressure Load and Cardiovascular Risk in Patients With Diabetes. J Am Coll Cardiol 2022; 80:1147–1155.
191. Mahfoud F, Mancia G, Schmieder RE, Ruilope L, Narkiewicz K, Schlaich M, et al. Cardiovascular Risk Reduction After Renal Denervation According to Time in Therapeutic Systolic Blood Pressure Range. J Am Coll Cardiol 2022; 80:1871–1880.
192. Mancia G, Parati G, Bilo G, Maronati A, Omboni S, Baurecht H, et al. Assessment of long-term antihypertensive treatment by clinic and ambulatory blood pressure: data from the European Lacidipine Study on Atherosclerosis. J Hypertens 2007; 25:1087–1094.
193. Kasiakogias A, Rosei EA, Camafort M, Ehret G, Faconti L, Ferreira JP, et al. Hypertension and heart failure with preserved ejection fraction: position paper by the European Society of Hypertension. J Hypertens 2021; 39:1522–1545.
194. Okin PM, Oikarinen L, Viitasalo M, Toivonen L, Kjeldsen SE, Nieminen MS, et al. Prognostic value of changes in the electrocardiographic strain pattern during antihypertensive treatment: the Losartan Intervention for End-Point Reduction in Hypertension Study (LIFE). Circulation 2009; 119:1883–1891.
195. Sparapani R, Dabbouseh NM, Gutterman D, Zhang J, Chen H, Bluemke DA, et al. Detection of Left Ventricular Hypertrophy Using Bayesian Additive Regression Trees: The MESA. J Am Heart Assoc 2019; 8:e009959.
196. Bacharova L, Schocken D, Estes EH, Strauss D. The role of ECG in the diagnosis of left ventricular hypertrophy. Curr Cardiol Rev 2014; 10:257–261.
197. Caroli A, Remuzzi A, Lerman LO. Basic principles and new advances in kidney imaging. Kidney Int 2021; 100:1001–1011.
198. Park BK. Gray-scale, color doppler, spectral doppler, and contrast-enhanced renal artery ultrasound: imaging techniques and features. J Clin Med 2022; 11:
199. Tsimikas S. Elevated lipoprotein(a) and the risk of stroke in children, young adults, and the elderly. Eur Heart J 2021; 42:2197–2200.
200. Kamstrup PR. Lipoprotein(a) and cardiovascular disease. Clin Chem 2021; 67:154–166.
201. Kronenberg F, Mora S, Stroes ESG, Ference BA, Arsenault BJ, Berglund L, et al. Lipoprotein(a) in atherosclerotic cardiovascular disease and aortic stenosis: a European Atherosclerosis Society consensus statement. Eur Heart J 2022; 43:3925–3946.
202. Ledwidge M, Gallagher J, Conlon C, Tallon E, O’Connell E, Dawkins I, et al. Natriuretic peptide-based screening and collaborative care for heart failure: the STOP-HF randomized trial. JAMA 2013; 310:66–74.
203. McDonagh TA, Metra M, Adamo M, Gardner RS, Baumbach A, Bohm M, et al. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J 2021; 42:3599–3726.
204. Collet JP, Thiele H, Barbato E, Barthelemy O, Bauersachs J, Bhatt DL, et al. 2020 ESC Guidelines for the management of acute coronary syndromes in patients presenting without persistent ST-segment elevation. Eur Heart J 2021; 42:1289–1367.
205. Lyon AR, Lopez-Fernandez T, Couch LS, Asteggiano R, Aznar MC, Bergler-Klein J, et al. 2022 ESC Guidelines on cardio-oncology developed in collaboration with the European Hematology Association (EHA), the European Society for Therapeutic Radiology and Oncology (ESTRO) and the International Cardio-Oncology Society (IC-OS). Eur Heart J 2022; 43:4229–4361.
206. Teaford HR, Barreto JN, Vollmer KJ, Rule AD, Barreto EF. Cystatin C: A Primer for Pharmacists. Pharmacy (Basel) 2020; 8:
207. Ebert N, Bevc S, Bokenkamp A, Gaillard F, Hornum M, Jager KJ, et al. Assessment of kidney function: clinical indications for measured GFR. Clin Kidney J 2021; 14:1861–1870.
208. Delgado C, Baweja M, Crews DC, Eneanya ND, Gadegbeku CA, Inker LA, et al. A Unifying Approach for GFR Estimation: Recommendations of the NKF-ASN Task Force on Reassessing the Inclusion of Race in Diagnosing Kidney Disease. Am J Kidney Dis 2022; 79:268–288. e261.
209. Benoit SW, Ciccia EA, Devarajan P. Cystatin C as a biomarker of chronic kidney disease: latest developments. Expert Rev Mol Diagn 2020; 20:1019–1026.
210. Nadkarni GN, Chauhan K, Rao V, Ix JH, Shlipak MG, Parikh CR, et al. Effect of Intensive Blood Pressure Lowering on Kidney Tubule Injury: Findings From the ACCORD Trial Study Participants. Am J Kidney Dis 2019; 73:31–38.
211. Ix JH, Shlipak MG. The Promise of Tubule Biomarkers in Kidney Disease: A Review. Am J Kidney Dis 2021; 78:719–727.
212. Bullen AL, Ascher SB, Scherzer R, Garimella PS, Katz R, Hallan SI, et al. Markers of Kidney Tubular Secretion and Risk of Adverse Events in SPRINT Participants with CKD. J Am Soc Nephrol 2022.
213. Lenders JWM, Kerstens MN, Amar L, Prejbisz A, Robledo M, Taieb D, et al. Genetics, diagnosis, management and future directions of research of phaeochromocytoma and paraganglioma: a position statement and consensus of the Working Group on Endocrine Hypertension of the European Society of Hypertension. J Hypertens 2020; 38:1443–1456.
214. Riveros-Mckay F, Weale ME, Moore R, Selzam S, Krapohl E, Sivley RM, et al. Integrated Polygenic Tool Substantially Enhances Coronary Artery Disease Prediction. Circ Genom Precis Med 2021; 14:e003304.
215. O'Sullivan JW, Raghavan S, Marquez-Luna C, Luzum JA, Damrauer SM, Ashley EA, et al. Polygenic Risk Scores for Cardiovascular Disease: A Scientific Statement From the American Heart Association. Circulation 2022; 146:e93–e118.
216. Perrone-Filardi P, Coca A, Galderisi M, Paolillo S, Alpendurada F, de Simone G, et al. Noninvasive cardiovascular imaging for evaluating subclinical target organ damage in hypertensive patients: a consensus article from the European Association of Cardiovascular Imaging, the European Society of Cardiology Council on Hypertension and the European Society of Hypertension. J Hypertens 2017; 35:1727–1741.
217. Mancia G, Facchetti R, Vanoli J, Dell’Oro R, Seravalle G, Grassi G. White-Coat Hypertension Without Organ Damage: Impact on Long-Term Mortality, New Hypertension, and New Organ Damage. Hypertension (Dallas, Tex: 1979) 2022; 79:1057–1066.
218. Greve SV, Blicher MK, Sehestedt T, Gram-Kampmann EM, Rasmussen S, Vishram JK, et al. Effective risk stratification in patients with moderate cardiovascular risk using albuminuria and atherosclerotic plaques in the carotid arteries. J Hypertens 2015; 33:1563–1570.
219. Bombelli M, Facchetti R, Carugo S, Madotto F, Arenare F, Quarti-Trevano F, et al. Left ventricular hypertrophy increases cardiovascular risk independently of in-office and out-of-office blood pressure values. J Hypertens 2009; 27:2458–2464.
220. Lonnebakken MT, Izzo R, Mancusi C, Gerdts E, Losi MA, Canciello G, et al. Left Ventricular Hypertrophy Regression During Antihypertensive Treatment in an Outpatient Clinic (the Campania Salute Network). J Am Heart Assoc 2017; 6:
221. Levy D, Garrison RJ, Savage DD, Kannel WB, Castelli WP. Prognostic implications of echocardiographically determined left ventricular mass in the Framingham Heart Study. N Engl J Med 1990; 322:1561–1566.
222. Koren MJ, Devereux RB, Casale PN, Savage DD, Laragh JH. Relation of left ventricular mass and geometry to morbidity and mortality in uncomplicated essential hypertension. Ann Intern Med 1991; 114:345–352.
223. Kawel-Boehm N, Kronmal R, Eng J, Folsom A, Burke G, Carr JJ, et al. Left Ventricular Mass at MRI and Long-term Risk of Cardiovascular Events: The Multi-Ethnic Study of Atherosclerosis (MESA). Radiology 2019; 293:107–114.
224. Armstrong AC, Jacobs DR Jr, Gidding SS, Colangelo LA, Gjesdal O, Lewis CE, et al. Framingham score and LV mass predict events in young adults: CARDIA study. Int J Cardiol 2014; 172:350–355.
225. Du Z, Xing L, Ye N, Lin M, Sun Y. Complementary value of ECG and echocardiographic left ventricular hypertrophy for prediction of adverse outcomes in the general population. J Hypertens 2021; 39:548–555.
226. Zalawadiya SK, Gunasekaran PC, Bavishi CP, Veeranna V, Panaich S, Afonso L. Left ventricular hypertrophy and risk reclassification for coronary events in multi-ethnic adults. Eur J Prev Cardiol 2015; 22:673–679.
227. Modin D, Biering-Sorensen SR, Mogelvang R, Landler N, Jensen JS, Biering-Sorensen T. Prognostic Value of Echocardiography in Hypertensive Versus Nonhypertensive Participants From the General Population. Hypertension (Dallas, Tex: 1979) 2018; 71:742–751.
228. Okin PM, Devereux RB, Jern S, Kjeldsen SE, Julius S, Nieminen MS, et al. Regression of electrocardiographic left ventricular hypertrophy during antihypertensive treatment and the prediction of major cardiovascular events. JAMA 2004; 292:2343–2349.
229. Devereux RB, Wachtell K, Gerdts E, Boman K, Nieminen MS, Papademetriou V, et al. Prognostic significance of left ventricular mass change during treatment of hypertension. JAMA 2004; 292:2350–2356.
230. Muiesan ML, Salvetti M, Monteduro C, Bonzi B, Paini A, Viola S, et al. Left ventricular concentric geometry during treatment adversely affects cardiovascular prognosis in hypertensive patients. Hypertension (Dallas, Tex: 1979) 2004; 43:731–738.
231. Wachtell K, Dahlöf B, Rokkedal J, Papademetriou V, Nieminen MS, Smith G, et al. Change of left ventricular geometric pattern after 1 year of antihypertensive treatment: the Losartan Intervention For Endpoint reduction in hypertension (LIFE) study. Am Heart J 2002; 144:1057–1064.
232. Zhou J, Du M, Chang S, Chen Z. Artificial intelligence in echocardiography: detection, functional evaluation, and disease diagnosis. Cardiovasc Ultrasound 2021; 19:29.
233. Marwick TH, Gillebert TC, Aurigemma G, Chirinos J, Derumeaux G, Galderisi M, et al. Recommendations on the use of echocardiography in adult hypertension: a report from the European Association of Cardiovascular Imaging (EACVI) and the American Society of Echocardiography (ASE)dagger. Eur Heart J Cardiovasc Imaging 2015; 16:577–605.
234. Garg S, de Lemos JA, Ayers C, Khouri MG, Pandey A, Berry JD, et al. Association of a 4-Tiered Classification of LV Hypertrophy With Adverse CV Outcomes in the General Population. JACC Cardiovasc Imaging 2015; 8:1034–1041.
235. Barron AJ, Hughes AD, Sharp A, Baksi AJ, Surendran P, Jabbour RJ, et al. Long-term antihypertensive treatment fails to improve E/e’ despite regression of left ventricular mass: an Anglo-Scandinavian cardiac outcomes trial substudy. Hypertension (Dallas, Tex: 1979) 2014; 63:252–258.
236. Mizukoshi K, Takeuchi M, Nagata Y, Addetia K, Lang RM, Akashi YJ, et al. Normal Values of Left Ventricular Mass Index Assessed by Transthoracic Three-Dimensional Echocardiography. J Am Soc Echocardiogr 2016; 29:51–61.
237. Lang RM, Badano LP, Tsang W, Adams DH, Agricola E, Buck T, et al. EAE/ASE recommendations for image acquisition and display using three-dimensional echocardiography. J Am Soc Echocardiogr 2012; 25:3–46.
238. Chaowu Y, Li L. Histopathological basis of myocardial late gadolinium enhancement in patients with systemic hypertension. Circulation 2014; 130:2210–2212.
239. Bhuva AN, Treibel TA, Fontana M, Herrey AS, Manisty CH, Moon JC. T1 mapping: non-invasive evaluation of myocardial tissue composition by cardiovascular magnetic resonance. Expert Rev Cardiovasc Ther 2014; 12:1455–1464.
240. Bellenger NG, Davies LC, Francis JM, Coats AJ, Pennell DJ. Reduction in sample size for studies of remodeling in heart failure by the use of cardiovascular magnetic resonance. J Cardiovasc Magn Reson 2000; 2:271–278.
241. Yaghi S, Moon YP, Mora-McLaughlin C, Willey JZ, Cheung K, Di Tullio MR, et al. Left atrial enlargement and stroke recurrence: the Northern Manhattan Stroke Study. Stroke 2015; 46:1488–1493.
242. Gerdts E, Wachtell K, Omvik P, Otterstad JE, Oikarinen L, Boman K, et al. Left atrial size and risk of major cardiovascular events during antihypertensive treatment: losartan intervention for endpoint reduction in hypertension trial. Hypertension (Dallas, Tex: 1979) 2007; 49:311–316.
243. Leone D, Airale L, Bernardi S, Mingrone G, Astarita A, Cesareo M, et al. Prognostic role of the ascending aorta dilatation in patients with arterial hypertension. J Hypertens 2021; 39:1163–1169.
244. Nagueh SF, Smiseth OA, Appleton CP, Byrd BF 3rd, Dokainish H, Edvardsen T, et al. Recommendations for the Evaluation of Left Ventricular Diastolic Function by Echocardiography: An Update from the American Society of Echocardiography and the European Association of Cardiovascular Imaging. Eur Heart J Cardiovasc Imaging 2016; 17:1321–1360.
245. Gimelli A, Lancellotti P, Badano LP, Lombardi M, Gerber B, Plein S, et al. Non-invasive cardiac imaging evaluation of patients with chronic systolic heart failure: a report from the European Association of Cardiovascular Imaging (EACVI). Eur Heart J 2014; 35:3417–3425.
246. Mor-Avi V, Lang RM, Badano LP, Belohlavek M, Cardim NM, Derumeaux G, et al. Current and evolving echocardiographic techniques for the quantitative evaluation of cardiac mechanics: ASE/EAE consensus statement on methodology and indications endorsed by the Japanese Society of Echocardiography. Eur J Echocardiogr 2011; 12:167–205.
247. Weinberg RL, Rubenfire M, Brook RD. Coronary artery calcium scoring in patients with hypertension. J Hum Hypertens 2020; 34:609–616.
248. Valenti V, Oh B, Heo R, Schulman-Marcus J, Cho I, Kalra DK, et al. Long-term prognosis for individuals with hypertension undergoing coronary artery calcium scoring. Int J Cardiol 2015; 187:534–540.
249. Saydam CD. Subclinical cardiovascular disease and utility of coronary artery calcium score. Int J Cardiol Heart Vasc 2021; 37:100909.
250. Maniar Y, Blumenthal RS, Alfaddagh A. The role of coronary artery calcium in allocating pharmacotherapy for primary prevention of cardiovascular disease: The ABCs of CAC. Clin Cardiol 2022.
251. Libby P, Ridker PM, Hansson GK. Progress and challenges in translating the biology of atherosclerosis. Nature 2011; 473:317–325.
252. Sehestedt T, Jeppesen J, Hansen TW, Wachtell K, Ibsen H, Torp-Pedersen C, et al. Risk prediction is improved by adding markers of subclinical organ damage to SCORE. Eur Heart J 2010; 31:883–891.
253. Nambi V, Chambless L, Folsom AR, He M, Hu Y, Mosley T, et al. Carotid intima-media thickness and presence or absence of plaque improves prediction of coronary heart disease risk: the ARIC (Atherosclerosis Risk In Communities) study. J Am Coll Cardiol 2010; 55:1600–1607.
254. Vlachopoulos C, Xaplanteris P, Aboyans V, Brodmann M, Cífková R, Cosentino F, et al. The role of vascular biomarkers for primary and secondary prevention. A position paper from the European Society of Cardiology Working Group on peripheral circulation: Endorsed by the Association for Research into Arterial Structure and Physiology (ARTERY) Society. Atherosclerosis 2015; 241:507–532.
255. Den Ruijter HM, Peters SA, Anderson TJ, Britton AR, Dekker JM, Eijkemans MJ, et al. Common carotid intima-media thickness measurements in cardiovascular risk prediction: a meta-analysis. JAMA 2012; 308:796–803.
256. Lorenz MW, Schaefer C, Steinmetz H, Sitzer M. Is carotid intima media thickness useful for individual prediction of cardiovascular risk? Ten-year results from the Carotid Atherosclerosis Progression Study (CAPS). Eur Heart J 2010; 31:2041–2048.
257. Willeit P, Tschiderer L, Allara E, Reuber K, Seekircher L, Gao L, et al. Carotid Intima-Media Thickness Progression as Surrogate Marker for Cardiovascular Risk: Meta-Analysis of 119 Clinical Trials Involving 100 667 Patients. Circulation 2020; 142:621–642.
258. Touboul PJ, Hennerici MG, Meairs S, Adams H, Amarenco P, Bornstein N, et al. Mannheim carotid intima-media thickness and plaque consensus (2004-2006-2011). An update on behalf of the advisory board of the 3rd, 4th and 5th watching the risk symposia, at the 13th, 15th and 20th European Stroke Conferences, Mannheim, Germany, 2004, Brussels, Belgium, 2006, and Hamburg, Germany, 2011. Cerebrovasc Dis 2012; 34:290–296.
259. Inaba Y, Chen JA, Bergmann SR. Carotid plaque, compared with carotid intima-media thickness, more accurately predicts coronary artery disease events: a meta-analysis. Atherosclerosis 2012; 220:128–133.
260. Polak JF, Pencina MJ, Pencina KM, O’Donnell CJ, Wolf PA, D’Agostino RB Sr. Carotid-wall intima-media thickness and cardiovascular events. N Engl J Med 2011; 365:213–221.
261. Laurent S, Cockcroft J, Van Bortel L, Boutouyrie P, Giannattasio C, Hayoz D, et al. Expert consensus document on arterial stiffness: methodological issues and clinical applications. Eur Heart J 2006; 27:2588–2605.
262. Benetos A, Adamopoulos C, Bureau JM, Temmar M, Labat C, Bean K, et al. Determinants of accelerated progression of arterial stiffness in normotensive subjects and in treated hypertensive subjects over a 6-year period. Circulation 2002; 105:1202–1207.
263. Agbaje AO, Barker AR, Tuomainen TP. Effects of arterial stiffness and carotid intima-media thickness progression on the risk of overweight/obesity and elevated blood pressure/hypertension: a cross-lagged cohort study. Hypertension (Dallas, Tex: 1979) 2022; 79:159–169.
264. Kaess BM, Rong J, Larson MG, Hamburg NM, Vita JA, Levy D, et al. Aortic stiffness, blood pressure progression, and incident hypertension. JAMA 2012; 308:875–881.
265. Najjar SS, Scuteri A, Shetty V, Wright JG, Muller DC, Fleg JL, et al. Pulse wave velocity is an independent predictor of the longitudinal increase in systolic blood pressure and of incident hypertension in the Baltimore Longitudinal Study of Aging. J Am Coll Cardiol 2008; 51:1377–1383.
266. Van Bortel LM, Laurent S, Boutouyrie P, Chowienczyk P, Cruickshank JK, De Backer T, et al. Expert consensus document on the measurement of aortic stiffness in daily practice using carotid-femoral pulse wave velocity. J Hypertens 2012; 30:445–448.
267. Reference Values for Arterial Stiffness C. Determinants of pulse wave velocity in healthy people and in the presence of cardiovascular risk factors: ’establishing normal and reference values’. Eur Heart J 2010; 31:2338–2350.
268. Yiming G, Zhou X, Lv W, Peng Y, Zhang W, Cheng X, et al. Reference values of brachial-ankle pulse wave velocity according to age and blood pressure in a central Asia population. PLoS One 2017; 12:e0171737.
269. Lu Y, Pechlaner R, Cai J, Yuan H, Huang Z, Yang G, et al. Trajectories of Age-Related Arterial Stiffness in Chinese Men and Women. J Am Coll Cardiol 2020; 75:870–880.
270. Baier D, Teren A, Wirkner K, Loeffler M, Scholz M. Parameters of pulse wave velocity: determinants and reference values assessed in the population-based study LIFE-Adult. Clin Res Cardiol 2018; 107:1050–1061.
271. Spronck B, Heusinkveld MH, Vanmolkot FH, Roodt JO, Hermeling E, Delhaas T, et al. Pressure-dependence of arterial stiffness: potential clinical implications. J Hypertens 2015; 33:330–338.
272. Giannattasio C, Salvi P, Valbusa F, Kearney-Schwartz A, Capra A, Amigoni M, et al. Simultaneous measurement of beat-to-beat carotid diameter and pressure changes to assess arterial mechanical properties. Hypertension (Dallas, Tex: 1979) 2008; 52:896–902.
273. Failla M, Grappiolo A, Emanuelli G, Vitale G, Fraschini N, Bigoni M, et al. Sympathetic tone restrains arterial distensibility of healthy and atherosclerotic subjects. J Hypertens 1999; 17:1117–1123.
274. Mangoni AA, Mircoli L, Giannattasio C, Mancia G, Ferrari AU. Effect of sympathectomy on mechanical properties of common carotid and femoral arteries. Hypertension (Dallas, Tex: 1979) 1997; 30:1085–1088.
275. Gerhard-Herman MD, Gornik HL, Barrett C, Barshes NR, Corriere MA, Drachman DE, et al. 2016 AHA/ACC Guideline on the Management of Patients With Lower Extremity Peripheral Artery Disease: Executive Summary: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines. Circulation 2017; 135:e686–e725.
276. Vasan RS, Song RJ, Xanthakis V, Beiser A, DeCarli C, Mitchell GF, et al. Hypertension-Mediated Organ Damage: Prevalence, Correlates, and Prognosis in the Community. Hypertension (Dallas, Tex: 1979) 2022; 79:505–515.
277. Ohkuma T, Ninomiya T, Tomiyama H, Kario K, Hoshide S, Kita Y, et al. Brachial-Ankle Pulse Wave Velocity and the Risk Prediction of Cardiovascular Disease: An Individual Participant Data Meta-Analysis. Hypertension (Dallas, Tex: 1979) 2017; 69:1045–1052.
278. Ben-Shlomo Y, Spears M, Boustred C, May M, Anderson SG, Benjamin EJ, et al. Aortic pulse wave velocity improves cardiovascular event prediction: an individual participant meta-analysis of prospective observational data from 17,635 subjects. J Am Coll Cardiol 2014; 63:636–646.
279. Koivistoinen T, Lyytikainen LP, Aatola H, Luukkaala T, Juonala M, Viikari J, et al. Pulse Wave Velocity Predicts the Progression of Blood Pressure and Development of Hypertension in Young Adults. Hypertension (Dallas, Tex: 1979) 2018; 71:451–456.
280. Climie RE, Alastruey J, Mayer CC, Schwarz A, Laucyte-Cibulskiene A, Voicehovska J, et al. Vascular ageing: moving from bench towards bedside. Eur J Prev Cardiol 2023; 30:1101–1117.
281. Ait-Oufella H, Collin C, Bozec E, Laloux B, Ong KT, Dufouil C, et al. Long-term reduction in aortic stiffness: a 5.3-year follow-up in routine clinical practice. J Hypertens 2010; 28:2336–2341.
282. Shahin Y, Khan JA, Chetter I. Angiotensin converting enzyme inhibitors effect on arterial stiffness and wave reflections: a meta-analysis and meta-regression of randomised controlled trials. Atherosclerosis 2012; 221:18–33.
283. Edwards NC, Steeds RP, Stewart PM, Ferro CJ, Townend JN. Effect of spironolactone on left ventricular mass and aortic stiffness in early-stage chronic kidney disease: a randomized controlled trial. J Am Coll Cardiol 2009; 54:505–512.
284. Tropeano AI, Boutouyrie P, Pannier B, Joannides R, Balkestein E, Katsahian S, et al. Brachial pressure-independent reduction in carotid stiffness after long-term angiotensin-converting enzyme inhibition in diabetic hypertensives. Hypertension (Dallas, Tex: 1979) 2006; 48:80–86.
285. Laurent S, Chatellier G, Azizi M, Calvet D, Choukroun G, Danchin N, et al. SPARTE Study: Normalization of Arterial Stiffness and Cardiovascular Events in Patients With Hypertension at Medium to Very High Risk. Hypertension (Dallas, Tex: 1979) 2021; 78:983–995.
286. Upadhya B, Pajewski NM, Rocco MV, Hundley WG, Aurigemma G, Hamilton CA, et al. Effect of Intensive Blood Pressure Control on Aortic Stiffness in the SPRINT-HEART. Hypertension (Dallas, Tex: 1979) 2021; 77:1571–1580.
287. Cardoso CRL, Salles GF. Prognostic Value of Changes in Aortic Stiffness for Cardiovascular Outcomes and Mortality in Resistant Hypertension: a Cohort Study. Hypertension (Dallas, Tex: 1979) 2022; 79:447–456.
288. Ankle Brachial Index C, Fowkes FG, Murray GD, Butcher I, Heald CL, Lee RJ, et al. Ankle brachial index combined with Framingham Risk Score to predict cardiovascular events and mortality: a meta-analysis. JAMA 2008; 300:197–208.
289. Summary of Recommendation Statements. Kidney Int Suppl 20112013; 3:5–14.
290. Levey AS, Stevens LA, Schmid CH, Zhang YL, Castro AF 3rd, Feldman HI, et al. A new equation to estimate glomerular filtration rate. Ann Intern Med 2009; 150:604–612.
291. Matsushita K, Mahmoodi BK, Woodward M, Emberson JR, Jafar TH, Jee SH, et al. Comparison of risk prediction using the CKD-EPI equation and the MDRD study equation for estimated glomerular filtration rate. JAMA 2012; 307:1941–1951.
292. Gerstein HC, Mann JF, Yi Q, Zinman B, Dinneen SF, Hoogwerf B, et al. Albuminuria and risk of cardiovascular events, death, and heart failure in diabetic and nondiabetic individuals. JAMA 2001; 286:421–426.
293. Schmieder RE, Schutte R, Schumacher H, Böhm M, Mancia G, Weber MA, et al. Mortality and morbidity in relation to changes in albuminuria, glucose status and systolic blood pressure: an analysis of the ONTARGET and TRANSCEND studies. Diabetologia 2014; 57:2019–2029.
294. Clase CM, Barzilay J, Gao P, Smyth A, Schmieder RE, Tobe S, et al. Acute change in glomerular filtration rate with inhibition of the renin-angiotensin system does not predict subsequent renal and cardiovascular outcomes. Kidney Int 2017; 91:683–690.
295. Schmidt M, Mansfield KE, Bhaskaran K, Nitsch D, Sorensen HT, Smeeth L, et al. Serum creatinine elevation after renin-angiotensin system blockade and long term cardiorenal risks: cohort study. Bmj 2017; 356:j791.
296. McTaggart MP, Newall RG, Hirst JA, Bankhead CR, Lamb EJ, Roberts NW, et al. Diagnostic accuracy of point-of-care tests for detecting albuminuria: a systematic review and meta-analysis. Ann Intern Med 2014; 160:550–557.
297. Ponte B, Pruijm M, Ackermann D, Vuistiner P, Eisenberger U, Guessous I, et al. Reference values and factors associated with renal resistive index in a family-based population study. Hypertension (Dallas, Tex: 1979) 2014; 63:136–142.
298. Andrikou I, Tsioufis C, Konstantinidis D, Kasiakogias A, Dimitriadis K, Leontsinis I, et al. Renal resistive index in hypertensive patients. J Clin Hypertens (Greenwich) 2018; 20:1739–1744.
299. Viazzi F, Grassi G, Pontremoli R. Can we predict outcome by noninvasive assessment of renal haemodynamics in hypertension? The role of renal resistive index. J Hypertens 2016; 34:1047–1049.
300. Scuteri A, Benetos A, Sierra C, Coca A, Chicherio C, Frisoni GB, et al. Routine assessment of cognitive function in older patients with hypertension seen by primary care physicians: why and how-a decision-making support from the working group on ’hypertension and the brain’ of the European Society of Hypertension and from the European Geriatric Medicine Society. J Hypertens 2021; 39:90–100.
301. Triantafyllou A, Ferreira JP, Kobayashi M, Micard E, Xie Y, Kearney-Schwartz A, et al. Longer Duration of Hypertension and MRI Microvascular Brain Alterations Are Associated with Lower Hippocampal Volumes in Older Individuals with Hypertension. J Alzheimers Dis 2020; 74:227–235.
302. Ungvari Z, Toth P, Tarantini S, Prodan CI, Sorond F, Merkely B, et al. Hypertension-induced cognitive impairment: from pathophysiology to public health. Nat Rev Nephrol 2021; 17:639–654.
303. Vermeer SE, Longstreth WT Jr, Koudstaal PJ. Silent brain infarcts: a systematic review. Lancet Neurol 2007; 6:611–619.
304. Iadecola C, Duering M, Hachinski V, Joutel A, Pendlebury ST, Schneider JA, et al. Vascular Cognitive Impairment and Dementia: JACC Scientific Expert Panel. J Am Coll Cardiol 2019; 73:3326–3344.
305. Laurent S, Agabiti-Rosei C, Bruno RM, Rizzoni D. Microcirculation and Macrocirculation in Hypertension: A Dangerous Cross-Link? Hypertension (Dallas, Tex: 1979) 2022; 79:479–490.
306. Mitchell GF, van Buchem MA, Sigurdsson S, Gotal JD, Jonsdottir MK, Kjartansson O, et al. Arterial stiffness, pressure and flow pulsatility and brain structure and function: the Age, Gene/Environment Susceptibility--Reykjavik study. Brain 2011; 134:3398–3407.
307. Webb AJ, Simoni M, Mazzucco S, Kuker W, Schulz U, Rothwell PM. Increased cerebral arterial pulsatility in patients with leukoaraiosis: arterial stiffness enhances transmission of aortic pulsatility. Stroke 2012; 43:2631–2636.
308. Chiesa ST, Masi S, Shipley MJ, Ellins EA, Fraser AG, Hughes AD, et al. Carotid artery wave intensity in mid- to late-life predicts cognitive decline: the Whitehall II study. Eur Heart J 2019; 40:2300–2309.
309. Watfa G, Benetos A, Kearney-Schwartz A, Labat C, Gautier S, Hanon O, et al. Do Arterial Hemodynamic Parameters Predict Cognitive Decline Over a Period of 2 Years in Individuals Older Than 80 Years Living in Nursing Homes? The PARTAGE Study. J Am Med Dir Assoc 2015; 16:598–602.
310. Alvarez-Bueno C, Cunha PG, Martinez-Vizcaino V, Pozuelo-Carrascosa DP, Visier-Alfonso ME, Jimenez-Lopez E, et al. Arterial stiffness and cognition among adults: a systematic review and meta-analysis of observational and longitudinal studies. J Am Heart Assoc 2020; 9:e014621.
311. Zhang B, Huo Y, Yang Z, Lv H, Wang Y, Feng J, et al. Day to Day Blood Pressure Variability Associated With Cerebral Arterial Dilation and White Matter Hyperintensity. Hypertension (Dallas, Tex: 1979) 2022; 79:1455–1465.
312. Li C, Zhu Y, Ma Y, Hua R, Zhong B, Xie W. Association of Cumulative Blood Pressure With Cognitive Decline, Dementia, and Mortality. J Am Coll Cardiol 2022; 79:1321–1335.
313. Breslin DJ, Gifford RW Jr, Fairbairn JF 2nd, Kearns TP. Prognostic importance of ophthalmoscopic findings in essential hypertension. JAMA 1966; 195:335–338.
314. Sairenchi T, Iso H, Yamagishi K, Irie F, Okubo Y, Gunji J, et al. Mild retinopathy is a risk factor for cardiovascular mortality in Japanese with and without hypertension: the Ibaraki Prefectural Health Study. Circulation 2011; 124:2502–2511.
315. Wong TY, Mitchell P. The eye in hypertension. Lancet 2007; 369:425–435.
316. Muiesan ML, Salvetti M, Paini A, Riviera M, Pintossi C, Bertacchini F, et al. Ocular fundus photography with a smartphone device in acute hypertension. J Hypertens 2017; 35:1660–1665.
317. Rizzoni D, Mengozzi A, Masi S, Agabiti Rosei C, De Ciuceis C, Virdis A. New noninvasive methods to evaluate microvascular structure and function. Hypertension (Dallas, Tex: 1979) 2022; 79:874–886.
318. De Ciuceis C, Agabiti Rosei C, Caletti S, Trapletti V, Coschignano MA, Tiberio GAM, et al. Comparison between invasive and noninvasive techniques of evaluation of microvascular structural alterations. J Hypertens 2018; 36:1154–1163.
319. Rizzoni D, Porteri E, Boari GE, De Ciuceis C, Sleiman I, Muiesan ML, et al. Prognostic significance of small-artery structure in hypertension. Circulation 2003; 108:2230–2235.
320. Ben-Shlomo Y, Spears M, Boustred C, May M, Anderson SG, Benjamin EJ, et al. Aortic pulse wave velocity improves cardiovascular event prediction: an individual participant meta-analysis of prospective observational data from 17,635 subjects. J Am Coll Cardiol 2014; 63:636-646.
321. Zanchetti A, Hennig M, Hollweck R, Bond G, Tang R, Cuspidi C, et al. Baseline values but not treatment-induced changes in carotid intima-media thickness predict incident cardiovascular events in treated hypertensive patients: findings in the European Lacidipine Study on Atherosclerosis (ELSA). Circulation 2009; 120:1084–1090.
322. Yeboah J, McClelland RL, Polonsky TS, Burke GL, Sibley CT, O’Leary D, et al. Comparison of novel risk markers for improvement in cardiovascular risk assessment in intermediate-risk individuals. JAMA 2012; 308:788–795.
323. Bang CN, Devereux RB, Okin PM. Regression of electrocardiographic left ventricular hypertrophy or strain is associated with lower incidence of cardiovascular morbidity and mortality in hypertensive patients independent of blood pressure reduction - A LIFE review. J Electrocardiol 2014; 47:630–635.
324. Ibsen H, Olsen MH, Wachtell K, Borch-Johnsen K, Lindholm LH, Mogensen CE, et al. Reduction in albuminuria translates to reduction in cardiovascular events in hypertensive patients: losartan intervention for endpoint reduction in hypertension study. Hypertension (Dallas, Tex: 1979) 2005; 45:198–202.
325. Pascual JM, Rodilla E, Costa JA, Garcia-Escrich M, Gonzalez C, Redon J. Prognostic value of microalbuminuria during antihypertensive treatment in essential hypertension. Hypertension (Dallas, Tex: 1979) 2014; 64:1228–1234.
326. Bakris GL, Sarafidis PA, Weir MR, Dahlof B, Pitt B, Jamerson K, et al. Renal outcomes with different fixed-dose combination therapies in patients with hypertension at high risk for cardiovascular events (ACCOMPLISH): a prespecified secondary analysis of a randomised controlled trial. Lancet 2010; 375:1173–1181.
327. Harrison TG, Tam-Tham H, Hemmelgarn BR, Elliott M, James MT, Ronksley PE, et al. Change in Proteinuria or Albuminuria as a Surrogate for Cardiovascular and Other Major Clinical Outcomes: A Systematic Review and Meta-analysis. Can J Cardiol 2019; 35:77–91.
328. Holtkamp FA, de Zeeuw D, de Graeff PA, Laverman GD, Berl T, Remuzzi G, et al. Albuminuria and blood pressure, independent targets for cardioprotective therapy in patients with diabetes and nephropathy: a post hoc analysis of the combined RENAAL and IDNT trials. Eur Heart J 2011; 32:1493–1499.
329. Inker LA, Levey AS, Pandya K, Stoycheff N, Okparavero A, Greene T, et al. Early change in proteinuria as a surrogate end point for kidney disease progression: an individual patient meta-analysis. Am J Kidney Dis 2014; 64:74–85.
330. Matsushita K, Selvin E, Bash LD, Franceschini N, Astor BC, Coresh J. Change in estimated GFR associates with coronary heart disease and mortality. J Am Soc Nephrol 2009; 20:2617–2624.
331. Chowdhury EK, Langham RG, Ademi Z, Owen A, Krum H, Wing LM, et al. Rate of change in renal function and mortality in elderly treated hypertensive patients. Clin J Am Soc Nephrol 2015; 10:1154–1161.
332. Costanzo P, Perrone-Filardi P, Vassallo E, Paolillo S, Cesarano P, Brevetti G, et al. Does carotid intima-media thickness regression predict reduction of cardiovascular events? A meta-analysis of 41 randomized trials. J Am Coll Cardiol 2010; 56:2006–2020.
333. Guerin AP, Blacher J, Pannier B, Marchais SJ, Safar ME, London GM. Impact of aortic stiffness attenuation on survival of patients in end-stage renal failure. Circulation 2001; 103:987–992.
334. Dorans KS, He H, Chen J, Dobre M, Go AS, Hamm LL, et al. Change in ankle-brachial index and mortality among individuals with chronic kidney disease: findings from the Chronic Renal Insufficiency Cohort Study. Nephrol Dial Transplant 2021; 36:2224–2231.
335. Januszewicz A, Mulatero P, Dobrowolski P, Monticone S, Van der Niepen P, Sarafidis P, et al. Cardiac Phenotypes in Secondary Hypertension: JACC State-of-the-Art Review. J Am Coll Cardiol 2022; 80:1480–1497.
336. Veasey SC, Rosen IM. Obstructive Sleep Apnea in Adults. N Engl J Med 2019; 380:1442–1449.
337. Seravalle G, Grassi G. Sleep Apnea and Hypertension. High Blood Press Cardiovasc Prev 2022; 29:23–31.
338. Mulatero P, Sechi LA, Williams TA, Lenders JWM, Reincke M, Satoh F, et al. Subtype diagnosis, treatment, complications and outcomes of primary aldosteronism and future direction of research: a position statement and consensus of the Working Group on Endocrine Hypertension of the European Society of Hypertension. J Hypertens 2020; 38:1929–1936.
339. Mulatero P, Monticone S, Deinum J, Amar L, Prejbisz A, Zennaro MC, et al. Genetics, prevalence, screening and confirmation of primary aldosteronism: a position statement and consensus of the Working Group on Endocrine Hypertension of The European Society of Hypertension. J Hypertens 2020; 38:1919–1928.
340. Gornik HL, Persu A, Adlam D, Aparicio LS, Azizi M, Boulanger M, et al. First international consensus on the diagnosis and management of fibromuscular dysplasia. J Hypertens 2019; 37:229–252.
341. Fallo F, Di Dalmazi G, Beuschlein F, Biermasz NR, Castinetti F, Elenkova A, et al. Diagnosis and management of hypertension in patients with Cushing's syndrome: a position statement and consensus of the Working Group on Endocrine Hypertension of the European Society of Hypertension. J Hypertens 2022; 40:2085–2101.
342. Ercu M, Markó L, Schächterle C, Tsvetkov D, Cui Y, Maghsodi S, et al. Phosphodiesterase 3A and Arterial Hypertension. Circulation 2020; 142:133–149.
343. Monticone S, Losano I, Tetti M, Buffolo F, Veglio F, Mulatero P. Diagnostic approach to low-renin hypertension. Clin Endocrinol (Oxf) 2018; 89:385–396.
344. Piepoli MF, Hoes AW, Agewall S, Albus C, Brotons C, Catapano AL, et al. 2016 European Guidelines on cardiovascular disease prevention in clinical practice: The Sixth Joint Task Force of the European Society of Cardiology and Other Societies on Cardiovascular Disease Prevention in Clinical Practice (constituted by representatives of 10 societies and by invited experts)Developed with the special contribution of the European Association for Cardiovascular Prevention & Rehabilitation (EACPR). Eur Heart J 2016; 37:2315–2381.
345. Pazoki R, Dehghan A, Evangelou E, Warren H, Gao H, Caulfield M, et al. Genetic Predisposition to High Blood Pressure and Lifestyle Factors: Associations With Midlife Blood Pressure Levels and Cardiovascular Events. Circulation 2018; 137:653–661.
346. Pescatello LS, Wu Y, Gao S, Livingston J, Sheppard BB, Chen MH. Do the combined blood pressure effects of exercise and antihypertensive medications add up to the sum of their parts? A systematic meta-review. BMJ Open Sport Exerc Med 2021; 7:e000895.
347. Filippou CD, Tsioufis CP, Thomopoulos CG, Mihas CC, Dimitriadis KS, Sotiropoulou LI, et al. Dietary Approaches to Stop Hypertension (DASH) Diet and Blood Pressure Reduction in Adults with and without Hypertension: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Adv Nutr 2020; 11:1150–1160.
348. He FJ, Tan M, Ma Y, MacGregor GA. Salt Reduction to Prevent Hypertension and Cardiovascular Disease: JACC State-of-the-Art Review. J Am Coll Cardiol 2020; 75:632–647.
349. Neter JE, Stam BE, Kok FJ, Grobbee DE, Geleijnse JM. Influence of Weight Reduction on Blood Pressure. Hypertension (Dallas, Tex: 1979) 2003; 42:878–884.
350. Blumenthal JA, Babyak MA, Sherwood A, Craighead L, Lin PH, Johnson J, et al. Effects of the dietary approaches to stop hypertension diet alone and in combination with exercise and caloric restriction on insulin sensitivity and lipids. Hypertension (Dallas, Tex: 1979) 2010; 55:1199–1205.
351. Appel LJ, Moore TJ, Obarzanek E, Vollmer WM, Svetkey LP, Sacks FM, et al. A clinical trial of the effects of dietary patterns on blood pressure. DASH Collaborative Research Group. N Engl J Med 1997; 336:1117–1124.
352. Filippini T, Malavolti M, Whelton PK, Naska A, Orsini N, Vinceti M. Blood Pressure Effects of Sodium Reduction: Dose-Response Meta-Analysis of Experimental Studies. Circulation 2021; 143:1542–1567.
353. Binia A, Jaeger J, Hu Y, Singh A, Zimmermann D. Daily potassium intake and sodium-to-potassium ratio in the reduction of blood pressure: a meta-analysis of randomized controlled trials. J Hypertens 2015; 33:1509–1520.
354. Geleijnse JM, Kok FJ, Grobbee DE. Blood pressure response to changes in sodium and potassium intake: a metaregression analysis of randomised trials. J Hum Hypertens 2003; 17:471–480.
355. Filippini T, Naska A, Kasdagli MI, Torres D, Lopes C, Carvalho C, et al. Potassium Intake and Blood Pressure: A Dose-Response Meta-Analysis of Randomized Controlled Trials. J Am Heart Assoc 2020; 9:e015719.
356. Cornelissen VA, Buys R, Smart NA. Endurance exercise beneficially affects ambulatory blood pressure: a systematic review and meta-analysis. J Hypertens 2013; 31:639–648.
357. Cornelissen VA, Smart NA. Exercise training for blood pressure: a systematic review and meta-analysis. J Am Heart Assoc 2013; 2:e004473.
358. Naci H, Salcher-Konrad M, Dias S, Blum MR, Sahoo SA, Nunan D, et al. How does exercise treatment compare with antihypertensive medications? A network meta-analysis of 391 randomised controlled trials assessing exercise and medication effects on systolic blood pressure. Br J Sports Med 2019; 53:859–869.
359. Noone C, Leahy J, Morrissey EC, Newell J, Newell M, Dwyer CP, et al. Comparative efficacy of exercise and anti-hypertensive pharmacological interventions in reducing blood pressure in people with hypertension: A network meta-analysis. Eur J Prev Cardiol 2020; 27:247–255.
360. Roerecke M, Kaczorowski J, Tobe SW, Gmel G, Hasan OSM, Rehm J. The effect of a reduction in alcohol consumption on blood pressure: a systematic review and meta-analysis. Lancet Public Health 2017; 2:e108–e120.
361. Zhong L, Chen W, Wang T, Zeng Q, Lai L, Lai J, et al. Alcohol and Health Outcomes: An Umbrella Review of Meta-Analyses Base on Prospective Cohort Studies. Front Public Health 2022; 10:859947.
362. Huang C, Zhan J, Liu YJ, Li DJ, Wang SQ, He QQ. Association between alcohol consumption and risk of cardiovascular disease and all-cause mortality in patients with hypertension: a meta-analysis of prospective cohort studies. Mayo Clin Proc 2014; 89:1201–1210.
363. Harsha DW, Bray GA. Weight loss and blood pressure control (Pro). Hypertension (Dallas, Tex: 1979) 2008; 51:1420–1425. discussion 1425.
364. Nguyen B, Bauman A, Ding D. Association between lifestyle risk factors and incident hypertension among middle-aged and older Australians. Prev Med 2019; 118:73–80.
365. Gay HC, Rao SG, Vaccarino V, Ali MK. Effects of Different Dietary Interventions on Blood Pressure: Systematic Review and Meta-Analysis of Randomized Controlled Trials. Hypertension (Dallas, Tex: 1979) 2016; 67:733–739.
366. Fu J, Liu Y, Zhang L, Zhou L, Li D, Quan H, et al. Nonpharmacologic Interventions for Reducing Blood Pressure in Adults With Prehypertension to Established Hypertension. J Am Heart Assoc 2020; 9:e016804.
367. Grassi G, Seravalle G, Colombo M, Bolla G, Cattaneo BM, Cavagnini F, et al. Body weight reduction, sympathetic nerve traffic, and arterial baroreflex in obese normotensive humans. Circulation 1998; 97:2037–2042.
368. Kritchevsky SB, Beavers KM, Miller ME, Shea MK, Houston DK, Kitzman DW, et al. Intentional weight loss and all-cause mortality: a meta-analysis of randomized clinical trials. PLoS One 2015; 10:e0121993.
369. Hall ME, Cohen JB, Ard JD, Egan BM, Hall JE, Lavie CJ, et al. Weight-Loss Strategies for Prevention and Treatment of Hypertension: A Scientific Statement From the American Heart Association. Hypertension (Dallas, Tex: 1979) 2021; 78:e38–e50.
370. Chopra S, Malhotra A, Ranjan P, Vikram NK, Sarkar S, Siddhu A, et al. Predictors of successful weight loss outcomes amongst individuals with obesity undergoing lifestyle interventions: A systematic review. Obes Rev 2021; 22:e13148.
371. Rhee EJ. Weight Cycling and Its Cardiometabolic Impact. J Obes Metab Syndr 2017; 26:237–242.
372. Siebenhofer A, Winterholer S, Jeitler K, Horvath K, Berghold A, Krenn C, et al. Long-term effects of weight-reducing drugs in people with hypertension. Cochrane Database Syst Rev 2021; 1:Cd007654.
373. Arterburn DE, Telem DA, Kushner RF, Courcoulas AP. Benefits and Risks of Bariatric Surgery in Adults: A Review. JAMA 2020; 324:879–887.
374. Campbell NRC, Whelton PK, Orias M, Wainford RD, Cappuccio FP, Ide N, et al. 2022 World Hypertension League, Resolve To Save Lives and International Society of Hypertension dietary sodium (salt) global call to action. J Hum Hypertens 2022.
375. GBD Diet Collaborators. Health effects of dietary risks in 195 countries, 1990–2017: a systematic analysis for the Global Burden of Disease Study 2017. Lancet 2019; 393:1958-1972.
376. Mente A, O’Donnell M, Rangarajan S, Dagenais G, Lear S, McQueen M, et al. Associations of urinary sodium excretion with cardiovascular events in individuals with and without hypertension: a pooled analysis of data from four studies. Lancet 2016; 388:465–475.
377. Graudal N, Hubeck-Graudal T, Jürgens G, Taylor RS. Dose-response relation between dietary sodium and blood pressure: a meta-regression analysis of 133 randomized controlled trials. Am J Clin Nutr 2019; 109:1273–1278.
378. Huang L, Trieu K, Yoshimura S, Neal B, Woodward M, Campbell NRC, et al. Effect of dose and duration of reduction in dietary sodium on blood pressure levels: systematic review and meta-analysis of randomised trials. BMJ 2020; 368:m315.
379. Suckling RJ, He FJ, Markandu ND, MacGregor GA. Modest salt reduction lowers blood pressure and albumin excretion in impaired glucose tolerance and type 2 diabetes mellitus: a randomized double-blind trial. Hypertension (Dallas, Tex: 1979) 2016; 67:1189–1195.
380. Pimenta E, Gaddam KK, Oparil S, Aban I, Husain S, Dell’Italia LJ, et al. Effects of dietary sodium reduction on blood pressure in subjects with resistant hypertension: results from a randomized trial. Hypertension (Dallas, Tex: 1979) 2009; 54:475–481.
381. Graudal NA, Hubeck-Graudal T, Jürgens G. Effects of low-sodium diet vs. high-sodium diet on blood pressure, renin, aldosterone, catecholamines, cholesterol, and triglyceride (Cochrane Review). Am J Hypertens 2012; 25:1–15.
382. Cutler JA, Follmann D, Allender PS. Randomized trials of sodium reduction: an overview. Am J Clin Nutr 1997; 65:643s–651s.
383. National Academies of Sciences Engineering and Medicine. Dietary Reference Intakes for Sodium and Potassium. In: Oria, M., Harrison, M., Stallings, V.A., editors. Dietary Reference Intakes for Sodium and Potassium. Washington (DC) 2019.
384. Mancia G, Oparil S, Whelton PK, McKee M, Dominiczak A, Luft FC, et al. The technical report on sodium intake and cardiovascular disease in low- and middle-income countries by the joint working group of the World Heart Federation, the European Society of Hypertension and the European Public Health Association. Eur Heart J 2017; 38:712–719.
385. O’Donnell M, Mente A, Alderman MH, Brady AJB, Diaz R, Gupta R, et al. Salt and cardiovascular disease: insufficient evidence to recommend low sodium intake. Eur Heart J 2020; 41:3363–3373.
386. O’Donnell M, Mente A, Rangarajan S, McQueen MJ, Wang X, Liu L, et al. Urinary sodium and potassium excretion, mortality, and cardiovascular events. N Engl J Med 2014; 371:612–623.
387. Zhu Y, Zhang J, Li Z, Liu Y, Fan X, Zhang Y. Association of sodium intake and major cardiovascular outcomes: a dose-response meta-analysis of prospective cohort studies. BMC Cardiovasc Disord 2018; 18:192.
388. Grassi G, Cattaneo BM, Seravalle G, Lanfranchi A, Bolla G, Mancia G. Baroreflex impairment by low sodium diet in mild or moderate essential hypertension. Hypertension (Dallas, Tex: 1979) 1997; 29:802–807.
389. Göthberg G, Lundin S, Aurell M, Folkow B. Response to slow, graded bleeding in salt-depleted rats. J Hypertens Suppl 1983; 1:24–26.
390. Campbell NRC, Whelton PK, Orias M, Cobb LL, Jones ESW, Garg R, et al. It is strongly recommended to not conduct, fund, or publish research studies that use spot urine samples with estimating equations to assess individuals’ sodium (salt) intake in association with health outcomes: a policy statement of the World Hypertension League, International Society of Hypertension and Resolve to Save Lives. J Hypertens 2023; 41:683–686.
391. He FJ, Campbell NRC, Woodward M, MacGregor GA. Salt reduction to prevent hypertension: the reasons of the controversy. Eur Heart J 2021; 42:2501–2505.
392. Neal B, Wu Y, Feng X, Zhang R, Zhang Y, Shi J, et al. Effect of Salt Substitution on Cardiovascular Events and Death. N Engl J Med 2021; 385:1067–1077.
393. Tsai YC, Tsao YP, Huang CJ, Tai YH, Su YC, Chiang CE, et al. Effectiveness of salt substitute on cardiovascular outcomes: A systematic review and meta-analysis. J Clin Hypertens (Greenwich) 2022; 24:1147–1160.
394. Poorolajal J, Zeraati F, Soltanian AR, Sheikh V, Hooshmand E, Maleki A. Oral potassium supplementation for management of essential hypertension: A meta-analysis of randomized controlled trials. PLoS One 2017; 12:e0174967.
395. Aburto NJ, Hanson S, Gutierrez H, Hooper L, Elliott P, Cappuccio FP. Effect of increased potassium intake on cardiovascular risk factors and disease: systematic review and meta-analyses. BMJ 2013; 346:f1378.
396. Liu X, Zhang D, Liu Y, Sun X, Han C, Wang B, et al. Dose-Response Association Between Physical Activity and Incident Hypertension: A Systematic Review and Meta-Analysis of Cohort Studies. Hypertension (Dallas, Tex: 1979) 2017; 69:813–820.
397. Rijal A, Nielsen EE, Adhikari TB, Dhakal S, Maagaard M, Piri R, et al. Effects of adding exercise to usual care in patients with either hypertension, type 2 diabetes or cardiovascular disease: a systematic review with meta-analysis and trial sequential analysis. Br J Sports Med 2023; 57:930–939.
398. Cornelissen VA, Fagard RH. Effects of endurance training on blood pressure, blood pressure-regulating mechanisms, and cardiovascular risk factors. Hypertension (Dallas, Tex: 1979) 2005; 46:667–675.
399. Whelton SP, Chin A, Xin X, He J. Effect of aerobic exercise on blood pressure: a meta-analysis of randomized, controlled trials. Ann Intern Med 2002; 136:493–503.
400. Saco-Ledo G, Valenzuela PL, Ruiz-Hurtado G, Ruilope LM, Lucia A. Exercise Reduces Ambulatory Blood Pressure in Patients With Hypertension: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. J Am Heart Assoc 2020; 9:e018487.
401. Cornelissen VA, Fagard RH, Coeckelberghs E, Vanhees L. Impact of resistance training on blood pressure and other cardiovascular risk factors: a meta-analysis of randomized, controlled trials. Hypertension (Dallas, Tex: 1979) 2011; 58:950–958.
402. Lopez-Valenciano A, Ruiz-Perez I, Ayala F, Sanchez-Meca J, Vera-Garcia FJ. Updated systematic review and meta-analysis on the role of isometric resistance training for resting blood pressure management in adults. J Hypertens 2019; 37:1320–1333.
403. Smart NA, Way D, Carlson D, Millar P, McGowan C, Swaine I, et al. Effects of isometric resistance training on resting blood pressure: individual participant data meta-analysis. J Hypertens 2019; 37:1927–1938.
404. Cohen DD, Aroca-Martinez G, Carreño-Robayo J, Castañeda-Hernández A, Herazo-Beltran Y, Camacho PA, et al. Reductions in systolic blood pressure achieved by hypertensives with three isometric training sessions per week are maintained with a single session per week. J Clin Hypertens (Greenwich) 2023; 25:380–387.
405. Williamson W, Foster C, Reid H, Kelly P, Lewandowski AJ, Boardman H, et al. Will exercise advice be sufficient for treatment of young adults with prehypertension and hypertension? a systematic review and meta-analysis. Hypertension (Dallas, Tex: 1979) 2016; 68:78–87.
406. Igarashi Y, Akazawa N, Maeda S. Regular aerobic exercise and blood pressure in East Asians: A meta-analysis of randomized controlled trials. Clin Exp Hypertens 2018; 40:378–389.
407. Kelley GA, Kelley KS. Aerobic exercise and resting blood pressure in women: a meta-analytic review of controlled clinical trials. J Womens Health Gend Based Med 1999; 8:787–803.
408. Bersaoui M, Baldew SM, Cornelis N, Toelsie J, Cornelissen VA. The effect of exercise training on blood pressure in African and Asian populations: A systematic review and meta-analysis of randomized controlled trials. Eur J Prev Cardiol 2020; 27:457–472.
409. Hansen D, Abreu A, Ambrosetti M, Cornelissen V, Gevaert A, Kemps H, et al. Exercise intensity assessment and prescription in cardiovascular rehabilitation and beyond: why and how: a position statement from the Secondary Prevention and Rehabilitation Section of the European Association of Preventive Cardiology. Eur J Prev Cardiol 2022; 29:230–245.
410. Lee LL, Mulvaney CA, Wong YKY, Chan ES, Watson MC, Lin HH. Walking for hypertension. Cochrane Database Syst Rev 2021; 2:CD008823.
411. Casonatto J, Goessler KF, Cornelissen VA, Cardoso JR, Polito MD. The blood pressure-lowering effect of a single bout of resistance exercise: A systematic review and meta-analysis of randomised controlled trials. Eur J Prev Cardiol 2016; 23:1700–1714.
412. Kenney MJ, Seals DR. Postexercise hypotension. Key features, mechanisms, and clinical significance. Hypertension (Dallas, Tex: 1979) 1993; 22:653–664.
413. Marçal IR, Goessler KF, Buys R, Casonatto J, Ciolac EG, Cornelissen VA. Post-exercise Hypotension Following a Single Bout of High Intensity Interval Exercise vs. a Single Bout of Moderate Intensity Continuous Exercise in Adults With or Without Hypertension: A Systematic Review and Meta-Analysis of Randomized Clinical Trials. Front Physiol 2021; 12:675289.
414. Dempsey PC, Sacre JW, Larsen RN, Straznicky NE, Sethi P, Cohen ND, et al. Interrupting prolonged sitting with brief bouts of light walking or simple resistance activities reduces resting blood pressure and plasma noradrenaline in type 2 diabetes. J Hypertens 2016; 34:2376–2382.
415. Warburton DER, Bredin SSD. Health benefits of physical activity: a systematic review of current systematic reviews. Curr Opin Cardiol 2017; 32:541–556.
416. Ekelund U, Tarp J, Steene-Johannessen J, Hansen BH, Jefferis B, Fagerland MW, et al. Dose-response associations between accelerometry measured physical activity and sedentary time and all cause mortality: systematic review and harmonised meta-analysis. BMJ 2019; 366:l4570.
417. Kraus WE, Powell KE, Haskell WL, Janz KF, Campbell WW, Jakicic JM, et al. Physical Activity, All-Cause and Cardiovascular Mortality, and Cardiovascular Disease. Med Sci Sports Exerc 2019; 51:1270–1281.
418. Sattelmair J, Pertman J, Ding EL, Kohl HW, Haskell W, Lee IM. Dose response between physical activity and risk of coronary heart disease: a meta-analysis. Circulation 2011; 124:789–795.
419. Joseph G, Marott JL, Torp-Pedersen C, Biering-Sørensen T, Nielsen G, Christensen AE, et al. Dose-Response Association Between Level of Physical Activity and Mortality in Normal, Elevated, and High Blood Pressure. Hypertension (Dallas, Tex: 1979) 2019; 74:1307–1315.
420. Abramson JL, Lewis C, Murrah NV. Relationship of self-reported alcohol consumption to ambulatory blood pressure in a sample of healthy adults. Am J Hypertens 2010; 23:994–999.
421. Puddey IB, Beilin LJ, Vandongen R, Rouse IL, Rogers P. Evidence for a direct effect of alcohol consumption on blood pressure in normotensive men. A randomized controlled trial. Hypertension (Dallas, Tex: 1979) 1985; 7:707–713.
422. Global burden of 87 risk factors in 204 countries and territories, 1990–2019: a systematic analysis for the Global Burden of Disease Study 2019. Lancet 2020; 396:1223-1249.
423. Xi B, Veeranki SP, Zhao M, Ma C, Yan Y, Mi J. Relationship of Alcohol Consumption to All-Cause, Cardiovascular, and Cancer-Related Mortality in U.S. Adults. J Am Coll Cardiol 2017; 70:913–922.
424. Mukamal KJ, Chen CM, Rao SR, Breslow RA. Alcohol consumption and cardiovascular mortality among U.S. adults, 1987 to 2002. J Am Coll Cardiol 2010; 55:1328–1335.
425. Biddinger KJ, Emdin CA, Haas ME, Wang M, Hindy G, Ellinor PT, et al. Association of Habitual Alcohol Intake With Risk of Cardiovascular Disease. JAMA Netw Open 2022; 5:e223849.
426. Hu C, Huang C, Li J, Liu F, Huang K, Liu Z, et al. Causal associations of alcohol consumption with cardiovascular diseases and all-cause mortality among Chinese males. Am J Clin Nutr 2022; 116:771–779.
427. Roerecke M, Tobe SW, Kaczorowski J, Bacon SL, Vafaei A, Hasan OSM, et al. Sex-Specific Associations Between Alcohol Consumption and Incidence of Hypertension: A Systematic Review and Meta-Analysis of Cohort Studies. J Am Heart Assoc 2018; 7:
428. Seppä K, Sillanaukee P. Binge drinking and ambulatory blood pressure. Hypertension (Dallas, Tex: 1979) 1999; 33:79–82.
429. Ariesen MJ, Claus SP, Rinkel GJ, Algra A. Risk factors for intracerebral hemorrhage in the general population: a systematic review. Stroke 2003; 34:2060–2065.
430. Anderson BO, Berdzuli N, Ilbawi A, Kestel D, Kluge HP, Krech R, et al. Health and cancer risks associated with low levels of alcohol consumption. Lancet Public Health 2023; 8:e6–e7.
431. Saha SP, Bhalla DK, Whayne TF, Gairola C. Cigarette smoke and adverse health effects: An overview of research trends and future needs. Int J Angiol 2007; 16:77–83.
432. National Center for Chronic Disease Prevention and Health Promotion (US) Office on Smoking and Health. The Health Consequences of Smoking—50 Years of Progress: A Report of the Surgeon General. The Health Consequences of Smoking-50 Years of Progress: A Report of the Surgeon General. Atlanta (GA) 2014.
433. Groppelli A, Giorgi DM, Omboni S, Parati G, Mancia G. Persistent blood pressure increase induced by heavy smoking. J Hypertens 1992; 10:495–499.
434. Grassi G, Seravalle G, Calhoun DA, Bolla GB, Giannattasio C, Marabini M, et al. Mechanisms responsible for sympathetic activation by cigarette smoking in humans. Circulation 1994; 90:248–253.
435. Omvik P. How smoking affects blood pressure. Blood Press 1996; 5:71–77.
436. Mahmud A, Feely J. Effects of passive smoking on blood pressure and aortic pressure waveform in healthy young adults--influence of gender. Br J Clin Pharmacol 2004; 57:37–43.
437. Cheung YTD, Jiang N, Jiang CQ, Zhuang RS, Gao WH, Zhou J, et al. Physicians’ very brief (30-sec) intervention for smoking cessation on 13 671 smokers in China: a pragmatic randomized controlled trial. Addiction 2021; 116:1172–1185.
438. Hartmann-Boyce J, Livingstone-Banks J, Ordóñez-Mena JM, Fanshawe TR, Lindson N, Freeman SC, et al. Behavioural interventions for smoking cessation: an overview and network meta-analysis. Cochrane Database Syst Rev 2021; 1:CD013229.
439. Maziak W, Taleb ZB, Bahelah R, Islam F, Jaber R, Auf R, et al. The global epidemiology of waterpipe smoking. Tob Control 2015; 24: (Suppl 1): i3–i12.
440. Larue F, Tasbih T, Ribeiro PAB, Lavoie KL, Dolan E, Bacon SL. Immediate physiological effects of acute electronic cigarette use in humans: A systematic review and meta-analysis. Respir Med 2021; 190:106684.
441. Kennedy CD, van Schalkwyk MCI, McKee M, Pisinger C. The cardiovascular effects of electronic cigarettes: A systematic review of experimental studies. Prev Med 2019; 127:105770.
442. Skotsimara G, Antonopoulos AS, Oikonomou E, Siasos G, Ioakeimidis N, Tsalamandris S, et al. Cardiovascular effects of electronic cigarettes: A systematic review and meta-analysis. Eur J Prev Cardiol 2019; 26:1219–1228.
443. Al Ali R, Vukadinović D, Maziak W, Katmeh L, Schwarz V, Mahfoud F, et al. Cardiovascular effects of waterpipe smoking: a systematic review and meta-analysis. Rev Cardiovasc Med 2020; 21:453–468.
444. Saffar Soflaei S, Darroudi S, Tayefi M, Nosrati Tirkani A, Moohebati M, Ebrahimi M, et al. Hookah smoking is strongly associated with diabetes mellitus, metabolic syndrome and obesity: a population-based study. Diabetol Metab Syndr 2018; 10:33.
445. Toledo E, Hu FB, Estruch R, Buil-Cosiales P, Corella D, Salas-Salvadó J, et al. Effect of the Mediterranean diet on blood pressure in the PREDIMED trial: results from a randomized controlled trial. BMC Med 2013; 11:207.
446. Doménech M, Roman P, Lapetra J, García de la Corte FJ, Sala-Vila A, de la Torre R, et al. Mediterranean diet reduces 24-hour ambulatory blood pressure, blood glucose, and lipids: one-year randomized, clinical trial. Hypertension (Dallas, Tex: 1979) 2014; 64:69–76.
447. Filippou CD, Thomopoulos CG, Kouremeti MM, Sotiropoulou LI, Nihoyannopoulos PI, Tousoulis DM, et al. Mediterranean diet and blood pressure reduction in adults with and without hypertension: A systematic review and meta-analysis of randomized controlled trials. Clin Nutr 2021; 40:3191–3200.
448. Soltani S, Arablou T, Jayedi A, Salehi-Abargouei A. Adherence to the dietary approaches to stop hypertension (DASH) diet in relation to all-cause and cause-specific mortality: a systematic review and dose-response meta-analysis of prospective cohort studies. Nutr J 2020; 19:37.
449. Sofi F, Abbate R, Gensini GF, Casini A. Accruing evidence on benefits of adherence to the Mediterranean diet on health: an updated systematic review and meta-analysis. Am J Clin Nutr 2010; 92:1189–1196.
450. Cicero AFG, Veronesi M, Fogacci F. Dietary Intervention to Improve Blood Pressure Control: Beyond Salt Restriction. High Blood Press Cardiovasc Prev 2021; 28:547–553.
451. Dinu M, Pagliai G, Angelino D, Rosi A, Dall’Asta M, Bresciani L, et al. Effects of Popular Diets on Anthropometric and Cardiometabolic Parameters: An Umbrella Review of Meta-Analyses of Randomized Controlled Trials. Adv Nutr 2020; 11:815–833.
452. Borghi C. Coffee and blood pressure: exciting news!. Blood Press 2022; 31:284–287.
453. Marcus GM, Rosenthal DG, Nah G, Vittinghoff E, Fang C, Ogomori K, et al. Acute Effects of Coffee Consumption on Health among Ambulatory Adults. N Engl J Med 2023; 388:1092–1100.
454. Guessous I, Dobrinas M, Kutalik Z, Pruijm M, Ehret G, Maillard M, et al. Caffeine intake and CYP1A2 variants associated with high caffeine intake protect non-smokers from hypertension. Hum Mol Genet 2012; 21:3283–3292.
455. Albus C, Waller C, Fritzsche K, Gunold H, Haass M, Hamann B, et al. Significance of psychosocial factors in cardiology: update 2018: Position paper of the German Cardiac Society. Clin Res Cardiol 2019; 108:1175–1196.
456. Liu M-Y, Li N, Li WA, Khan H. Association between psychosocial stress and hypertension: a systematic review and meta-analysis. Neurological Research 2017; 39:573–580.
457. Gaffey AE, Rosman L, Sico JJ, Haskell SG, Brandt CA, Bathulapalli H, et al. Military sexual trauma and incident hypertension: a 16-year cohort study of young and middle-aged men and women. J Hypertens 2022; 40:2307–2315.
458. Mendlowicz V, Garcia-Rosa ML, Gekker M, Wermelinger L, Berger W, Luz MP, et al. Post-traumatic stress disorder as a predictor for incident hypertension: a 3-year retrospective cohort study. Psychol Med 2021; 1–8.
459. Dyball D, Evans S, Boos CJ, Stevelink SAM, Fear NT. The association between PTSD and cardiovascular disease and its risk factors in male veterans of the Iraq/Afghanistan conflicts: a systematic review. Int Rev Psychiatry 2019; 31:34–48.
460. Conversano C, Orrù G, Pozza A, Miccoli M, Ciacchini R, Marchi L, et al. Is mindfulness-based stress reduction effective for people with hypertension? a systematic review and meta-analysis of 30 years of Evidence. Int J Environ Res Public Health 2021; 18:
461. Shi L, Zhang D, Wang L, Zhuang J, Cook R, Chen L. Meditation and blood pressure: a meta-analysis of randomized clinical trials. J Hypertens 2017; 35:696–706.
462. Schneider JK, Reangsing C, Willis DG. Effects of Transcendental Meditation on Blood Pressure: A Meta-analysis. J Cardiovasc Nurs 2022; 37:E11–E21.
463. Hahad O, Rajagopalan S, Lelieveld J, Sorensen M, Kuntic M, Daiber A, et al. Noise and Air Pollution as Risk Factors for Hypertension: Part II-Pathophysiologic Insight. Hypertension (Dallas, Tex: 1979) 2023; 80:1384–1392.
464. Wojciechowska W, Januszewicz A, Drożdż T, Rojek M, Bączalska J, Terlecki M, et al. Blood pressure and arterial stiffness in association with aircraft noise exposure: long-term observation and potential effect of COVID-19 lockdown. Hypertension (Dallas, Tex: 1979) 2022; 79:325–334.
465. Prabhakaran D, Anand S, Watkins D, Gaziano T, Wu Y, Mbanya JC, et al. Cardiovascular, respiratory, and related disorders: key messages from Disease Control Priorities. 3rd edition. Lancet 2018; 391:1224–1236.
466. Ettehad D, Emdin CA, Kiran A, Anderson SG, Callender T, Emberson J, et al. Blood pressure lowering for prevention of cardiovascular disease and death: a systematic review and meta-analysis. Lancet 2016; 387:957–967.
467. Tsai WC, Wu HY, Peng YS, Yang JY, Chen HY, Chiu YL, et al. Association of Intensive Blood Pressure Control and Kidney Disease Progression in Nondiabetic Patients With Chronic Kidney Disease: A Systematic Review and Meta-analysis. JAMA Intern Med 2017; 177:792–799.
468. Xie X, Atkins E, Lv J, Bennett A, Neal B, Ninomiya T, et al. Effects of intensive blood pressure lowering on cardiovascular and renal outcomes: updated systematic review and meta-analysis. Lancet 2016; 387:435–443.
469. Agabiti Rosei E, Muiesan, M.L. Evaluation of Cardiac Damage in Hypertension: Echocardiography. In: Mancia EA-RaG, editor. Asessment of Prclinical Organ Damage in Hypertension: Springer; 2015. pp. 13-25.
470. White WB, Wakefield DB, Moscufo N, Guttmann CRG, Kaplan RF, Bohannon RW, et al. Effects of Intensive Versus Standard Ambulatory Blood Pressure Control on Cerebrovascular Outcomes in Older People (INFINITY). Circulation 2019; 140:1626–1635.
471. Williamson JD, Pajewski NM, Auchus AP, Bryan RN, Chelune G, Cheung AK, et al. Effect of Intensive vs Standard Blood Pressure Control on Probable Dementia: A Randomized Clinical Trial. JAMA 2019; 321:553–561.
472. Brunstrom M, Carlberg B. Association of Blood Pressure Lowering With Mortality and Cardiovascular Disease Across Blood Pressure Levels: A Systematic Review and Meta-analysis. JAMA Intern Med 2018; 178:28–36.
473. Collins R, MacMahon S. Blood pressure, antihypertensive drug treatment and the risks of stroke and of coronary heart disease. Br Med Bull 1994; 50:272–298.
474. Abegaz TM, Shehab A, Gebreyohannes EA, Bhagavathula AS, Elnour AA. Nonadherence to antihypertensive drugs: A systematic review and meta-analysis. Medicine (Baltimore) 2017; 96:e5641.
475. Kulkarni S, Rao R, Goodman JDH, Connolly K, O'Shaughnessy KM. Nonadherence to antihypertensive medications amongst patients with uncontrolled hypertension: A retrospective study. Medicine (Baltimore) 2021; 100:e24654.
476. Naderi SH, Bestwick JP, Wald DS. Adherence to drugs that prevent cardiovascular disease: meta-analysis on 376,162 patients. Am J Med 2012; 125:882–887. e881.
477. Corrao G, Mancia G. Research strategies in treatment of hypertension: value of retrospective real-life data. Eur Heart J 2022; 43:3312–3322.
478. Corrao G, Parodi A, Nicotra F, Zambon A, Merlino L, Cesana G, et al. Better compliance to antihypertensive medications reduces cardiovascular risk. J Hypertens 2011; 29:610–618.
479. Burnier M, Egan BM. Adherence in Hypertension. Circ Res 2019; 124:1124–1140.
480. Derington CG, King JB, Bryant KB, McGee BT, Moran AE, Weintraub WS, et al. Cost-Effectiveness and Challenges of Implementing Intensive Blood Pressure Goals and Team-Based Care. Curr Hypertens Rep 2019; 21:91.
481. Health NIf, Excellence C. Hypertension in adults: diagnosis and management. Clinical guideline CG127 National Clininal Guideline Centre London 2011.
482. Moise N, Huang C, Rodgers A, Kohli-Lynch CN, Tzong KY, Coxson PG, et al. Comparative cost-effectiveness of conservative or intensive blood pressure treatment guidelines in adults aged 35-74 years: the cardiovascular disease policy model. Hypertension (Dallas, Tex: 1979) 2016; 68:88–96.
483. Moran AE, Odden MC, Thanataveerat A, Tzong KY, Rasmussen PW, Guzman D, et al. Cost-effectiveness of hypertension therapy according to 2014 guidelines. N Engl J Med 2015; 372:447–455.
484. Blood Pressure Lowering Treatment Trialists C. Pharmacological blood pressure lowering for primary and secondary prevention of cardiovascular disease across different levels of blood pressure: an individual participant-level data meta-analysis. Lancet 2021; 397:1625–1636.
485. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure lowering on outcome incidence in hypertension: 3. Effects in patients at different levels of cardiovascular risk--overview and meta-analyses of randomized trials. J Hypertens 2014; 32:2305–2314.
486. Diao D, Wright JM, Cundiff DK, Gueyffier F. Pharmacotherapy for mild hypertension. Cochrane Database Syst Rev 2012; 2012:Cd006742.
487. Kjeldsen S, Feldman RD, Lisheng L, Mourad JJ, Chiang CE, Zhang W, et al. Updated national and international hypertension guidelines: a review of current recommendations. Drugs 2014; 74:2033–2051.
488. Whelton PK, Carey RM, Aronow WS, Casey DE Jr, Collins KJ, Dennison Himmelfarb C, et al. 2017 ACC/AHA/AAPA/ABC/ACPM/AGS/APhA/ASH/ASPC/NMA/PCNA Guideline for the Prevention, Detection, Evaluation, and Management of High Blood Pressure in Adults: Executive Summary: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines. Hypertension (Dallas, Tex: 1979) 2018; 71:1269–1324.
489. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure-lowering treatment on cardiovascular outcomes and mortality: 13 - benefits and adverse events in older and younger patients with hypertension: overview, meta-analyses and meta-regression analyses of randomized trials. J Hypertens 2018; 36:1622–1636.
490. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure lowering on outcome incidence in hypertension: 2. Effects at different baseline and achieved blood pressure levels--overview and meta-analyses of randomized trials. J Hypertens 2014; 32:2296–2304.
491. Sundstrom J, Arima H, Jackson R, Turnbull F, Rahimi K, Chalmers J, et al. Effects of blood pressure reduction in mild hypertension: a systematic review and meta-analysis. Ann Intern Med 2015; 162:184–191.
492. Lonn EM, Bosch J, Lopez-Jaramillo P, Zhu J, Liu L, Pais P, et al. Blood-Pressure Lowering in Intermediate-Risk Persons without Cardiovascular Disease. N Engl J Med 2016; 374:2009–2020.
493. Thomopoulos C, Parati G, Zanchetti A. Effects of blood-pressure-lowering treatment on outcome incidence. 12. Effects in individuals with high-normal and normal blood pressure: overview and meta-analyses of randomized trials. J Hypertens 2017; 35:2150–2160.
494. Hirakawa Y, Arima H, Webster R, Zoungas S, Li Q, Harrap S, et al. Risks associated with permanent discontinuation of blood pressure-lowering medications in patients with type 2 diabetes. J Hypertens 2016; 34:781–787.
495. Carlberg B. What do we know about the risks of stopping antihypertensive treatment? J Hypertens 2014; 32:1400–1401.
496. Zhang W, Zhang S, Deng Y, Wu S, Ren J, Sun G, et al. Trial of Intensive Blood-Pressure Control in Older Patients with Hypertension. N Engl J Med 2021; 385:1268–1279.
497. Perry HM Jr, Davis BR, Price TR, Applegate WB, Fields WS, Guralnik JM, et al. Effect of treating isolated systolic hypertension on the risk of developing various types and subtypes of stroke: the Systolic Hypertension in the Elderly Program (SHEP). JAMA 2000; 284:465–471.
498. Staessen JA, Fagard R, Thijs L, Celis H, Arabidze GG, Birkenhäger WH, et al. Randomised double-blind comparison of placebo and active treatment for older patients with isolated systolic hypertension. The Systolic Hypertension in Europe (Syst-Eur) Trial Investigators. Lancet 1997; 350:757–764.
499. Roush GC, Zubair A, Singh K, Kostis WJ, Sica DA, Kostis JB. Does the benefit from treating to lower blood pressure targets vary with age? A systematic review and meta-analysis. J Hypertens 2019; 37:1558–1566.
500. Brunström M, Carlberg B, Kjeldsen SE. Effect of antihypertensive treatment in isolated systolic hypertension (ISH) - systematic review and meta-analysis of randomised controlled trials. Blood Press 2023; 32:2226757.
501. Brunström M, Carlberg B, Kjeldsen SE. Effect of antihypertensive treatment in isolated systolic hypertension (ISH) – systematic review and meta-analysis of randomized controlled trials. Blood Press 2023; In press.
502. Beckett NS, Peters R, Fletcher AE, Staessen JA, Liu L, Dumitrascu D, et al. Treatment of hypertension in patients 80 years of age or older. N Engl J Med 2008; 358:1887–1898.
503. Williamson JD, Supiano MA, Applegate WB, Berlowitz DR, Campbell RC, Chertow GM, et al. Intensive vs Standard Blood Pressure Control and Cardiovascular Disease Outcomes in Adults Aged ≥75 Years: A Randomized Clinical Trial. Jama 2016; 315:2673–2682.
504. Odden MC, Peralta CA, Haan MN, Covinsky KE. Rethinking the association of high blood pressure with mortality in elderly adults: the impact of frailty. Arch Intern Med 2012; 172:1162–1168.
505. Kremer KM, Braisch U, Rothenbacher D, Denkinger M, Dallmeier D. Systolic blood pressure and mortality in community-dwelling older adults: frailty as an effect modifier. Hypertension (Dallas, Tex: 1979) 2022; 79:24–32.
506. Odden MC, Anderson TS. How low should you go in the presence of frailty? Hypertension (Dallas, Tex: 1979) 2022; 79:33–35.
507. Aparicio LS, Thijs L, Boggia J, Jacobs L, Barochiner J, Odili AN, et al. Defining thresholds for home blood pressure monitoring in octogenarians. Hypertension (Dallas, Tex: 1979) 2015; 66:865–873.
508. Benetos A, Labat C, Rossignol P, Fay R, Rolland Y, Valbusa F, et al. Treatment With Multiple Blood Pressure Medications, Achieved Blood Pressure, and Mortality in Older Nursing Home Residents: The PARTAGE Study. JAMA Intern Med 2015; 175:989–995.
509. Mossello E, Pieraccioli M, Nesti N, Bulgaresi M, Lorenzi C, Caleri V, et al. Effects of low blood pressure in cognitively impaired elderly patients treated with antihypertensive drugs. JAMA Intern Med 2015; 175:578–585.
510. Streit S, Poortvliet RKE, Gussekloo J. Lower blood pressure during antihypertensive treatment is associated with higher all-cause mortality and accelerated cognitive decline in the oldest-old. Data from the Leiden 85-plus Study. Age Ageing 2018; 47:545–550.
511. Douros A, Tölle M, Ebert N, Gaedeke J, Huscher D, Kreutz R, et al. Control of blood pressure and risk of mortality in a cohort of older adults: the Berlin Initiative Study. Eur Heart J 2019; 40:2021–2028.
512. Rea F, Cantarutti A, Merlino L, Ungar A, Corrao G, Mancia G. Antihypertensive treatment in elderly frail patients: evidence from a large italian database. Hypertension (Dallas, Tex: 1979) 2020; 76:442–449.
513. Rea F, Mancia G, Corrao G. Statin treatment reduces the risk of death among elderly frail patients: evidence from a large population-based cohort. Eur J Prev Cardiol 2022; 28:1885–1894.
514. Corrao G, Zambon A, Parodi A, Poluzzi E, Baldi I, Merlino L, et al. Discontinuation of and changes in drug therapy for hypertension among newly-treated patients: a population-based study in Italy. J Hypertens 2008; 26:819–824.
515. Arguedas JA, Leiva V, Wright JM. Blood pressure targets in adults with hypertension. Cochrane Database Syst Rev 2020; 12:Cd004349.
516. Salvetti M, Paini A, Aggiusti C, Bertacchini F, Stassaldi D, Capellini S, et al. Unattended versus attended blood pressure measurement. Hypertension (Dallas, Tex: 1979) 2019; 73:736–742.
517. Saiz LC, Gorricho J, Garjon J, Celaya MC, Erviti J, Leache L. Blood pressure targets for the treatment of people with hypertension and cardiovascular disease. Cochrane Database Syst Rev 2020; 9:CD010315.
518. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure lowering on outcome incidence in hypertension: 7. Effects of more vs. less intensive blood pressure lowering and different achieved blood pressure levels - updated overview and meta-analyses of randomized trials. J Hypertens 2016; 34:613–622.
519. Schrader J, Lüders S, Züchner C, Herbold M, Schrandt G. Practice vs ambulatory blood pressure measurement under treatment with ramipril (PLUR Study): a randomised, prospective long-term study to evaluate the benefits of ABPM in patients on antihypertensive treatment. J Hum Hypertens 2000; 14:435–440.
520. Mancia G, Zanchetti A, Agabiti-Rosei E, Benemio G, De Cesaris R, Fogari R, et al. Ambulatory blood pressure is superior to clinic blood pressure in predicting treatment-induced regression of left ventricular hypertrophy. SAMPLE Study Group. Study on Ambulatory Monitoring of Blood Pressure and Lisinopril Evaluation. Circulation 1997; 95:1464–1470.
521. Mortensen RN, Gerds TA, Jeppesen JL, Torp-Pedersen C. Office blood pressure or ambulatory blood pressure for the prediction of cardiovascular events. Eur Heart J 2017; 38:3296–3304.
522. Mancia G, Parati G, Bilo G, Gao P, Fagard R, Redon J, et al. Ambulatory blood pressure values in the Ongoing Telmisartan Alone and in Combination with Ramipril Global Endpoint Trial (ONTARGET). Hypertension (Dallas, Tex: 1979) 2012; 60:1400–1406.
523. Weber MA, Julius S, Kjeldsen SE, Brunner HR, Ekman S, Hansson L, et al. Blood pressure dependent and independent effects of antihypertensive treatment on clinical events in the VALUE Trial. Lancet 2004; 363:2049–2051.
524. Julius S, Kjeldsen SE, Weber M, Brunner HR, Ekman S, Hansson L, et al. Outcomes in hypertensive patients at high cardiovascular risk treated with regimens based on valsartan or amlodipine: the VALUE randomised trial. Lancet 2004; 363:2022–2031.
525. Doumas M, Tsioufis C, Fletcher R, Amdur R, Faselis C, Papademetriou V. Time in therapeutic range, as a determinant of all-cause mortality in patients with hypertension. J Am Heart Assoc 2017; 6:
526. Andersson OK, Almgren T, Persson B, Samuelsson O, Hedner T, Wilhelmsen L. Survival in treated hypertension: follow up study after two decades. BMJ 1998; 317:167–171.
527. Blacher J, Evans A, Arveiler D, Amouyel P, Ferrières J, Bingham A, et al. Residual cardiovascular risk in treated hypertension and hyperlipidaemia: the PRIME Study. J Hum Hypertens 2010; 24:19–26.
528. Zhou D, Xi B, Zhao M, Wang L, Veeranki SP. Uncontrolled hypertension increases risk of all-cause and cardiovascular disease mortality in US adults: the NHANES III Linked Mortality Study. Sci Rep 2018; 8:9418.
529. Zanchetti A. Bottom blood pressure or bottom cardiovascular risk? How far can cardiovascular risk be reduced? J Hypertens 2009; 27:1509–1520.
530. Bhatt DL, Steg PG, Ohman EM, Hirsch AT, Ikeda Y, Mas JL, et al. International prevalence, recognition, and treatment of cardiovascular risk factors in outpatients with atherothrombosis. JAMA 2006; 295:180–189.
531. Yusuf S, Hawken S, Ounpuu S, Dans T, Avezum A, Lanas F, et al. Effect of potentially modifiable risk factors associated with myocardial infarction in 52 countries (the INTERHEART study): case-control study. Lancet 2004; 364:937–952.
532. Kotseva K, De Backer G, De Bacquer D, Ryden L, Hoes A, Grobbee D, et al. Lifestyle and impact on cardiovascular risk factor control in coronary patients across 27 countries: Results from the European Society of Cardiology ESC-EORP EUROASPIRE V registry. Eur J Prev Cardiol 2019; 26:824–835.
533. Lee HH, Lee H, Cho SMJ, Kim DW, Park S, Kim HC. On-treatment blood pressure and cardiovascular outcomes in adults with hypertension and left ventricular hypertrophy. J Am Coll Cardiol 2021; 78:1485–1495.
534. Okin PM, Hille DA, Kjeldsen SE, Dahlöf B, Devereux RB. Impact of lower achieved blood pressure on outcomes in hypertensive patients. J Hypertens 2012; 30:802–810. discussion 810.
535. Heimark S, Mehlum MH, Mancia G, Søraas CL, Liestøl K, Wachtell K, et al. Middle-Aged and Older Patients With Left Ventricular Hypertrophy: Higher Mortality With Drug Treated Systolic Blood Pressure Below 130 mm Hg. Hypertension (Dallas, Tex: 1979) 2023; 80:1739–1748.
536. Heimark S, Mehlum MH, Mancia G, Søraas CL, Liestöl K, Wachtell W, et al. Middle-aged and older patients with left ventricular hypertrophy: Higher mortality with drug treated blood pressure below 130/80 mmHg. Hypertension 2023; in press.
537. Polese A, De Cesare N, Montorsi P, Fabbiocchi F, Guazzi M, Loaldi A, et al. Upward shift of the lower range of coronary flow autoregulation in hypertensive patients with hypertrophy of the left ventricle. Circulation 1991; 83:845–853.
538. Redon J, Mancia G, Sleight P, Schumacher H, Gao P, Pogue J, et al. Safety and efficacy of low blood pressures among patients with diabetes: subgroup analyses from the ONTARGET (ONgoing Telmisartan Alone and in combination with Ramipril Global Endpoint Trial). J Am Coll Cardiol 2012; 59:74–83.
539. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure-lowering on outcome incidence in hypertension: 5. Head-to-head comparisons of various classes of antihypertensive drugs - overview and meta-analyses. J Hypertens 2015; 33:1321–1341.
540. Psaty BM, Lumley T, Furberg CD, Schellenbaum G, Pahor M, Alderman MH, et al. Health outcomes associated with various antihypertensive therapies used as first-line agents: a network meta-analysis. JAMA 2003; 289:2534–2544.
541. Wei J, Galaviz KI, Kowalski AJ, Magee MJ, Haw JS, Narayan KMV, et al. Comparison of Cardiovascular Events Among Users of Different Classes of Antihypertension Medications: A Systematic Review and Network Meta-analysis. JAMA Netw Open 2020; 3:e1921618.
542. Baker WL, Buckley LF, Kelly MS, Bucheit JD, Parod ED, Brown R, et al. Effects of sodium-glucose cotransporter 2 inhibitors on 24-hour ambulatory blood pressure: a systematic review and meta-analysis. J Am Heart Assoc 2017; 6:
543. Wanner C, Lachin JM, Inzucchi SE, Fitchett D, Mattheus M, George J, et al. Empagliflozin and clinical outcomes in patients with type 2 diabetes mellitus, established cardiovascular disease, and chronic kidney disease. Circulation 2018; 137:119–129.
544. Zinman B, Wanner C, Lachin JM, Fitchett D, Bluhmki E, Hantel S, et al. Empagliflozin, cardiovascular outcomes, and mortality in type 2 diabetes. N Engl J Med 2015; 373:2117–2128.
545. Agarwal R, Filippatos G, Pitt B, Anker SD, Rossing P, Joseph A, et al. Cardiovascular and kidney outcomes with finerenone in patients with type 2 diabetes and chronic kidney disease: the FIDELITY pooled analysis. Eur Heart J 2022; 43:474–484.
546. Bakris GL, Agarwal R, Anker SD, Pitt B, Ruilope LM, Rossing P, et al. Effect of Finerenone on Chronic Kidney Disease Outcomes in Type 2 Diabetes. N Engl J Med 2020; 383:2219–2229.
547. Pitt B, Filippatos G, Agarwal R, Anker SD, Bakris GL, Rossing P, et al. Cardiovascular Events with Finerenone in Kidney Disease and Type 2 Diabetes. N Engl J Med 2021; 385:2252–2263.
548. Mancia G, Parodi A, Merlino L, Corrao G. Heterogeneity in antihypertensive treatment discontinuation between drugs belonging to the same class. J Hypertens 2011; 29:1012–1018.
549. van Vark LC, Bertrand M, Akkerhuis KM, Brugts JJ, Fox K, Mourad JJ, et al. Angiotensin-converting enzyme inhibitors reduce mortality in hypertension: a meta-analysis of randomized clinical trials of renin-angiotensin-aldosterone system inhibitors involving 158,998 patients. Eur Heart J 2012; 33:2088–2097.
550. Savarese G, Costanzo P, Cleland JG, Vassallo E, Ruggiero D, Rosano G, et al. A meta-analysis reporting effects of angiotensin-converting enzyme inhibitors and angiotensin receptor blockers in patients without heart failure. J Am Coll Cardiol 2013; 61:131–142.
551. Burnett H, Earley A, Voors AA, Senni M, McMurray JJ, Deschaseaux C, et al. Thirty Years of Evidence on the Efficacy of Drug Treatments for Chronic Heart Failure With Reduced Ejection Fraction: A Network Meta-Analysis. Circ Heart Fail 2017; 10:e003529.
552. Woo KS, Nicholls MG. High prevalence of persistent cough with angiotensin converting enzyme inhibitors in Chinese. Br J Clin Pharmacol 1995; 40:141–144.
553. Tseng DS, Kwong J, Rezvani F, Coates AO. Angiotensin-converting enzyme-related cough among Chinese-Americans. Am J Med 2010; 123:111–185. 183.e.
554. Messerli FH, Bangalore S, Bavishi C, Rimoldi SF. Angiotensin-Converting Enzyme Inhibitors in Hypertension: To Use or Not to Use? J Am Coll Cardiol 2018; 71:1474–1482.
555. Gallo G, Volpe M, Rubattu S. Angiotensin Receptor Blockers in the Management of Hypertension: A Real-World Perspective and Current Recommendations. Vasc Health Risk Manag 2022; 18:507–515.
556. Reboldi G, Angeli F, Cavallini C, Gentile G, Mancia G, Verdecchia P. Comparison between angiotensin-converting enzyme inhibitors and angiotensin receptor blockers on the risk of myocardial infarction, stroke and death: a meta-analysis. J Hypertens 2008; 26:1282–1289.
557. Musini VM, Lawrence KA, Fortin PM, Bassett K, Wright JM. Blood pressure lowering efficacy of renin inhibitors for primary hypertension. Cochrane Database Syst Rev 2017; 4:Cd007066.
558. Pantzaris ND, Karanikolas E, Tsiotsios K, Velissaris D. Renin inhibition with aliskiren: a decade of clinical experience. J Clin Med 2017; 6:
559. Ontarget Investigators, Yusuf S, Teo KK, Pogue J, Dyal L, Copland I, et al. Telmisartan, ramipril, or both in patients at high risk for vascular events. N Engl J Med 2008; 358:1547–1559.
560. Silva AR, Martini AG, Canto GDL, Guerra ENdS, Neves FdAR. Effects of dual blockade in heart failure and renal dysfunction: Systematic review and meta-analysis. Journal of the Renin-Angiotensin-Aldosterone System 2019; 20:1470320319882656.
561. Parving HH, Brenner BM, McMurray JJ, de Zeeuw D, Haffner SM, Solomon SD, et al. Cardiorenal end points in a trial of aliskiren for type 2 diabetes. N Engl J Med 2012; 367:2204–2213.
562. Wang YC, Hsieh TC, Chou CL, Wu JL, Fang TC. Risks of Adverse Events Following Coprescription of Statins and Calcium Channel Blockers: A Nationwide Population-Based Study. Medicine (Baltimore) 2016; 95:e2487.
563. Xu Y, Chang AR, Inker LA, McAdams-DeMarco M, Grams ME, Shin JI. Concomitant Use of Diltiazem With Direct Oral Anticoagulants and Bleeding Risk in Atrial Fibrillation. J Am Heart Assoc 2022; 11:e025723.
564. Pham P, Schmidt S, Lesko L, Lip GYH, Brown JD. Association of Oral Anticoagulants and Verapamil or Diltiazem With Adverse Bleeding Events in Patients With Nonvalvular Atrial Fibrillation and Normal Kidney Function. JAMA Netw Open 2020; 3:e203593.
565. Chalmers J, Mourad JJ, Brzozowska-Villatte R, De Champvallins M, Mancia G. Benefit of treatment based on indapamide mostly combined with perindopril on mortality and cardiovascular outcomes: a pooled analysis of four trials. J Hypertens 2023; 41:508–515.
566. Major outcomes in high-risk hypertensive patients randomized to angiotensin-converting enzyme inhibitor or calcium channel blocker vs diuretic: the antihypertensive and lipid-lowering treatment to prevent heart attack trial (ALLHAT). JAMA 2002; 288:2981–2997.
567. Roush GC, Ernst ME, Kostis JB, Tandon S, Sica DA. Head-to-head comparisons of hydrochlorothiazide with indapamide and chlorthalidone: antihypertensive and metabolic effects. Hypertension (Dallas, Tex: 1979) 2015; 65:1041–1046.
568. Olde Engberink RH, Frenkel WJ, van den Bogaard B, Brewster LM, Vogt L, van den Born BJ. Effects of thiazide-type and thiazide-like diuretics on cardiovascular events and mortality: systematic review and meta-analysis. Hypertension (Dallas, Tex: 1979) 2015; 65:1033–1040.
569. Hripcsak G, Suchard MA, Shea S, Chen R, You SC, Pratt N, et al. Comparison of Cardiovascular and Safety Outcomes of Chlorthalidone vs Hydrochlorothiazide to Treat Hypertension. JAMA Intern Med 2020; 180:542–551.
570. Ishani A, Cushman WC, Leatherman SM, Lew RA, Woods P, Glassman PA, et al. Chlorthalidone vs. Hydrochlorothiazide for Hypertension-Cardiovascular Events. N Engl J Med 2022; 387:2401–2410.
571. Brown MJ, Williams B, Morant SV, Webb DJ, Caulfield MJ, Cruickshank JK, et al. Effect of amiloride, or amiloride plus hydrochlorothiazide, versus hydrochlorothiazide on glucose tolerance and blood pressure (PATHWAY-3): a parallel-group, double-blind randomised phase 4 trial. Lancet Diabetes Endocrinol 2016; 4:136–147.
572. Zillich AJ, Garg J, Basu S, Bakris GL, Carter BL. Thiazide diuretics, potassium, and the development of diabetes: a quantitative review. Hypertension (Dallas, Tex: 1979) 2006; 48:219–224.
573. Agarwal R, Sinha AD, Cramer AE, Balmes-Fenwick M, Dickinson JH, Ouyang F, et al. Chlorthalidone for Hypertension in Advanced Chronic Kidney Disease. N Engl J Med 2021; 385:2507–2519.
574. Mentz RJ, Anstrom KJ, Eisenstein EL, Sapp S, Greene SJ, Morgan S, et al. Effect of Torsemide vs Furosemide After Discharge on All-Cause Mortality in Patients Hospitalized With Heart Failure: The TRANSFORM-HF Randomized Clinical Trial. JAMA 2023; 329:214–223.
575. Sun Q, Sever P. Amiloride: A review. J Renin Angiotensin Aldosterone Syst 2020; 21:1470320320975893.
576. Williams B, MacDonald TM, Morant SV, Webb DJ, Sever P, McInnes GT, et al. Endocrine and haemodynamic changes in resistant hypertension, and blood pressure responses to spironolactone or amiloride: the PATHWAY-2 mechanisms substudies. Lancet Diabetes Endocrinol 2018; 6:464–475.
577. Tu W, Decker BS, He Z, Erdel BL, Eckert GJ, Hellman RN, et al. Triamterene Enhances the Blood Pressure Lowering Effect of Hydrochlorothiazide in Patients with Hypertension. J Gen Intern Med 2016; 31:30–36.
578. Hu LX, Wang D, Liu HL, Zhang QT, Sun DS, Zhang L, et al. A double-blind, placebo-controlled trial on the antihypertensive treatment effect of a quadruple single-pill combination. J Clin Hypertens (Greenwich) 2021; 23:815–822.
579. Chen C, Zhu XY, Li D, Lin Q, Zhou K. Clinical efficacy and safety of spironolactone in patients with resistant hypertension: A systematic review and meta-analysis. Medicine (Baltimore) 2020; 99:e21694.
580. Williams B, MacDonald TM, Morant S, Webb DJ, Sever P, McInnes G, et al. Spironolactone versus placebo, bisoprolol, and doxazosin to determine the optimal treatment for drug-resistant hypertension (PATHWAY-2): a randomised, double-blind, crossover trial. Lancet 2015; 386:2059–2068.
581. Agarwal R, Ruilope LM, Ruiz-Hurtado G, Haller H, Schmieder RE, Anker SD, et al. Effect of finerenone on ambulatory blood pressure in chronic kidney disease in type 2 diabetes. J Hypertens 2023; 41:295–302.
582. Ruilope LM, Agarwal R, Anker SD, Filippatos G, Pitt B, Rossing P, et al. Blood Pressure and Cardiorenal Outcomes With Finerenone in Chronic Kidney Disease in Type 2 Diabetes. Hypertension (Dallas, Tex: 1979) 2022; 79:2685–2695.
583. Law MR, Morris JK, Wald NJ. Use of blood pressure lowering drugs in the prevention of cardiovascular disease: meta-analysis of 147 randomised trials in the context of expectations from prospective epidemiological studies. BMJ 2009; 338:b1665.
584. Thomopoulos C, Bazoukis G, Tsioufis C, Mancia G. Beta-blockers in hypertension: overview and meta-analysis of randomized outcome trials. J Hypertens 2020; 38:1669–1681.
585. do Vale GT, Ceron CS, Gonzaga NA, Simplicio JA, Padovan JC. Three Generations of β-blockers: History, Class Differences and Clinical Applicability. Current hypertension reviews 2019; 15:22–31.
586. Viigimaa M, Vlachopoulos C, Doumas M, Wolf J, Imprialos K, Terentes-Printzios D, et al. Update of the position paper on arterial hypertension and erectile dysfunction. J Hypertens 2020; 38:1220–1234.
587. Rashid AM, Khan MS, Fudim M, DeWald TA, DeVore A, Butler J. Management of Heart Failure with Reduced Ejection Fraction. Curr Probl Cardiol 2023; 101596.
588. Chan You S, Krumholz HM, Suchard MA, Schuemie MJ, Hripcsak G, Chen R, et al. Comprehensive comparative effectiveness and safety of first-line beta-blocker monotherapy in hypertensive patients: a large-scale multicenter observational study. Hypertension (Dallas, Tex: 1979) 2021; 77:1528–1538.
589. Huck DM, Rosenberg MA, Stauffer BL. Nebivolol and incident cardiovascular events in hypertensive patients compared with nonvasodilatory beta blockers. J Hypertens 2022; 40:1019–1029.
590. Basile J, Egan B, Punzi H, Ali S, Li Q, Patel M, et al. Risk of Hospitalization for Cardiovascular Events with β-Blockers in Hypertensive Patients: A Retrospective Cohort Study. Cardiol Ther 2018; 7:173–183.
591. Mancia G, Kjeldsen SE, Kreutz R, Pathak A, Grassi G, Esler M. Individualized beta-blocker treatment for high blood pressure dictated by medical comorbidities: indications beyond the 2018 european society of cardiology/european society of hypertension guidelines. Hypertension (Dallas, Tex: 1979) 2022; 79:1153–1166.
592. Riemer TG, Villagomez Fuentes LE, Algharably EAE, Schäfer MS, Mangelsen E, Fürtig MA, et al. Do β-blockers cause depression?: systematic review and meta-analysis of psychiatric adverse events during β-blocker therapy. Hypertension (Dallas, Tex: 1979) 2021; 77:1539–1548.
593. Grassi G, Quarti-Trevano F, Seravalle G, Dell’Oro R, Facchetti R, Mancia G. Association between the european society of cardiology/european society of hypertension heart rate thresholds for cardiovascular risk and neuroadrenergic markers. Hypertension (Dallas, Tex: 1979) 2020; 76:577–582.
594. Gillman MW, Kannel WB, Belanger A, D’Agostino RB. Influence of heart rate on mortality among persons with hypertension: the Framingham Study. American heart journal 1993; 125:1148–1154.
595. Benetos A, Rudnichi A, Thomas F, Safar M, Guize L. Influence of heart rate on mortality in a French population: role of age, gender, and blood pressure. Hypertension (Dallas, Tex: 1979) 1999; 33:44–52.
596. Benetos A, Thomas F, Bean K, Albaladejo P, Palatini P, Guize L. Resting heart rate in older people: a predictor of survival to age 85. J Am Geriatr Soc 2003; 51:284–285.
597. Julius S, Palatini P, Kjeldsen SE, Zanchetti A, Weber MA, McInnes GT, et al. Usefulness of heart rate to predict cardiac events in treated patients with high-risk systemic hypertension. Am J Cardiol 2012; 109:685–692.
598. Kolloch R, Legler UF, Champion A, Cooper-Dehoff RM, Handberg E, Zhou Q, et al. Impact of resting heart rate on outcomes in hypertensive patients with coronary artery disease: findings from the INternational VErapamil-SR/trandolapril STudy (INVEST). Eur Heart J 2008; 29:1327–1334.
599. Esler M, Kjeldsen SE, Pathak A, Grassi G, Kreutz R, Mancia G. Diverse pharmacological properties, trial results, comorbidity prescribing and neural pathophysiology suggest European hypertension guideline downgrading of beta-blockers is not justified. Blood Press 2022; 31:210–224.
600. Dahlöf B, Sever PS, Poulter NR, Wedel H, Beevers DG, Caulfield M, et al. Prevention of cardiovascular events with an antihypertensive regimen of amlodipine adding perindopril as required versus atenolol adding bendroflumethiazide as required, in the Anglo-Scandinavian Cardiac Outcomes Trial-Blood Pressure Lowering Arm (ASCOT-BPLA): a multicentre randomised controlled trial. Lancet 2005; 366:895–906.
601. Li H, Xu TY, Li Y, Chia YC, Buranakitjaroen P, Cheng HM, et al. Role of alpha1-blockers in the current management of hypertension. J Clin Hypertens (Greenwich) 2022; 24:1180–1186.
602. Corrao G, Mazzola P, Monzio Compagnoni M, Rea F, Merlino L, Annoni G, et al. Antihypertensive medications, loop diuretics, and risk of hip fracture in the elderly: a population-based cohort study of 81,617 Italian patients newly treated between 2005 and 2009. Drugs Aging 2015; 32:927–936.
603. Krieger EM, Drager LF, Giorgi DMA, Pereira AC, Barreto-Filho JAS, Nogueira AR, et al. Spironolactone versus clonidine as a fourth-drug therapy for resistant hypertension: the rehot randomized study (resistant hypertension optimal treatment). Hypertension (Dallas, Tex: 1979) 2018; 71:681–690.
604. Chua SK, Lai WT, Chen LC, Hung HF. The antihypertensive effects and safety of LCZ696 in patients with hypertension: a systemic review and meta-analysis of randomized controlled trials. J Clin Med 2021; 10:
605. Ruilope LM, Dukat A, Bohm M, Lacourciere Y, Gong J, Lefkowitz MP. Blood-pressure reduction with LCZ696, a novel dual-acting inhibitor of the angiotensin II receptor and neprilysin: a randomised, double-blind, placebo-controlled, active comparator study. Lancet 2010; 375:1255–1266.
606. Judd E, Calhoun DA. Apparent and true resistant hypertension: definition, prevalence and outcomes. J Hum Hypertens 2014; 28:463–468.
607. Pallares-Carratala V, Bonig-Trigueros I, Palazon-Bru A, Esteban-Giner MJ, Gil-Guillen VF, Giner-Galvan V. Clinical inertia in hypertension: a new holistic and practical concept within the cardiovascular continuum and clinical care process. Blood Press 2019; 28:217–228.
608. Rea F, Corrao G, Merlino L, Mancia G. Initial antihypertensive treatment strategies and therapeutic inertia. Hypertension (Dallas, Tex: 1979) 2018; 72:846–853.
609. Kjeldsen SE, Julius S, Dahlof B, Weber MA. Physician (investigator) inertia in apparent treatment-resistant hypertension - insights from large randomized clinical trials. Lennart Hansson Memorial Lecture. Blood Press 2015; 24:1–6.
610. Milman T, Joundi RA, Alotaibi NM, Saposnik G. Clinical inertia in the pharmacological management of hypertension: A systematic review and meta-analysis. Medicine (Baltimore) 2018; 97:e11121.
611. Berra E, Azizi M, Capron A, Høieggen A, Rabbia F, Kjeldsen SE, et al. Evaluation of adherence should become an integral part of assessment of patients with apparently treatment-resistant hypertension. Hypertension (Dallas, Tex: 1979) 2016; 68:297–306.
612. Mancia G, Rea F, Corrao G, Grassi G. Two-drug combinations as first-step antihypertensive treatment. Circ Res 2019; 124:1113–1123.
613. Feng Y, Han M, Qie R, Huang S, Li Q, Guo C, et al. Adherence to antihypertensive medications for secondary prevention of cardiovascular disease events: a dose-response meta-analysis. Public Health 2021; 196:179–185.
614. Savaré L, Rea F, Corrao G, Mancia G. Use of initial and subsequent antihypertensive combination treatment in the last decade: analysis of a large Italian database. J Hypertens 2022; 40:1768–1775.
615. Bakris G, Sarafidis P, Agarwal R, Ruilope L. Review of blood pressure control rates and outcomes. J Am Soc Hypertens 2014; 8:127–141.
616. Choudhry NK, Kronish IM, Vongpatanasin W, Ferdinand KC, Pavlik VN, Egan BM, et al. Medication Adherence and Blood Pressure Control: A Scientific Statement From the American Heart Association. Hypertension (Dallas, Tex: 1979) 2022; 79:e1–e14.
617. Parati G, Kjeldsen S, Coca A, Cushman W C, Wang J. Adherence to single-pill versus free-equivalent combination therapy in hypertension: a systematic review and meta-analysis. Hypertension (Dallas, Tex: 1979) 2021; 77:692–705.
618. Corrao G, Parodi A, Zambon A, Heiman F, Filippi A, Cricelli C, et al. Reduced discontinuation of antihypertensive treatment by two-drug combination as first step. Evidence from daily life practice. J Hypertens 2010; 28:1584–1590.
619. Mancia G, Zambon A, Soranna D, Merlino L, Corrao G. Factors involved in the discontinuation of antihypertensive drug therapy: an analysis from real life data. J Hypertens 2014; 32:1708–1715. discussion 1716.
620. Egan BM, Bandyopadhyay D, Shaftman SR, Wagner CS, Zhao Y, Yu-Isenberg KS. Initial monotherapy and combination therapy and hypertension control the first year. Hypertension (Dallas, Tex: 1979) 2012; 59:1124–1131.
621. Egan BM, Kjeldsen SE, Narkiewicz K, Kreutz R, Burnier M. Single-pill combinations, hypertension control and clinical outcomes: potential, pitfalls and solutions. Blood Press 2022; 31:164–168.
622. Gradman AH, Parisé H, Lefebvre P, Falvey H, Lafeuille MH, Duh MS. Initial combination therapy reduces the risk of cardiovascular events in hypertensive patients: a matched cohort study. Hypertension (Dallas, Tex: 1979) 2013; 61:309–318.
623. Corrao G, Nicotra F, Parodi A, Zambon A, Heiman F, Merlino L, et al. Cardiovascular protection by initial and subsequent combination of antihypertensive drugs in daily life practice. Hypertension (Dallas, Tex: 1979) 2011; 58:566–572.
624. Thomopoulos C, Bazoukis G, Grassi G, Tsioufis C, Mancia G. Monotherapy vs combination treatments of different complexity: a meta-analysis of blood pressure lowering randomized outcome trials. J Hypertens 2021; 39:846–855.
625. Dahlof B, Devereux RB, Kjeldsen SE, Julius S, Beevers G, de Faire U, et al. Cardiovascular morbidity and mortality in the Losartan Intervention For Endpoint reduction in hypertension study (LIFE): a randomised trial against atenolol. Lancet 2002; 359:995–1003.
626. Jamerson K, Weber MA, Bakris GL, Dahlof B, Pitt B, Shi V, et al. Benazepril plus amlodipine or hydrochlorothiazide for hypertension in high-risk patients. N Engl J Med 2008; 359:2417–2428.
627. Matsuzaki M, Ogihara T, Umemoto S, Rakugi H, Matsuoka H, Shimada K, et al. Prevention of cardiovascular events with calcium channel blocker-based combination therapies in patients with hypertension: a randomized controlled trial. J Hypertens 2011; 29:1649–1659.
628. Ogihara T, Saruta T, Rakugi H, Saito I, Shimamoto K, Matsuoka H, et al. Combination therapy of hypertension in the elderly: a subgroup analysis of the Combination of OLMesartan and a calcium channel blocker or diuretic in Japanese elderly hypertensive patients trial. Hypertens Res 2015; 38:89–96.
629. Randomised trial of a perindopril-based blood-pressure-lowering regimen among 6,105 individuals with previous stroke or transient ischaemic attack. Lancet 2001; 358:1033–1041.
630. Patel A, MacMahon S, Chalmers J, Neal B, Woodward M, Billot L, et al. Effects of a fixed combination of perindopril and indapamide on macrovascular and microvascular outcomes in patients with type 2 diabetes mellitus (the ADVANCE trial): a randomised controlled trial. Lancet 2007; 370:829–840.
631. Lithell H, Hansson L, Skoog I, Elmfeldt D, Hofman A, Olofsson B, et al. The Study on Cognition and Prognosis in the Elderly (SCOPE): principal results of a randomized double-blind intervention trial. J Hypertens 2003; 21:875–886.
632. Ogawa H, Kim-Mitsuyama S, Matsui K, Jinnouchi T, Jinnouchi H, Arakawa K. Angiotensin II receptor blocker-based therapy in Japanese elderly, high-risk, hypertensive patients. Am J Med 2012; 125:981–990.
633. Liu L, Zhang Y, Liu G, Li W, Zhang X, Zanchetti A. The Felodipine Event Reduction (FEVER) Study: a randomized long-term placebo-controlled trial in Chinese hypertensive patients. J Hypertens 2005; 23:2157–2172.
634. Coope J, Warrender TS. Randomised trial of treatment of hypertension in elderly patients in primary care. Br Med J (Clin Res Ed) 1986; 293:1145–1151.
635. Prevention of stroke by antihypertensive drug treatment in older persons with isolated systolic hypertension. Final results of the Systolic Hypertension in the Elderly Program (SHEP). SHEP Cooperative Research Group. JAMA 1991; 265:3255–3264.
636. Dahlöf B, Lindholm LH, Hansson L, Scherstén B, Ekbom T, Wester PO. Morbidity and mortality in the Swedish Trial in Old Patients with Hypertension (STOP-Hypertension). Lancet 1991; 338:1281–1285.
637. Hansson L, Lindholm LH, Ekbom T, Dahlöf B, Lanke J, Scherstén B, et al. Randomised trial of old and new antihypertensive drugs in elderly patients: cardiovascular mortality and morbidity the Swedish Trial in Old Patients with Hypertension-2 study. Lancet 1999; 354:1751–1756.
638. Yusuf S, Teo KK, Pogue J, Dyal L, Copland I, Schumacher H, et al. Telmisartan, ramipril, or both in patients at high risk for vascular events. N Engl J Med 2008; 358:1547–1559.
639. Hansson L, Lindholm LH, Niskanen L, Lanke J, Hedner T, Niklason A, et al. Effect of angiotensin-converting-enzyme inhibition compared with conventional therapy on cardiovascular morbidity and mortality in hypertension: the Captopril Prevention Project (CAPPP) randomised trial. Lancet 1999; 353:611–616.
640. Dahlöf B, Devereux RB, Kjeldsen SE, Julius S, Beevers G, de Faire U, et al. Cardiovascular morbidity and mortality in the Losartan Intervention For Endpoint reduction in hypertension study (LIFE): a randomised trial against atenolol. Lancet 2002; 359:995–1003.
641. Zanchetti A, Hennig M, Baurecht H, Tang R, Cuspidi C, Carugo S, et al. Prevalence and incidence of the metabolic syndrome in the European Lacidipine Study on Atherosclerosis (ELSA) and its relation with carotid intima-media thickness. J Hypertens 2007; 25:2463–2470.
642. Black HR, Elliott WJ, Grandits G, Grambsch P, Lucente T, White WB, et al. Principal results of the Controlled Onset Verapamil Investigation of Cardiovascular End Points (CONVINCE) trial. Jama 2003; 289:2073–2082.
643. Ojji DB, Mayosi B, Francis V, Badri M, Cornelius V, Smythe W, et al. Comparison of Dual Therapies for Lowering Blood Pressure in Black Africans. N Engl J Med 2019; 380:2429–2439.
644. Hansson L, Hedner T, Lund-Johansen P, Kjeldsen SE, Lindholm LH, Syvertsen JO, et al. Randomised trial of effects of calcium antagonists compared with diuretics and beta-blockers on cardiovascular morbidity and mortality in hypertension: the Nordic Diltiazem (NORDIL) study. Lancet 2000; 356:359–365.
645. Pepine CJ, Handberg EM, Cooper-DeHoff RM, Marks RG, Kowey P, Messerli FH, et al. A calcium antagonist vs a non-calcium antagonist hypertension treatment strategy for patients with coronary artery disease. The International Verapamil-Trandolapril Study (INVEST): a randomized controlled trial. JAMA 2003; 290:2805–2816.
646. Xu W, Goldberg SI, Shubina M, Turchin A. Optimal systolic blood pressure target, time to intensification, and time to follow-up in treatment of hypertension: population based retrospective cohort study. BMJ 2015; 350:h158.
647. Rea F, Savaré L, Franchi M, Corrao G, Mancia G. Adherence to Treatment by Initial Antihypertensive Mono and Combination Therapies. Am J Hypertens 2021; 34:1083–1091.
648. Rea F, Corrao G, Merlino L, Mancia G. Early cardiovascular protection by initial two-drug fixed-dose combination treatment vs. monotherapy in hypertension. Eur Heart J 2018; 39:3654–3661.
649. Brant LCC, Passaglia LG, Pinto-Filho MM, de Castilho FM, Ribeiro ALP, Nascimento BR. The Burden of Resistant Hypertension Across the World. Curr Hypertens Rep 2022; 24:55–66.
650. Webster R, Salam A, de Silva HA, Selak V, Stepien S, Rajapakse S, et al. Fixed Low-Dose Triple Combination Antihypertensive Medication vs Usual Care for Blood Pressure Control in Patients With Mild to Moderate Hypertension in Sri Lanka: A Randomized Clinical Trial. JAMA 2018; 320:566–579.
651. Schmieder RE, Wassmann S, Predel HG, Weisser B, Blettenberg J, Gillessen A, et al. Improved persistence to medication, decreased cardiovascular events and reduced all-cause mortality in hypertensive patients with use of single-pill combinations: results from the START-study. Hypertension (Dallas, Tex: 1979) 2023; 80:1127–1135.
652. Persu A, Lopez-Sublet M, Algharably EAE, Kreutz R. Starting antihypertensive drug treatment with combination therapy: controversies in hypertension - pro side of the argument. Hypertension (Dallas, Tex: 1979) 2021; 77:800–805.
653. Giles TD, Weber MA, Basile J, Gradman AH, Bharucha DB, Chen W, et al. Efficacy and safety of nebivolol and valsartan as fixed-dose combination in hypertension: a randomised, multicentre study. Lancet 2014; 383:1889–1898.
654. Chow CK, Thakkar J, Bennett A, Hillis G, Burke M, Usherwood T, et al. Quarter-dose quadruple combination therapy for initial treatment of hypertension: placebo-controlled, crossover, randomised trial and systematic review. Lancet 2017; 389:1035–1042.
655. Chow CK, Atkins ER, Hillis GS, Nelson MR, Reid CM, Schlaich MP, et al. Initial treatment with a single pill containing quadruple combination of quarter doses of blood pressure medicines versus standard dose monotherapy in patients with hypertension (QUARTET): a phase 3, randomised, double-blind, active-controlled trial. Lancet 2021; 398:1043–1052.
656. Coca A, Castellano JM, Camafort M, Fuster V. Polypill in cardiovascular disease prevention: recent advances. Pol Arch Intern Med 2023; 133.
657. Tomaszewski M, White C, Patel P, Masca N, Damani R, Hepworth J, et al. High rates of non-adherence to antihypertensive treatment revealed by high-performance liquid chromatography-tandem mass spectrometry (HP LC-MS/MS) urine analysis. Heart 2014; 100:855–861.
658. Gupta P, Patel P, Strauch B, Lai FY, Akbarov A, Gulsin GS, et al. Biochemical screening for nonadherence is associated with blood pressure reduction and improvement in adherence. Hypertension (Dallas, Tex: 1979) 2017; 70:1042–1048.
659. Lane D, Lawson A, Burns A, Azizi M, Burnier M, Jones DJL, et al. Nonadherence in hypertension: how to develop and implement chemical adherence testing. Hypertension (Dallas, Tex: 1979) 2022; 79:12–23.
660. Ibanez B, James S, Agewall S, Antunes MJ, Bucciarelli-Ducci C, Bueno H, et al. 2017 ESC Guidelines for the management of acute myocardial infarction in patients presenting with ST-segment elevation: The Task Force for the management of acute myocardial infarction in patients presenting with ST-segment elevation of the European Society of Cardiology (ESC). Eur Heart J 2018; 39:119–177.
661. Yusuf S, Joseph P, Dans A, Gao P, Teo K, Xavier D, et al. Polypill with or without Aspirin in Persons without Cardiovascular Disease. N Engl J Med 2021; 384:216–228.
662. Castellano JM, Pocock SJ, Bhatt DL, Quesada AJ, Owen R, Fernandez-Ortiz A, et al. Polypill Strategy in Secondary Cardiovascular Prevention. N Engl J Med 2022; 387:967–977.
663. Joseph P, Roshandel G, Gao P, Pais P, Lonn E, Xavier D, et al. Fixed-dose combination therapies with and without aspirin for primary prevention of cardiovascular disease: an individual participant data meta-analysis. Lancet 2021; 398:1133–1146.
664. Coca A, Agabiti-Rosei E, Cifkova R, Manolis AJ, Redon J, Mancia G. The polypill in cardiovascular prevention: evidence, limitations and perspective - position paper of the European Society of Hypertension. J Hypertens 2017; 35:1546–1553.
665. Ambrosioni E, Leonetti G, Pessina AC, Rappelli A, Trimarco B, Zanchetti A. Patterns of hypertension management in Italy: results of a pharmacoepidemiological survey on antihypertensive therapy. Scientific Committee of the Italian Pharmacoepidemiological Survey on Antihypertensive Therapy. J Hypertens 2000; 18:1691–1699.
666. Algharably EA-H, Kreutz R. Antihypertensive Drug Classes. Manual of Hypertension of the European Society of Hypertension: CRC Press; 2019. pp. 325-335.
667. Kreutz R, Algharably EAe-H. Antihypertensive Drugs. In: Offermanns, S., Rosenthal, W., editors. Encyclopedia of Molecular Pharmacology. Cham: Springer International Publishing; 2021. pp. 165-174.
668. Ruben MD, Smith DF, FitzGerald GA, Hogenesch JB. Dosing time matters. Science 2019; 365:547–549.
669. Gupta P, Patel P, Štrauch B, Lai FY, Akbarov A, Marešová V, et al. Risk factors for nonadherence to antihypertensive treatment. Hypertension (Dallas, Tex: 1979) 2017; 69:1113–1120.
670. Mackenzie IS, Rogers A, Poulter NR, Williams B, Brown MJ, Webb DJ, et al. Cardiovascular outcomes in adults with hypertension with evening versus morning dosing of usual antihypertensives in the UK (TIME study): a prospective, randomised, open-label, blinded-endpoint clinical trial. Lancet 2022; 400:1417–1425.
671. Burnier M, Kreutz R, Narkiewicz K, Kjeldsen S, Oparil S, Mancia G. Circadian variations in blood pressure and their implications for the administration of antihypertensive drugs: is dosing in the evening better than in the morning? J Hypertens 2020; 38:1396–1406.
672. Brunström M, Kjeldsen SE, Kreutz R, Gjesdal K, Narkiewicz K, Burnier M, et al. Missing verification of source data in hypertension research: The HYGIA PROJECT in perspective. Hypertension (Dallas, Tex: 1979) 2021; 78:555–558.
673. Stergiou G, Brunström M, MacDonald T, Kyriakoulis KG, Bursztyn M, Khan N, et al. Bedtime dosing of antihypertensive medications: systematic review and consensus statement: International Society of Hypertension position paper endorsed by World Hypertension League and European Society of Hypertension. J Hypertens 2022; 40:1847–1858.
674. Kjeldsen SE, Egan BM, Narkiewicz K, Kreutz R, Burnier M, Oparil S, et al. TIME to face the reality about evening dosing of antihypertensive drugs in hypertension. Blood Press 2023; 32:1–3.
675. Vrijens B, Vincze G, Kristanto P, Urquhart J, Burnier M. Adherence to prescribed antihypertensive drug treatments: longitudinal study of electronically compiled dosing histories. BMJ 2008; 336:1114–1117.
676. van der Wardt V, Harrison JK, Welsh T, Conroy S, Gladman J. Withdrawal of antihypertensive medication: a systematic review. J Hypertens 2017; 35:1742–1749.
677. Ekbom T, Lindholm LH, Odén A, Dahlöf B, Hansson L, Wester PO, et al. A 5-year prospective, observational study of the withdrawal of antihypertensive treatment in elderly people. J Intern Med 1994; 235:581–588.
678. Ikeda N, Sapienza D, Guerrero R, Aekplakorn W, Naghavi M, Mokdad AH, et al. Control of hypertension with medication: a comparative analysis of national surveys in 20 countries. Bull World Health Organ 2014; 92:10–19c.
679. Reeve E, Jordan V, Thompson W, Sawan M, Todd A, Gammie TM, et al. Withdrawal of antihypertensive drugs in older people. Cochrane Database Syst Rev 2020; 6:C012572.
680. Dyer AR, Stamler J, Berkson DM, Lindberg HA, Stevens E. High blood-pressure: a risk factor for cancer mortality? Lancet 1975; 1:1051–1056.
681. Kreutz R, Algharably EAH, Douros A. Reviewing the effects of thiazide and thiazide-like diuretics as photosensitizing drugs on the risk of skin cancer. J Hypertens 2019; 37:1950–1958.
682. Stocks T, Van Hemelrijck M, Manjer J, Bjørge T, Ulmer H, Hallmans G, et al. Blood pressure and risk of cancer incidence and mortality in the Metabolic Syndrome and Cancer Project. Hypertension (Dallas, Tex: 1979) 2012; 59:802–810.
683. Pottegård A, Hallas J, Olesen M, Svendsen MT, Habel LA, Friedman GD, et al. Hydrochlorothiazide use is strongly associated with risk of lip cancer. J Intern Med 2017; 282:322–331.
684. Pedersen SA, Gaist D, Schmidt SAJ, Hölmich LR, Friis S, Pottegård A. Hydrochlorothiazide use and risk of nonmelanoma skin cancer: A nationwide case-control study from Denmark. J Am Acad Dermatol 2018; 78:673–681. e679.
685. Bigagli E, Mugelli A, Mancia G. A reverse translational pharmacological approach to understand the underlying mechanisms of the reported association between hydrochlorothiazide and non-melanoma skin cancer. J Hypertens 2022; 40:1647–1649.
686. Pottegård A, Pedersen SA, Schmidt SAJ, Lee CN, Hsu CK, Liao TC, et al. Use of hydrochlorothiazide and risk of skin cancer: a nationwide Taiwanese case-control study. Br J Cancer 2019; 121:973–978.
687. Rouette J, Yin H, Pottegård A, Nirantharakumar K, Azoulay L. Use of Hydrochlorothiazide and Risk of Melanoma and Nonmelanoma Skin Cancer. Drug Saf 2021; 44:245–254.
688. Gotzinger F, Wilke T, Hardtstock F, Krieger J, Maywald U, Kunz M, et al. Association of hydrochlorothiazide treatment compared with alternative diuretics with overall and skin cancer risk: a propensity-matched cohort study. J Hypertens 2023; 41:926–933.
689. Copland E, Canoy D, Nazarzadeh M, Bidel Z, Ramakrishnan R, Woodward M, et al. Antihypertensive treatment and risk of cancer: an individual participant data meta-analysis. The Lancet Oncology 2021; 22:558–570.
690. Mahfoud F, Kieble M, Enners S, Kintscher U, Laufs U, Bohm M, et al. Use of fixed-dose combination antihypertensives in Germany between 2016 and 2020: an example of guideline inertia. Clin Res Cardiol 2023; 112:197–202.
691. Mach F, Baigent C, Catapano AL, Koskinas KC, Casula M, Badimon L, et al. 2019 ESC/EAS Guidelines for the management of dyslipidaemias: lipid modification to reduce cardiovascular risk. Eur Heart J 2020; 41:111–188.
692. Sever PS, Dahlöf B, Poulter NR, Wedel H, Beevers G, Caulfield M, et al. Prevention of coronary and stroke events with atorvastatin in hypertensive patients who have average or lower-than-average cholesterol concentrations, in the Anglo-Scandinavian Cardiac Outcomes Trial--Lipid Lowering Arm (ASCOT-LLA): a multicentre randomised controlled trial. Lancet 2003; 361:1149–1158.
693. Collins R, Reith C, Emberson J, Armitage J, Baigent C, Blackwell L, et al. Interpretation of the evidence for the efficacy and safety of statin therapy. Lancet 2016; 388:2532–2561.
694. Silverman MG, Ference BA, Im K, Wiviott SD, Giugliano RP, Grundy SM, et al. Association between lowering LDL-C and cardiovascular risk reduction among different therapeutic interventions: a systematic review and meta-analysis. JAMA 2016; 316:1289–1297.
695. Yusuf S, Bosch J, Dagenais G, Zhu J, Xavier D, Liu L, et al. Cholesterol Lowering in Intermediate-Risk Persons without Cardiovascular Disease. N Engl J Med 2016; 374:2021–2031.
696. Yusuf S, Lonn E, Pais P, Bosch J, Lopez-Jaramillo P, Zhu J, et al. Blood-Pressure and Cholesterol Lowering in Persons without Cardiovascular Disease. N Engl J Med 2016; 374:2032–2043.
697. Mihaylova B, Emberson J, Blackwell L, Keech A, Simes J, Barnes EH, et al. The effects of lowering LDL cholesterol with statin therapy in people at low risk of vascular disease: meta-analysis of individual data from 27 randomised trials. Lancet 2012; 380:581–590.
698. Wood FA, Howard JP, Finegold JA, Nowbar AN, Thompson DM, Arnold AD, et al. N-of-1 Trial of a Statin, Placebo, or No Treatment to Assess Side Effects. N Engl J Med 2020; 383:2182–2184.
699. Rea F, Savaré L, Corrao G, Mancia G. Adherence to Lipid-Lowering Treatment by Single-Pill Combination of Statin and Ezetimibe. Adv Ther 2021; 38:5270–5285.
700. Bhatt DL, Steg PG, Miller M, Brinton EA, Jacobson TA, Ketchum SB, et al. Cardiovascular risk reduction with icosapent ethyl for hypertriglyceridemia. N Engl J Med 2019; 380:11–22.
701. Nicholls SJ, Lincoff AM, Garcia M, Bash D, Ballantyne CM, Barter PJ, et al. Effect of high-dose omega-3 fatty acids vs corn oil on major adverse cardiovascular events in patients at high cardiovascular risk: the STRENGTH randomized clinical trial. JAMA 2020; 324:2268–2280.
702. Stark K, Massberg S. Interplay between inflammation and thrombosis in cardiovascular pathology. Nat Rev Cardiol 2021; 18:666–682.
703. Collaborative meta-analysis of randomised trials of antiplatelet therapy for prevention of death, myocardial infarction, and stroke in high risk patients. Bmj 2002; 324:71–86.
704. Shantsila E, Kozieł-Siołkowska M, Lip GY. Antiplatelet agents and anticoagulants for hypertension. Cochrane Database Syst Rev 2022; 7:
705. Hansson L, Zanchetti A, Carruthers SG, Dahlöf B, Elmfeldt D, Julius S, et al. Effects of intensive blood-pressure lowering and low-dose aspirin in patients with hypertension: principal results of the Hypertension Optimal Treatment (HOT) randomised trial. HOT Study Group. Lancet 1998; 351:1755–1762.
706. Carey RM, Calhoun DA, Bakris GL, Brook RD, Daugherty SL, Dennison-Himmelfarb CR, et al. Resistant hypertension: detection, evaluation, and management: a scientific statement from the American Heart Association. Hypertension (Dallas, Tex: 1979) 2018; 72:e53–e90.
707. Achelrod D, Wenzel U, Frey S. Systematic review and meta-analysis of the prevalence of resistant hypertension in treated hypertensive populations. Am J Hypertens 2015; 28:355–361.
708. Tanner RM, Calhoun DA, Bell EK, Bowling CB, Gutierrez OM, Irvin MR, et al. Prevalence of apparent treatment-resistant hypertension among individuals with CKD. Clin J Am Soc Nephrol 2013; 8:1583–1590.
709. Smith SM, Gurka MJ, Winterstein AG, Pepine CJ, Cooper-DeHoff RM. Incidence, prevalence, and predictors of treatment-resistant hypertension with intensive blood pressure lowering. J Clin Hypertens (Greenwich) 2019; 21:825–834.
710. de la Sierra A, Segura J, Banegas JR, Gorostidi M, de la Cruz JJ, Armario P, et al. Clinical features of 8295 patients with resistant hypertension classified on the basis of ambulatory blood pressure monitoring. Hypertension (Dallas, Tex: 1979) 2011; 57:898–902.
711. Carey RM, Sakhuja S, Calhoun DA, Whelton PK, Muntner P. Prevalence of apparent treatment-resistant hypertension in the United States. Hypertension (Dallas, Tex: 1979) 2019; 73:424–431.
712. Egan BM, Zhao Y, Li J, Brzezinski WA, Todoran TM, Brook RD, et al. Prevalence of optimal treatment regimens in patients with apparent treatment-resistant hypertension based on office blood pressure in a community-based practice network. Hypertension (Dallas, Tex: 1979) 2013; 62:691–697.
713. Gaddam KK, Nishizaka MK, Pratt-Ubunama MN, Pimenta E, Aban I, Oparil S, et al. Characterization of resistant hypertension: association between resistant hypertension, aldosterone, and persistent intravascular volume expansion. Arch Intern Med 2008; 168:1159–1164.
714. Martins LC, Figueiredo VN, Quinaglia T, Boer-Martins L, Yugar-Toledo JC, Martin JF, et al. Characteristics of resistant hypertension: ageing, body mass index, hyperaldosteronism, cardiac hypertrophy and vascular stiffness. J Hum Hypertens 2011; 25:532–538.
715. Gupta AK, Nasothimiou EG, Chang CL, Sever PS, Dahlof B, Poulter NR. Baseline predictors of resistant hypertension in the Anglo-Scandinavian Cardiac Outcome Trial (ASCOT): a risk score to identify those at high-risk. J Hypertens 2011; 29:2004–2013.
716. Calhoun DA. Hyperaldosteronism as a common cause of resistant hypertension. Annual review of medicine 2013; 64:233–247.
717. Freeman MW, Halvorsen YD, Marshall W, Pater M, Isaacsohn J, Pearce C, et al. Phase 2 Trial of Baxdrostat for Treatment-Resistant Hypertension. N Engl J Med 2023; 388:395–405.
718. Schlaich MP, Bellet M, Weber MA, Danaietash P, Bakris GL, Flack JM, et al. Dual endothelin antagonist aprocitentan for resistant hypertension (PRECISION): a multicentre, blinded, randomised, parallel-group, phase 3 trial. Lancet 2022; 400:1927–1937.
719. Mendes M, Dubourg J, Blanchard A, Bergerot D, Courand PY, Forni V, et al. Copeptin is increased in resistant hypertension. J Hypertens 2016; 34:2458–2464.
720. Grassi G, Seravalle G, Brambilla G, Pini C, Alimento M, Facchetti R, et al. Marked sympathetic activation and baroreflex dysfunction in true resistant hypertension. Int J Cardiol 2014; 177:1020–1025.
721. Dudenbostel T, Acelajado MC, Pisoni R, Li P, Oparil S, Calhoun DA. Refractory hypertension: evidence of heightened sympathetic activity as a cause of antihypertensive treatment failure. Hypertension (Dallas, Tex: 1979) 2015; 66:126–133.
722. Matanes F, Khan MB, Siddiqui M, Dudenbostel T, Calhoun D, Oparil S. An update on refractory hypertension. Curr Hypertens Rep 2022; 24:225–234.
723. Velasco A, Siddiqui M, Kreps E, Kolakalapudi P, Dudenbostel T, Arora G, et al. Refractory hypertension is not attributable to intravascular fluid retention as determined by intracardiac volumes. Hypertension (Dallas, Tex: 1979) 2018; 72:343–349.
724. Cuspidi C, Macca G, Sampieri L, Michev I, Salerno M, Fusi V, et al. High prevalence of cardiac and extracardiac target organ damage in refractory hypertension. J Hypertens 2001; 19:2063–2070.
725. De Nicola L, Gabbai FB, Agarwal R, Chiodini P, Borrelli S, Bellizzi V, et al. Prevalence and prognostic role of resistant hypertension in chronic kidney disease patients. J Am Coll Cardiol 2013; 61:2461–2467.
726. Daugherty SL, Powers JD, Magid DJ, Tavel HM, Masoudi FA, Margolis KL, et al. Incidence and prognosis of resistant hypertension in hypertensive patients. Circulation 2012; 125:1635–1642.
727. Kario K, Hoshide S, Narita K, Okawara Y, Kanegae H. Cardiovascular prognosis in drug-resistant hypertension stratified by 24-hour ambulatory blood pressure: the JAMP study. Hypertension (Dallas, Tex: 1979) 2021; 78:1781–1790.
728. Reincke M, Bancos I, Mulatero P, Scholl UI, Stowasser M, Williams TA. Diagnosis and treatment of primary aldosteronism. Lancet Diabetes Endocrinol 2021; 9:876–892.
729. Pedrosa RP, Drager LF, Gonzaga CC, Sousa MG, de Paula LK, Amaro AC, et al. Obstructive sleep apnea: the most common secondary cause of hypertension associated with resistant hypertension. Hypertension (Dallas, Tex: 1979) 2011; 58:811–817.
730. Carnethon MR, Johnson DA. Sleep and Resistant Hypertension. Curr Hypertens Rep 2019; 21:34.
731. Parati G, Ochoa JE, Bilo G, Mattaliano P, Salvi P, Kario K, et al. Obstructive sleep apnea syndrome as a cause of resistant hypertension. Hypertens Res 2014; 37:601–613.
732. Kario K, Hettrick DA, Prejbisz A, Januszewicz A. Obstructive sleep apnea-induced neurogenic nocturnal hypertension: a potential role of renal denervation? Hypertension (Dallas, Tex: 1979) 2021; 77:1047–1060.
733. Chedier B, Roderjan CN, Cavalcanti AH, de Souza F, Muxfeldt ES. Prevalence and associated factors of obstructive sleep apnea in refractory hypertension. J Hypertens 2022; 40:1327–1335.
734. Kably B, Billaud EM, Derobertmasure A, Blanchard A, Boutouyrie P, Azizi M. Urine N-acetyl-Ser-Asp-Lys-Pro measurement as a versatile biomarker to assess adherence to angiotensin-converting enzyme inhibitors. J Hypertens 2022; 40:348–355.
735. Postel-Vinay N, Bobrie G, Steichen O, Sosner P, Baguet JP, Gosse P, et al. HY-Quest, standardized patient questionnaire to be completed at home before a first visit for hypertension: a validation study in specialized centres in France. J Hypertens 2014; 32:693–698.
736. Agarwal R, Sinha AD, Tu W. Chlorthalidone for Resistant Hypertension in Advanced Chronic Kidney Disease. Circulation 2022; 146:718–720.
737. Tamargo J, Segura J, Ruilope LM. Diuretics in the treatment of hypertension. Part 2: loop diuretics and potassium-sparing agents. Expert opinion on pharmacotherapy 2014; 15:605–621.
738. Tsujimoto T, Kajio H. Spironolactone use and improved outcomes in patients with heart failure with preserved ejection fraction with resistant hypertension. J Am Heart Assoc 2020; 9:e018827.
739. Bolignano D, Palmer SC, Navaneethan SD, Strippoli GF. Aldosterone antagonists for preventing the progression of chronic kidney disease. Cochrane Database Syst Rev 2014; 4:CD007004.
740. Agarwal R, Rossignol P, Romero A, Garza D, Mayo MR, Warren S, et al. Patiromer versus placebo to enable spironolactone use in patients with resistant hypertension and chronic kidney disease (AMBER): a phase 2, randomised, double-blind, placebo-controlled trial. Lancet 2019; 394:1540–1550.
741. Hoy SM. Sodium zirconium cyclosilicate: a review in hyperkalaemia. Drugs 2018; 78:1605–1613.
742. Agarwal R, Kolkhof P, Bakris G, Bauersachs J, Haller H, Wada T, et al. Steroidal and non-steroidal mineralocorticoid receptor antagonists in cardiorenal medicine. Eur Heart J 2021; 42:152–161.
743. Azizi M, Sanghvi K, Saxena M, Gosse P, Reilly JP, Levy T, et al. Ultrasound renal denervation for hypertension resistant to a triple medication pill (RADIANCE-HTN TRIO): a randomised, multicentre, single-blind, sham-controlled trial. Lancet 2021; 397:2476–2486.
744. Bakris G, Pergola PE, Delgado B, Genov D, Doliashvili T, Vo N, et al. Effect of KBP-5074 on blood pressure in advanced chronic kidney disease: results of the BLOCK-CKD Study. Hypertension (Dallas, Tex: 1979) 2021; 78:74–81.
745. Azizi M, Schmieder RE, Mahfoud F, Weber MA, Daemen J, Davies J, et al. Endovascular ultrasound renal denervation to treat hypertension (RADIANCE-HTN SOLO): a multicentre, international, single-blind, randomised, sham-controlled trial. Lancet 2018; 391:2335–2345.
746. Bohm M, Kario K, Kandzari DE, Mahfoud F, Weber MA, Schmieder RE, et al. Efficacy of catheter-based renal denervation in the absence of antihypertensive medications (SPYRAL HTN-OFF MED Pivotal): a multicentre, randomised, sham-controlled trial. Lancet 2020; 395:1444–1451.
747. Mahfoud F, Bohm M, Schmieder R, Narkiewicz K, Ewen S, Ruilope L, et al. Effects of renal denervation on kidney function and long-term outcomes: 3-year follow-up from the Global SYMPLICITY Registry. Eur Heart J 2019; 40:3474–3482.
748. Barbato E, Azizi M, Schmieder RE, Lauder L, Bohm M, Brouwers S, et al. Renal Denervation in the Management of Hypertension in Adults. A Clinical Consensus Statement of the ESC Council on Hypertension and the European Association of Cardiovascular Interventions (EAPCI). Eur Heart J 2023; 44:1313–1330.
749. Pio-Abreu A, Drager LF. Resistant hypertension: time to consider the best fifth anti-hypertensive treatment. Curr Hypertens Rep 2018; 20:67.
750. Desai R, Park H, Brown JD, Mohandas R, Pepine CJ, Smith SM. Comparative safety and effectiveness of aldosterone antagonists versus beta-blockers as fourth agents in patients with apparent resistant hypertension. Hypertension (Dallas, Tex: 1979) 2022; 79:2305–2315.
751. Sinnott SJ, Smeeth L, Williamson E, Perel P, Nitsch D, Tomlinson LA, et al. The comparative effectiveness of fourth-line drugs in resistant hypertension: An application in electronic health record data. Pharmacoepidemiol Drug Saf 2019; 28:1267–1277.
752. Hu X, Fan J, Chen S, Yin Y, Zrenner B. The role of continuous positive airway pressure in blood pressure control for patients with obstructive sleep apnea and hypertension: a meta-analysis of randomized controlled trials. J Clin Hypertens (Greenwich) 2015; 17:215–222.
753. Labarca G, Schmidt A, Dreyse J, Jorquera J, Enos D, Torres G, et al. Efficacy of continuous positive airway pressure (CPAP) in patients with obstructive sleep apnea (OSA) and resistant hypertension (RH): Systematic review and meta-analysis. Sleep Med Rev 2021; 58:101446.
754. Martinez-Garcia MA, Pengo MF. Clinical phenotype of resistant hypertension responders to continuous positive airway pressure treatment: results from the HIPARCO randomized clinical trial. Hypertension (Dallas, Tex: 1979) 2021; 78:559–561.
755. Shi Q, Wang Y, Hao Q, Vandvik PO, Guyatt G, Li J, et al. Pharmacotherapy for adults with overweight and obesity: a systematic review and network meta-analysis of randomised controlled trials. Lancet 2022; 399:259–269.
756. Ferdinand KC, White WB, Calhoun DA, Lonn EM, Sager PT, Brunelle R, et al. Effects of the once-weekly glucagon-like peptide-1 receptor agonist dulaglutide on ambulatory blood pressure and heart rate in patients with type 2 diabetes mellitus. Hypertension (Dallas, Tex: 1979) 2014; 64:731–737.
757. Maringwa J, Sardu ML, Hang Y, Czerniak R, Vishnubhotla M, Vakilynejad M, et al. Characterizing Effects of Antidiabetic Drugs on Heart Rate, Systolic and Diastolic Blood Pressure. Clin Pharmacol Ther 2021; 109:1583–1592.
758. Giugliano D, Scappaticcio L, Longo M, Caruso P, Maiorino MI, Bellastella G, et al. GLP-1 receptor agonists and cardiorenal outcomes in type 2 diabetes: an updated meta-analysis of eight CVOTs. Cardiovasc Diabetol 2021; 20:189.
759. Draznin B, Aroda VR, Bakris G, Benson G, et al. American Diabetes Association Professional Practice C. 8. obesity and weight management for the prevention and treatment of type 2 diabetes: standards of medical care in diabetes-2022. Diabetes Care 2022; 45:S113–S124.
760. Schiavon CA, Ikeoka D, Santucci EV, Santos RN, Damiani LP, Bueno PT, et al. Effects of bariatric surgery versus medical therapy on the 24-hour ambulatory blood pressure and the prevalence of resistant hypertension. Hypertension (Dallas, Tex: 1979) 2019; 73:571–577.
761. Georgianos PI, Agarwal R. Ambulatory blood pressure reduction with SGLT-2 inhibitors: dose-response meta-analysis and comparative evaluation with low-dose hydrochlorothiazide. Diabetes Care 2019; 42:693–700.
762. Solomon SD, McMurray JJV, Anand IS, Ge J, Lam CSP, Maggioni AP, et al. Angiotensin-neprilysin inhibition in heart failure with preserved ejection fraction. N Engl J Med 2019; 381:1609–1620.
763. Jackson AM, Jhund PS, Anand IS, Düngen HD, Lam CSP, Lefkowitz MP, et al. Sacubitril-valsartan as a treatment for apparent resistant hypertension in patients with heart failure and preserved ejection fraction. Eur Heart J 2021; 42:3741–3752.
764. Mancia G, Grassi G. The autonomic nervous system and hypertension. Circ Res 2014; 114:1804-1814
765. Schmieder RE. Renal denervation in patients with chronic kidney disease: current evidence and future perspectives. Nephrol Dial Transplant 2023; 38:1089–1096.
766. DiBona GF, Esler M. Translational medicine: the antihypertensive effect of renal denervation. Am J Physiol Regul Integr Comp Physiol 2010; 298:R245–253.
767. DiBona GF. Neural control of the kidney: past, present, and future. Hypertension (Dallas, Tex: 1979) 2003; 41:621–624.
768. Stella A, Zanchetti A. Functional role of renal afferents. Physiol Rev 1991; 71:659–682.
769. Schmieder RE, Mahfoud F, Mancia G, Azizi M, Bohm M, Dimitriadis K, et al. European Society of Hypertension position paper on renal denervation 2021. J Hypertens 2021; 39:1733–1741.
770. Grassi G, Seravalle G, Brambilla G, Trabattoni D, Cuspidi C, Corso R, et al. Blood pressure responses to renal denervation precede and are independent of the sympathetic and baroreflex effects. Hypertension (Dallas, Tex: 1979) 2015; 65:1209–1216.
771. Hering D, Lambert EA, Marusic P, Walton AS, Krum H, Lambert GW, et al. Substantial reduction in single sympathetic nerve firing after renal denervation in patients with resistant hypertension. Hypertension (Dallas, Tex: 1979) 2013; 61:457–464.
772. Esler M. Illusions of truths in the Symplicity HTN-3 trial: generic design strengths but neuroscience failings. J Am Soc Hypertens 2014; 8:593–598.
773. Brinkmann J, Heusser K, Schmidt BM, Menne J, Klein G, Bauersachs J, et al. Catheter-based renal nerve ablation and centrally generated sympathetic activity in difficult-to-control hypertensive patients: prospective case series. Hypertension (Dallas, Tex: 1979) 2012; 60:1485–1490.
774. Biffi A, Dell’Oro R, Quarti-Trevano F, Cuspidi C, Corrao G, Mancia G, et al. Effects of renal denervation on sympathetic nerve traffic and correlates in drug-resistant and uncontrolled hypertension: a systematic review and meta-analysis. Hypertension (Dallas, Tex: 1979) 2023; 80:659–667.
775. Lauder L, Azizi M, Kirtane AJ, Bohm M, Mahfoud F. Device-based therapies for arterial hypertension. Nat Rev Cardiol 2020; 17:614–628.
776. Bhatt DL, Kandzari DE, O’Neill WW, D’Agostino R, Flack JM, Katzen BT, et al. A controlled trial of renal denervation for resistant hypertension. N Engl J Med 2014; 370:1393–1401.
777. Kandzari DE, Bhatt DL, Brar S, Devireddy CM, Esler M, Fahy M, et al. Predictors of blood pressure response in the SYMPLICITY HTN-3 trial. Eur Heart J 2015; 36:219–227.
778. Mahfoud F, Schmieder RE, Azizi M, Pathak A, Sievert H, Tsioufis C, et al. Proceedings from the 2nd European Clinical Consensus Conference for device-based therapies for hypertension: state of the art and considerations for the future. Eur Heart J 2017; 38:3272–3281.
779. Jordan J, Biaggioni I, Kotsis V, Nilsson P, Grassi G, Fedorowski A, et al. Consensus statement on the definition of orthostatic hypertension endorsed by the American Autonomic Society and the Japanese Society of Hypertension. Clin Auton Res 2022.
780. Townsend RR, Mahfoud F, Kandzari DE, Kario K, Pocock S, Weber MA, et al. Catheter-based renal denervation in patients with uncontrolled hypertension in the absence of antihypertensive medications (SPYRAL HTN-OFF MED): a randomised, sham-controlled, proof-of-concept trial. Lancet 2017; 390:2160–2170.
781. Azizi M, Mahfoud F, Weber MA, Sharp ASP, Schmieder RE, Lurz P, et al. Effects of renal denervation vs sham in resistant hypertension after medication escalation: prespecified analysis at 6 months of the RADIANCE-HTN TRIO randomized clinical trial. JAMA Cardiol 2022; 7:1244–1252.
782. Kandzari DE, Bohm M, Mahfoud F, Townsend RR, Weber MA, Pocock S, et al. Effect of renal denervation on blood pressure in the presence of antihypertensive drugs: 6-month efficacy and safety results from the SPYRAL HTN-ON MED proof-of-concept randomised trial. Lancet 2018; 391:2346–2355.
783. Kirtane A, Sharp A, Mahfoud F, Fisher N, Schmieder RE, Daemen J, et al. Patient-level pooled analysis of ultrasound renal denervation in the sham-controlled RADIANCE II, RADIANCE-HTN SOLO and RADIANCE-HTN TRIO trials. JAMA Cardiology 2023; 8:464–473.
784. Rosa J, Widimsky P, Tousek P, Petrak O, Curila K, Waldauf P, et al. Randomized comparison of renal denervation versus intensified pharmacotherapy including spironolactone in true-resistant hypertension: six-month results from the Prague-15 study. Hypertension (Dallas, Tex: 1979) 2015; 65:407–413.
785. Kario K, Yokoi Y, Okamura K, Fujihara M, Ogoyama Y, Yamamoto E, et al. Catheter-based ultrasound renal denervation in patients with resistant hypertension: the randomized, controlled REQUIRE trial. Hypertens Res 2022; 45:221–231.
786. Townsend RR, Walton A, Hettrick DA, Hickey GL, Weil J, Sharp ASP, et al. Review and meta-analysis of renal artery damage following percutaneous renal denervation with radiofrequency renal artery ablation. EuroIntervention 2020; 16:89–96.
787. Sanders MF, Reitsma JB, Morpey M, Gremmels H, Bots ML, Pisano A, et al. Renal safety of catheter-based renal denervation: systematic review and meta-analysis. Nephrol Dial Transplant 2017; 32:1440–1447.
788. Rader F, Kirtane AJ, Wang Y, Daemen J, Lurz P, Sayer J, et al. Durability of blood pressure reduction after ultrasound renal denervation: three-year follow-up of the treatment arm of the randomised RADIANCE-HTN SOLO trial. EuroIntervention 2022; 18:e677–e685.
789. Mahfoud F, Kandzari DE, Kario K, Townsend RR, Weber MA, Schmieder RE, et al. Long-term efficacy and safety of renal denervation in the presence of antihypertensive drugs (SPYRAL HTN-ON MED): a randomised, sham-controlled trial. Lancet 2022; 399:1401–1410.
790. Bhatt DL, Vaduganathan M, Kandzari DE, Leon MB, Rocha-Singh K, Townsend RR, et al. Long-term outcomes after catheter-based renal artery denervation for resistant hypertension: final follow-up of the randomised SYMPLICITY HTN-3 Trial. Lancet 2022; 400:1405–1416.
791. Kjeldsen SE, Burnier M, Narkiewicz K, Kreutz R, Mancia G. Key questions regarding the SYMPLICITY HTN-3 trial. Lancet 2023; 401:1336–1337.
792. Persu A, Stoenoiu, M.S., Maes, F., Kreutz, R., Mancia, G., Kjeldsen, S.E.. Late outcomes of renal denervation are more favorable than early ones. Facts or fancies? Clin Kid J 2023; in press.
793. Schmieder RE, Mahfoud F, Mancia G, Narkiewicz K, Ruilope L, Hutton D, et al. Clinical Event Reductions in High-Risk Patients After Renal Denervation Projected from the Global SYMPLICITY Registry. Eur Heart J Qual Care Clin Outcomes 2022.
794. Schmieder RE, Hogerl K, Jung S, Bramlage P, Veelken R, Ott C. Patient preference for therapies in hypertension: a cross-sectional survey of German patients. Clin Res Cardiol 2019; 108:1331–1342.
795. Mahfoud F, Schlaich MP, Lobo MD. Device therapy of hypertension. Circ Res 2021; 128:1080–1099.
796. Heusser K, Tank J, Engeli S, Diedrich A, Menne J, Eckert S, et al. Carotid baroreceptor stimulation, sympathetic activity, baroreflex function, and blood pressure in hypertensive patients. Hypertension (Dallas, Tex: 1979) 2010; 55:619–626.
797. Gronda E, Seravalle G, Brambilla G, Costantino G, Casini A, Alsheraei A, et al. Chronic baroreflex activation effects on sympathetic nerve traffic, baroreflex function, and cardiac haemodynamics in heart failure: a proof-of-concept study. Eur J Heart Fail 2014; 16:977–983.
798. Bisognano JD, Bakris G, Nadim MK, Sanchez L, Kroon AA, Schafer J, et al. Baroreflex activation therapy lowers blood pressure in patients with resistant hypertension: results from the double-blind, randomized, placebo-controlled rheos pivotal trial. J Am Coll Cardiol 2011; 58:765–773.
799. Spiering W, Williams B, Van der Heyden J, van Kleef M, Lo R, Versmissen J, et al. Endovascular baroreflex amplification for resistant hypertension: a safety and proof-of-principle clinical study. Lancet 2017; 390:2655–2661.
800. van Kleef M, Devireddy CM, van der Heyden J, Bates MC, Bakris GL, Stone GW, et al. Treatment of resistant hypertension with endovascular baroreflex amplification: 3-year results from the CALM-FIM study. JACC Cardiovasc Interv 2022; 15:321–332.
801. Lobo MD, Sobotka PA, Stanton A, Cockcroft JR, Sulke N, Dolan E, et al. Central arteriovenous anastomosis for the treatment of patients with uncontrolled hypertension (the ROX CONTROL HTN study): a randomised controlled trial. Lancet 2015; 385:1634–1641.
802. Neuzil P, Merkely B, Erglis A, Marinskis G, de Groot JR, Schmidinger H, et al. Pacemaker-mediated programmable hypertension control therapy. J Am Heart Assoc 2017; 6:
803. Kalarus Z, Merkely B, Neužil P, Grabowski M, Mitkowski P, Marinskis G, et al. Pacemaker-based cardiac neuromodulation therapy in patients with hypertension: a pilot study. J Am Heart Assoc 2021; 10:e020492.
804. Mancia G, Bertinieri G, Grassi G, Parati G, Pomidossi G, Ferrari A, et al. Effects of blood-pressure measurement by the doctor on patient's blood pressure and heart rate. Lancet 1983; 2:695–698.
805. Mancia G, Parati G, Pomidossi G, Grassi G, Casadei R, Zanchetti A. Alerting reaction and rise in blood pressure during measurement by physician and nurse. Hypertension (Dallas, Tex: 1979) 1987; 9:209–215.
806. Palatini P, Palomba D, Bertolo O, Minghetti R, Longo D, Sarlo M, et al. The white-coat effect is unrelated to the difference between clinic and daytime blood pressure and is associated with greater reactivity to public speaking. J Hypertens 2003; 21:545–553.
807. Parati G, Omboni S, Staessen J, Thijs L, Fagard R, Ulian L, et al. Limitations of the difference between clinic and daytime blood pressure as a surrogate measure of the ‘white-coat’ effect. Syst-Eur investigators. J Hypertens 1998; 16:23–29.
808. Grassi G, Pisano A, Bolignano D, Seravalle G, D’Arrigo G, Quarti-Trevano F, et al. Sympathetic nerve traffic activation in essential hypertension and its correlates: systematic reviews and meta-analyses. Hypertension (Dallas, Tex: 1979) 2018; 72:483–491.
809. Sega R, Trocino G, Lanzarotti A, Carugo S, Cesana G, Schiavina R, et al. Alterations of cardiac structure in patients with isolated office, ambulatory, or home hypertension: Data from the general population (Pressione Arteriose Monitorate E Loro Associazioni [PAMELA] Study). Circulation 2001; 104:1385–1392.
810. Briasoulis A, Androulakis E, Palla M, Papageorgiou N, Tousoulis D. White-coat hypertension and cardiovascular events: a meta-analysis. J Hypertens 2016; 34:593–599.
811. Huang Y, Huang W, Mai W, Cai X, An D, Liu Z, et al. White-coat hypertension is a risk factor for cardiovascular diseases and total mortality. J Hypertens 2017; 35:677–688.
812. Mancia G, Bombelli M, Facchetti R, Madotto F, Quarti-Trevano F, Polo Friz H, et al. Long-term risk of sustained hypertension in white-coat or masked hypertension. Hypertension (Dallas, Tex: 1979) 2009; 54:226–232.
813. Mancia G, Bombelli M, Facchetti R, Madotto F, Quarti-Trevano F, Grassi G, et al. Increased long-term risk of new-onset diabetes mellitus in white-coat and masked hypertension. J Hypertens 2009; 27:1672–1678.
814. Mancia G, Facchetti R, Bombelli M, Cuspidi C, Grassi G. White-coat hypertension: pathophysiological and clinical aspects: excellence award for hypertension research 2020. Hypertension (Dallas, Tex: 1979) 2021; 78:1677–1688.
815. Antza C, Vazakidis P, Doundoulakis I, Bouras E, Haidich AB, Stabouli S, et al. Masked and white coat hypertension, the double trouble of large arteries: A systematic review and meta-analysis. J Clin Hypertens (Greenwich) 2020; 22:802–811.
816. Asayama K, Li Y, Franklin SS, Thijs L, O’Brien E, Staessen JA. Cardiovascular risk associated with white-coat hypertension: con side of the argument. Hypertension (Dallas, Tex: 1979) 2017; 70:676–682.
817. Cohen JB, Denker MG, Cohen DL, Townsend RR. Cardiovascular events and mortality in white coat hypertension. Ann Intern Med 2019; 171:603–604.
818. Puato M, Palatini P, Zanardo M, Dorigatti F, Tirrito C, Rattazzi M, et al. Increase in carotid intima-media thickness in grade I hypertensive subjects: white-coat versus sustained hypertension. Hypertension (Dallas, Tex: 1979) 2008; 51:1300–1305.
819. Bulpitt CJ, Beckett N, Peters R, Staessen JA, Wang JG, Comsa M, et al. Does white coat hypertension require treatment over age 80?: Results of the hypertension in the very elderly trial ambulatory blood pressure side project. Hypertension (Dallas, Tex: 1979) 2013; 61:89–94.
820. Mancia G, Facchetti R, Parati G, Zanchetti A. Effect of long-term antihypertensive treatment on white-coat hypertension. Hypertension (Dallas, Tex: 1979) 2014; 64:1388–1398.
821. Mancia G, Bombelli M, Cuspidi C, Facchetti R, Grassi G. Cardiovascular risk associated with white-coat hypertension: pro side of the argument. Hypertension (Dallas, Tex: 1979) 2017; 70:668–675.
822. Anstey DE, Pugliese D, Abdalla M, Bello NA, Givens R, Shimbo D. An update on masked hypertension. Curr Hypertens Rep 2017; 19:94.
823. Melgarejo JD, Maestre GE, Thijs L, Asayama K, Boggia J, Casiglia E, et al. Prevalence, treatment, and control rates of conventional and ambulatory hypertension across 10 populations in 3 Continents. Hypertension (Dallas, Tex: 1979) 2017; 70:50–58.
824. Hung MH, Shih LC, Wang YC, Leu HB, Huang PH, Wu TC, et al. Prediction of masked hypertension and masked uncontrolled hypertension using machine learning. Front Cardiovasc Med 2021; 8:778306.
825. Shimbo D, Abdalla M, Falzon L, Townsend RR, Muntner P. Studies comparing ambulatory blood pressure and home blood pressure on cardiovascular disease and mortality outcomes: a systematic review. J Am Soc Hypertens 2016; 10:224–234. e217.
826. Palatini P, Winnicki M, Santonastaso M, Mos L, Longo D, Zaetta V, et al. Prevalence and clinical significance of isolated ambulatory hypertension in young subjects screened for stage 1 hypertension. Hypertension (Dallas, Tex: 1979) 2004; 44:170–174.
827. Cuspidi C, Facchetti R, Quarti-Trevano F, Dell’Oro R, Tadic M, Grassi G, et al. Left ventricular hypertrophy in isolated and dual masked hypertension. J Clin Hypertens (Greenwich) 2020; 22:673–677.
828. Cuspidi C, Sala C, Tadic M, Rescaldani M, Grassi G, Mancia G. Untreated masked hypertension and subclinical cardiac damage: a systematic review and meta-analysis. Am J Hypertens 2015; 28:806–813.
829. Cuspidi C, Facchetti R, Bombelli M, Sala C, Tadic M, Grassi G, et al. Risk of new-onset metabolic syndrome associated with white-coat and masked hypertension: data from a general population. J Hypertens 2018; 36:1833–1839.
830. Grassi G, Seravalle G, Trevano FQ, Dell’oro R, Bolla G, Cuspidi C, et al. Neurogenic abnormalities in masked hypertension. Hypertension (Dallas, Tex: 1979) 2007; 50:537–542.
831. Palatini P, Mos L, Rattazzi M, Spinella P, Ermolao A, Vriz O, et al. Blood pressure response to standing is a strong determinant of masked hypertension in young to middle-age individuals. J Hypertens 2022; 40:1927–1934.
832. Palla M, Saber H, Konda S, Briasoulis A. Masked hypertension and cardiovascular outcomes: an updated systematic review and meta-analysis. Integr Blood Press Control 2018; 11:11–24.
833. Zhang DY, Guo QH, An DW, Li Y, Wang JG. A comparative meta-analysis of prospective observational studies on masked hypertension and masked uncontrolled hypertension defined by ambulatory and home blood pressure. J Hypertens 2019; 37:1775–1785.
834. Tientcheu D, Ayers C, Das SR, McGuire DK, de Lemos JA, Khera A, et al. Target organ complications and cardiovascular events associated with masked hypertension and white-coat hypertension: analysis from the Dallas heart study. J Am Coll Cardiol 2015; 66:2159–2169.
835. Lurbe E, Redon J, Kesani A, Pascual JM, Tacons J, Alvarez V, et al. Increase in nocturnal blood pressure and progression to microalbuminuria in type 1 diabetes. N Engl J Med 2002; 347:797–805.
836. Antza C, Farmakis I, Doundoulakis I, Akrivos E, Stalikas N, Zafeiropoulos S, et al. Reproducibility of masked hypertension and office-based hypertension: a systematic review and meta-analysis. J Hypertens 2022; 40:1053–1059.
837. Paiva AMG, Gomes M, Campana É MG, Feitosa ADM, Sposito AC, Mota-Gomes MA, et al. Impact of hypertension phenotypes on the office and 24-h pulse wave velocity and augmentation index in individuals with or without antihypertensive medication use. Hypertens Res 2019; 42:1989–1995.
838. Bobrie G, Chatellier G, Genes N, Clerson P, Vaur L, Vaisse B, et al. Cardiovascular prognosis of “masked hypertension” detected by blood pressure self-measurement in elderly treated hypertensive patients. Jama 2004; 291:1342–1349.
839. Franklin SS, Thijs L, Li Y, Hansen TW, Boggia J, Liu Y, et al. Masked hypertension in diabetes mellitus: treatment implications for clinical practice. Hypertension (Dallas, Tex: 1979) 2013; 61:964–971.
840. Ghazi L, Cohen LP, Muntner P, Shimbo D, Drawz PE. Effects of intensive versus standard office-based hypertension treatment strategy on white-coat effect and masked uncontrolled hypertension: from the SPRINT ABPM Ancillary study. Hypertension (Dallas, Tex: 1979) 2020; 76:1090–1096.
841. Kim HJ, Shin JH, Lee Y, Kim JH, Hwang SH, Kim WS, et al. Clinical features and predictors of masked uncontrolled hypertension from the Korean Ambulatory Blood Pressure Monitoring Registry. Korean J Intern Med 2021; 36:1102–1114.
842. Fu X, Ren H, Xie J, Wang W, Li Y, Gao P, et al. Association of nighttime masked uncontrolled hypertension with left ventricular hypertrophy and kidney function among patients with chronic kidney disease not receiving dialysis. JAMA Netw Open 2022; 5:e2214460.
843. Pierdomenico SD, Pierdomenico AM, Coccina F, Clement DL, De Buyzere ML, De Bacquer DA, et al. Prognostic value of masked uncontrolled hypertension. Hypertension (Dallas, Tex: 1979) 2018; 72:862–869.
844. Mancia G, Facchetti R, Vanoli J, Dolfini V, Grassi G. Reproducibility of blood pressure phenotypes identified by office and ambulatory blood pressure in treated hypertensive patients. Data from the PHYLLIS study. Hypertens Res 2022; 45:1599–1608.
845. Siddiqui M, Judd EK, Jaeger BC, Bhatt H, Dudenbostel T, Zhang B, et al. Out-of-clinic sympathetic activity is increased in patients with masked uncontrolled hypertension. Hypertension (Dallas, Tex: 1979) 2019; 73:132–141.
846. Siddiqui M, Judd EK, Dudenbostel T, Zhang B, Gupta P, Tomaszewski M, et al. Masked uncontrolled hypertension is not attributable to medication nonadherence. Hypertension (Dallas, Tex: 1979) 2019; 74:652–659.
847. Grassi G, Mancia G. Masked uncontrolled hypertension. Hypertension (Dallas, Tex: 1979) 2019; 73:39–41.
848. McEniery CM, Yasmin, Wallace S, Maki-Petaja K, McDonnell B, Sharman JE, et al. Increased stroke volume and aortic stiffness contribute to isolated systolic hypertension in young adults. Hypertension (Dallas, Tex: 1979) 2005; 46:221–226.
849. Palatini P. Regular physical activity: a major component of isolated systolic hypertension in the young. Minerva Med 2022; 113:798–806.
850. Grebla RC, Rodriguez CJ, Borrell LN, Pickering TG. Prevalence and determinants of isolated systolic hypertension among young adults: the 1999–2004 US National Health And Nutrition Examination Survey. J Hypertens 2010; 28:15–23.
851. O’Rourke MF, Vlachopoulos C, Graham RM. Spurious systolic hypertension in youth. Vasc Med 2000; 5:141–145.
852. Mahmud A, Feely J. Spurious systolic hypertension of youth: fit young men with elastic arteries. Am J Hypertens 2003; 16:229–232.
853. Eeftinck Schattenkerk DW, van Gorp J, Vogt L, Peters RJ, van den Born BH. Isolated systolic hypertension of the young and its association with central blood pressure in a large multi-ethnic population. The HELIUS study. Eur J Prev Cardiol 2018; 25:1351–1359.
854. Saladini F, Fania C, Mos L, Mazzer A, Casiglia E, Palatini P. Office pulse pressure is a predictor of favorable outcome in young- to middle-aged subjects with stage 1 hypertension. Hypertension (Dallas, Tex: 1979) 2017.
855. Yano Y, Stamler J, Garside D B, Daviglus M L, Franklin S S, Carnethon M R, et al. Isolated systolic hypertension in young and middle-aged adults and 31-year risk for cardiovascular mortality: the Chicago Heart Association Detection Project in Industry study. J Am Coll Cardiol 2015; 65:327–335.
856. Atasoy S, Middeke M, Johar H, Peters A, Heier M, Ladwig K-H. Cardiovascular mortality risk in young adults with isolated systolic hypertension: findings from population-based MONICA/KORA cohort study. Journal of Human Hypertension 2022; 36:1059–1065.
857. Lee H, Yano Y, Cho SMJ, Park JH, Park S, Lloyd-Jones DM, et al. Cardiovascular Risk of Isolated Systolic or Diastolic Hypertension in Young Adults. Circulation 2020; 141:1778–1786.
858. Palatini P, Saladini F, Mos L, Fania C, Mazzer A, Casiglia E. Clinical characteristics and risk of hypertension needing treatment in young patients with systolic hypertension identified with ambulatory monitoring. J Hypertens 2018; 36:1810–1815.
859. Saladini F, Santonastaso M, Mos L, Benetti E, Zanatta N, Maraglino G, et al. Isolated systolic hypertension of young-to-middle-age individuals implies a relatively low risk of developing hypertension needing treatment when central blood pressure is low. J Hypertens 2011; 29:1311–1319.
860. Mancia G, Giannattasio C. Diagnostic and therapeutic problems of isolated systolic hypertension. J Hypertens 2015; 33:33–43.
861. Cheng S, Xanthakis V, Sullivan LM, Vasan RS. Blood pressure tracking over the adult life course: patterns and correlates in the Framingham heart study. Hypertension (Dallas, Tex: 1979) 2012; 60:1393–1399.
862. Wills AK, Lawlor DA, Matthews FE, Sayer AA, Bakra E, Ben-Shlomo Y, et al. Life course trajectories of systolic blood pressure using longitudinal data from eight UK cohorts. PLoS Med 2011; 8:e1000440.
863. Liu X, Rodriguez CJ, Wang K. Prevalence and trends of isolated systolic hypertension among untreated adults in the United States. J Am Soc Hypertens 2015; 9:197–205.
864. Franklin SS, Larson MG, Khan SA, Wong ND, Leip EP, Kannel WB, et al. Does the relation of blood pressure to coronary heart disease risk change with aging? The Framingham Heart Study. Circulation 2001; 103:1245–1249.
865. Flint AC, Conell C, Ren X, Banki NM, Chan SL, Rao VA, et al. Effect of systolic and diastolic blood pressure on cardiovascular outcomes. N Engl J Med 2019; 381:243–251.
866. Bavishi C, Goel S, Messerli FH. Isolated systolic hypertension: an update after SPRINT. Am J Med 2016; 129:1251–1258.
867. Li Y, Wei FF, Thijs L, Boggia J, Asayama K, Hansen TW, et al. Ambulatory hypertension subtypes and 24-hour systolic and diastolic blood pressure as distinct outcome predictors in 8341 untreated people recruited from 12 populations. Circulation 2014; 130:466–474.
868. O’Donnell CJ, Ridker PM, Glynn RJ, Berger K, Ajani U, Manson JE, et al. Hypertension and borderline isolated systolic hypertension increase risks of cardiovascular disease and mortality in male physicians. Circulation 1997; 95:1132–1137.
869. Kostis WJ, Sargsyan D, Mekkaoui C, Moreyra AE, Cabrera J, Cosgrove NM, et al. Association of orthostatic hypertension with mortality in the Systolic Hypertension in the Elderly Program. J Hum Hypertens 2019; 33:735–740.
870. Liu L, Wang JG, Gong L, Liu G, Staessen JA. Comparison of active treatment and placebo in older Chinese patients with isolated systolic hypertension. Systolic Hypertension in China (Syst-China) Collaborative Group. J Hypertens 1998; 16:1823–1829.
871. Staessen JA, Gasowski J, Wang JG, Thijs L, Den Hond E, Boissel JP, et al. Risks of untreated and treated isolated systolic hypertension in the elderly: meta-analysis of outcome trials. Lancet 2000; 355:865–872.
872. Ogihara T, Saruta T, Rakugi H, Matsuoka H, Shimamoto K, Shimada K, et al. Target blood pressure for treatment of isolated systolic hypertension in the elderly: valsartan in elderly isolated systolic hypertension study. Hypertension (Dallas, Tex: 1979) 2010; 56:196–202.
873. Sobieraj P, Lewandowski J, Siński M, Gaciong Z. Low on-treatment diastolic blood pressure and cardiovascular outcome: A post-hoc analysis using NHLBI SPRINT Research Materials. Sci Rep 2019; 9:13070.
874. Franklin SS, Chow VH, Mori AD, Wong ND. The significance of low DBP in US adults with isolated systolic hypertension. J Hypertens 2011; 29:1101–1108.
875. Mahajan S, Zhang D, He S, Lu Y, Gupta A, Spatz ES, et al. Prevalence, Awareness, and Treatment of Isolated Diastolic Hypertension: Insights From the China PEACE Million Persons Project. J Am Heart Assoc 2019; 8:e012954.
876. Yu S, Zhang Y. The association between isolated systolic or diastolic hypertension and cardiovascular risk. J Hypertens 2021; 39:1552–1554.
877. Romero CA, Tabares AH, Orias M. Is Isolated Diastolic Hypertension an Important Phenotype? Curr Cardiol Rep 2021; 23:177.
878. Franklin SS, Jacobs MJ, Wong ND, L’Italien GJ, Lapuerta P. Predominance of isolated systolic hypertension among middle-aged and elderly US hypertensives: analysis based on National Health and Nutrition Examination Survey (NHANES) III. Hypertension (Dallas, Tex: 1979) 2001; 37:869–874.
879. Xie K, Gao X, Bao L, Shan Y, Shi H, Li Y. The different risk factors for isolated diastolic hypertension and isolated systolic hypertension: a national survey. BMC Public Health 2021; 21:1672.
880. Jacobsen AP, Al Rifai M, Arps K, Whelton SP, Budoff MJ, Nasir K, et al. A cohort study and meta-analysis of isolated diastolic hypertension: searching for a threshold to guide treatment. Eur Heart J 2021; 42:2119–2129.
881. Saladini F, Dorigatti F, Santonastaso M, Mos L, Ragazzo F, Bortolazzi A, et al. Natural history of hypertension subtypes in young and middle-age adults. Am J Hypertens 2009; 22:531–537.
882. Li FR, He Y, Yang HL, Liu HM, Zhou R, Chen GC, et al. Isolated systolic and diastolic hypertension by the 2017 American College of Cardiology/American Heart Association guidelines and risk of cardiovascular disease: a large prospective cohort study. J Hypertens 2021; 39:1594–1601.
883. McGrath BP, Kundu P, Daya N, Coresh J, Selvin E, McEvoy JW, et al. Isolated diastolic hypertension in the UK Biobank: comparison of ACC/AHA and ESC/NICE guideline definitions. Hypertension (Dallas, Tex: 1979) 2020; 76:699–706.
884. Yue L, Chen H, Sun Q, Shi L, Sun J, Li G, et al. Prevalence of isolated diastolic hypertension and the risk of cardiovascular mortality among adults aged 40 years and older in northeast China: a prospective cohort study. BMJ Open 2022; 12:e061762.
885. Vishram-Nielsen JKK, Kristensen AMD, Pareek M, Laurent S, Nilsson PM, Linneberg A, et al. Predictive importance of blood pressure characteristics with increasing age in healthy men and women: the MORGAM Project. Hypertension (Dallas, Tex: 1979) 2021; 77:1076–1085.
886. Huang M, Long L, Tan L, Shen A, Deng M, Peng Y, et al. Isolated diastolic hypertension and risk of cardiovascular events: a systematic review and meta-analysis of cohort studies with 489,814 participants. Front Cardiovasc Med 2021; 8:810105.
887. McEvoy JW, Yang WY, Thijs L, Zhang ZY, Melgarejo JD, Boggia J, et al. Isolated diastolic hypertension in the IDACO Study: an age-stratified analysis using 24-hour ambulatory blood pressure measurements. Hypertension (Dallas, Tex: 1979) 2021; 78:1222–1231.
888. MacMahon SW, Cutler JA, Furberg CD, Payne GH. The effects of drug treatment for hypertension on morbidity and mortality from cardiovascular disease: a review of randomized controlled trials. Prog Cardiovasc Dis 1986; 29:99–118.
889. Cuspidi C, Tadic M, Sala C. Targeting nocturnal hypertension: the emerging role of home blood pressure. Am J Hypertens 2019; 32:727–729.
890. Fujiwara T, Hoshide S, Kanegae H, Kario K. Cardiovascular event risks associated with masked nocturnal hypertension defined by home blood pressure monitoring in the J-HOP Nocturnal blood pressure study. Hypertension (Dallas, Tex: 1979) 2020; 76:259–266.
891. Salles GF, Reboldi G, Fagard RH, Cardoso CR, Pierdomenico SD, Verdecchia P, et al. Prognostic effect of the nocturnal blood pressure fall in hypertensive patients: the ambulatory blood pressure collaboration in patients with hypertension (ABC-H) meta-Analysis. Hypertension (Dallas, Tex: 1979) 2016; 67:693–700.
892. Yang WY, Melgarejo JD, Thijs L, Zhang ZY, Boggia J, Wei FF, et al. Association of office and ambulatory blood pressure with mortality and cardiovascular outcomes. JAMA 2019; 322:409–420.
893. Cuspidi C, Sala C, Tadic M, Gherbesi E, Grassi G, Mancia G. Nocturnal hypertension and subclinical cardiac and carotid damage: an updated review and meta-analysis of echocardiographic studies. J Clin Hypertens (Greenwich) 2016; 18:913–920.
894. Fan HQ, Li Y, Thijs L, Hansen TW, Boggia J, Kikuya M, et al. Prognostic value of isolated nocturnal hypertension on ambulatory measurement in 8711 individuals from 10 populations. J Hypertens 2010; 28:2036–2045.
895. Li S, Schwartz JE, Shimbo D, Muntner P, Shikany JM, Booth JN 3rd, et al. Estimated prevalence of masked asleep hypertension in US adults. JAMA Cardiol 2021; 6:568–573.
896. Salazar MR, Espeche WG, Balbín E, Leiva Sisnieguez CE, Minetto J, Leiva Sisnieguez BC, et al. Prevalence of isolated nocturnal hypertension according to 2018 European Society of Cardiology and European Society of Hypertension office blood pressure categories. J Hypertens 2020; 38:434–440.
897. Cuspidi C, Paoletti F, Tadic M, Sala C, Gherbesi E, Dell’Oro R, et al. Nocturnal blood pressure: the dark side of white-coat hypertension. J Hypertens 2020; 38:2404–2408.
898. Palatini P, Verdecchia P, Beilin LJ, Eguchi K, Imai Y, Kario K, et al. Association of extreme nocturnal dipping with cardiovascular events strongly depends on age. Hypertension (Dallas, Tex: 1979) 2020; 75:324–330.
899. Omboni S, Parati G, Palatini P, Vanasia A, Muiesan ML, Cuspidi C, et al. Reproducibility and clinical value of nocturnal hypotension: prospective evidence from the SAMPLE study. Study on Ambulatory Monitoring of Pressure and Lisinopril Evaluation. J Hypertens 1998; 16:733–738.
900. Tadic M, Cuspidi C, Grassi G, Mancia G. Isolated nocturnal hypertension: what do we know and what can we do? Integr Blood Press Control 2020; 13:63–69.
901. Parati G, Omboni S, Rizzoni D, Agabiti-Rosei E, Mancia G. The smoothness index: a new, reproducible and clinically relevant measure of the homogeneity of the blood pressure reduction with treatment for hypertension. J Hypertens 1998; 16:1685–1691.
902. Dani M, Dirksen A, Taraborrelli P, Panagopolous D, Torocastro M, Sutton R, et al. Orthostatic hypotension in older people: considerations, diagnosis and management. Clin Med (Lond) 2021; 21:e275–e282.
903. Freeman R, Wieling W, Axelrod FB, Benditt DG, Benarroch E, Biaggioni I, et al. Consensus statement on the definition of orthostatic hypotension, neurally mediated syncope and the postural tachycardia syndrome. Clin Auton Res 2011; 21:69–72.
904. Jordan J, Ricci F, Hoffmann F, Hamrefors V, Fedorowski A. Orthostatic hypertension: critical appraisal of an overlooked condition. Hypertension (Dallas, Tex: 1979) 2020; 75:1151–1158.
905. Hoshide S, Matsui Y, Shibasaki S, Eguchi K, Ishikawa J, Ishikawa S, et al. Orthostatic hypertension detected by self-measured home blood pressure monitoring: a new cardiovascular risk factor for elderly hypertensives. Hypertens Res 2008; 31:1509–1516.
906. Kario K. Orthostatic hypertension-a new haemodynamic cardiovascular risk factor. Nat Rev Nephrol 2013; 9:726–738.
907. Robertson D, Hollister AS, Biaggioni I, Netterville JL, Mosqueda-Garcia R, Robertson RM. The diagnosis and treatment of baroreflex failure. N Engl J Med 1993; 329:1449–1455.
908. Heusser K, Tank J, Luft FC, Jordan J. Baroreflex failure. Hypertension (Dallas, Tex: 1979) 2005; 45:834–839.
909. Biaggioni I, Shibao CA, Diedrich A, Muldowney JAS 3rd, Laffer CL, Jordan J. Blood pressure management in afferent baroreflex failure: JACC review topic of the week. J Am Coll Cardiol 2019; 74:2939–2947.
910. Norcliffe-Kaufmann L, Axelrod F, Kaufmann H. Afferent baroreflex failure in familial dysautonomia. Neurology 2010; 75:1904–1911.
911. Jordan J, Shannon JR, Black B, Costa F, Ertl AC, Furlan R, et al. Malignant vagotonia due to selective baroreflex failure. Hypertension (Dallas, Tex: 1979) 1997; 30:1072–1077.
912. Fanciulli A, Jordan J, Biaggioni I, Calandra-Buonaura G, Cheshire WP, Cortelli P, et al. Consensus statement on the definition of neurogenic supine hypertension in cardiovascular autonomic failure by the American Autonomic Society (AAS) and the European Federation of Autonomic Societies (EFAS): Endorsed by the European Academy of Neurology (EAN) and the European Society of Hypertension (ESH). Clin Auton Res 2018; 28:355–362.
913. Golden EP, Vernino S. Autoimmune autonomic neuropathies and ganglionopathies: epidemiology, pathophysiology, and therapeutic advances. Clin Auton Res 2019; 29:277–288.
914. Juraschek SP, Daya N, Appel LJ, Miller ER 3rd, McEvoy JW, Matsushita K, et al. Orthostatic hypotension and risk of clinical and subclinical cardiovascular disease in middle-aged adults. J Am Heart Assoc 2018; 7:
915. Wieling W, Kaufmann H, Claydon VE, van Wijnen VK, Harms MPM, Juraschek SP, et al. Diagnosis and treatment of orthostatic hypotension. Lancet Neurol 2022; 21:735–746.
916. Jordan J, Fanciulli A, Tank J, Calandra-Buonaura G, Cheshire WP, Cortelli P, et al. Management of supine hypertension in patients with neurogenic orthostatic hypotension: scientific statement of the American Autonomic Society, European Federation of Autonomic Societies, and the European Society of Hypertension. J Hypertens 2019; 37:1541–1546.
917. Lurbe E, Agabiti-Rosei E, Cruickshank JK, Dominiczak A, Erdine S, Hirth A, et al. 2016 European Society of Hypertension guidelines for the management of high blood pressure in children and adolescents. J Hypertens 2016; 34:1887–1920.
918. Falkner B, Lurbe E. The USPSTF call to inaction on blood pressure screening in children and adolescents. Pediatr Nephrol 2021; 36:1327–1329.
919. de Simone G, Mancusi C, Hanssen H, Genovesi S, Lurbe E, Parati G, et al. Hypertension in children and adolescents. Eur Heart J 2022; 43:3290–3301.
920. Falkner B, Gidding SS, Baker-Smith CM, Brady TM, Flynn JT, Malle LM, et al. Pediatric primary hypertension: an underrecognized condition: a scientific statement from the american heart association. Hypertension (Dallas, Tex: 1979) 2023; 80:e101–e111.
921. The fourth report on the diagnosis, evaluation, and treatment of high blood pressure in children and adolescents. Pediatrics 2004; 114:555–576.
922. Stergiou GS, Dolan E, Kollias A, Poulter NR, Shennan A, Staessen JA, et al. Blood pressure measurement in special populations and circumstances. J Clin Hypertens (Greenwich) 2018; 20:1122–1127.
923. Lurbe E, Mancia G, Calpe J, Drozdz D, Erdine S, Fernandez-Aranda F, et al. Joint statement for assessing and managing high blood pressure in children and adolescents: Chapter 1. how to correctly measure blood pressure in children and adolescents. Front Pediatr 2023; 11:1140357.
924. Flynn JT, Daniels SR, Hayman LL, Maahs DM, McCrindle BW, Mitsnefes M, et al. Update: ambulatory blood pressure monitoring in children and adolescents: a scientific statement from the American Heart Association. Hypertension (Dallas, Tex: 1979) 2014; 63:1116–1135.
925. Wuhl E, Hadtstein C, Mehls O, Schaefer F, Escape Trial G. Home, clinic, and ambulatory blood pressure monitoring in children with chronic renal failure. Pediatr Res 2004; 55:492–497.
926. Stergiou GS, Bountzona I, Alamara C, Vazeou A, Kollias A, Ntineri A. Reproducibility of office and out-of-office blood pressure measurements in children: implications for clinical practice and research. Hypertension (Dallas, Tex: 1979) 2021; 77:993–1000.
927. Stergiou G, Stambolliu E, Bountzona I, Ntineri A, Kollias A, Vazeou A, et al. Home blood pressure monitoring in children and adolescents: systematic review of evidence on clinical utility. Curr Hypertens Rep 2019; 21:64.
928. Wühl E, Trivelli A, Picca S, Litwin M, Peco-Antic A, Zurowska A, et al. Strict blood-pressure control and progression of renal failure in children. N Engl J Med 2009; 361:1639–1650.
929. Dionne JM, Jiang S, Ng DK, Flynn JT, Mitsnefes MM, Furth SL, et al. Mean arterial pressure and chronic kidney disease progression in the CKiD cohort. Hypertension (Dallas, Tex: 1979) 2021; 78:65–73.
930. Flynn JT, Kaelber DC, Baker-Smith CM, Blowey D, Carroll AE, Daniels SR, et al. Clinical practice guideline for screening and management of high blood pressure in children and adolescents. Pediatrics 2017; 140.
931. Rabi DM, McBrien KA, Sapir-Pichhadze R, Nakhla M, Ahmed SB, Dumanski SM, et al. Hypertension Canada's 2020 comprehensive guidelines for the prevention, diagnosis, risk assessment, and treatment of hypertension in adults and children. Can J Cardiol 2020; 36:596–624.
932. Lee H, Yano Y, Cho SMJ, Heo JE, Kim DW, Park S, et al. Adherence to antihypertensive medication and incident cardiovascular events in young adults with hypertension. Hypertension (Dallas, Tex: 1979) 2021; 77:1341–1349.
933. De Venecia T, Lu M, Figueredo VM. Hypertension in young adults. Postgrad Med 2016; 128:201–207.
934. Vasan RS. High blood pressure in young adulthood and risk of premature cardiovascular disease: calibrating treatment benefits to potential harm. JAMA 2018; 320:1760–1763.
935. Wang EA, Pletcher M, Lin F, Vittinghoff E, Kertesz SG, Kiefe CI, et al. Incarceration, incident hypertension, and access to health care: findings from the coronary artery risk development in young adults (CARDIA) study. Arch Intern Med 2009; 169:687–693.
936. Yano Y, Reis JP, Colangelo LA, Shimbo D, Viera AJ, Allen NB, et al. Association of blood pressure classification in young adults using the 2017 American College of Cardiology/American Heart Association Blood Pressure Guideline With Cardiovascular Events Later in Life. JAMA 2018; 320:1774–1782.
937. Safar ME, Levy BI, Struijker-Boudier H. Current perspectives on arterial stiffness and pulse pressure in hypertension and cardiovascular diseases. Circulation 2003; 107:2864–2869.
938. Abellan van Kan G, Rolland Y, Bergman H, Morley JE, Kritchevsky SB, Vellas B. The I.A.N A Task Force on frailty assessment of older people in clinical practice. J Nutr Health Aging 2008; 12:29–37.
939. European Commission: https://ec.europa.eu/eurostat/web/main. Accessed October 10.
940. Insua JT, Sacks HS, Lau TS, Lau J, Reitman D, Pagano D, et al. Drug treatment of hypertension in the elderly: a meta-analysis. Ann Intern Med 1994; 121:355–362.
941. Turnbull F, Neal B, Ninomiya T, Algert C, Arima H, Barzi F, et al. Effects of different regimens to lower blood pressure on major cardiovascular events in older and younger adults: meta-analysis of randomised trials. BMJ 2008; 336:1121–1123.
942. Kjeldsen SE, Lund-Johansen P, Nilsson PM, Mancia G. Unattended blood pressure measurements in the systolic blood pressure intervention trial: implications for entry and achieved blood pressure values compared with other trials. Hypertension (Dallas, Tex: 1979) 2016; 67:808–812.
943. Benetos A, Persu A, Kreutz R. Hypertension in older patients: a STEP forward?[Formula: see text]. Blood Press 2022; 31:118–120.
944. Mangoni AA, Jackson SH. Age-related changes in pharmacokinetics and pharmacodynamics: basic principles and practical applications. Br J Clin Pharmacol 2004; 57:6–14.
945. Benetos A, Rossignol P, Cherubini A, Joly L, Grodzicki T, Rajkumar C, et al. Polypharmacy in the aging patient: management of hypertension in octogenarians. JAMA 2015; 314:170–180.
946. Schneider J, Algharably EAE, Budnick A, Wenzel A, Dräger D, Kreutz R. High prevalence of multimorbidity and polypharmacy in elderly patients with chronic pain receiving home care are associated with multiple medication-related problems. Front Pharmacol 2021; 12:686990.
947. O’Mahony D, O'Sullivan D, Byrne S, O’Connor MN, Ryan C, Gallagher P. STOPP/START criteria for potentially inappropriate prescribing in older people: version 2. Age Ageing 2015; 44:213–218.
948. Warwick J, Falaschetti E, Rockwood K, Mitnitski A, Thijs L, Beckett N, et al. No evidence that frailty modifies the positive impact of antihypertensive treatment in very elderly people: an investigation of the impact of frailty upon treatment effect in the HYpertension in the Very Elderly Trial (HYVET) study, a double-blind, placebo-controlled study of antihypertensives in people with hypertension aged 80 and over. BMC Med 2015; 13:78.
949. Hanlon P, Corcoran N, Rughani G, Shah ASV, Mair FS, Guthrie B, et al. Observed and expected serious adverse event rates in randomised clinical trials for hypertension: an observational study comparing trials that do and do not focus on older people. Lancet Healthy Longev 2021; 2:e398–e406.
950. Sheppard JP, Lown M, Burt J, Temple E, Lowe R, Ashby H, et al. Generalizability of Blood Pressure Lowering Trials to Older Patients: Cross-Sectional Analysis. J Am Geriatr Soc 2020; 68:2508–2515.
951. Rockwood K, Song X, MacKnight C, Bergman H, Hogan DB, McDowell I, et al. A global clinical measure of fitness and frailty in elderly people. Cmaj 2005; 173:489–495.
952. Katz S. Assessing self-maintenance: activities of daily living, mobility, and instrumental activities of daily living. J Am Geriatr Soc 1983; 31:721–727.
953. Benetos A, Bulpitt CJ, Petrovic M, Ungar A, Agabiti Rosei E, Cherubini A, et al. An Expert Opinion From the European Society of Hypertension-European Union Geriatric Medicine Society Working Group on the Management of Hypertension in Very Old, Frail Subjects. Hypertension (Dallas, Tex: 1979) 2016; 67:820–825.
954. Applegate WB, Williamson JD, Berlowitz D. Deprescribing antihypertensive medication in elderly adults. JAMA 2020; 324:1682.
955. Folstein MF, Folstein SE, McHugh PR. “Mini-mental state”. A practical method for grading the cognitive state of patients for the clinician. J Psychiatr Res 1975; 12:189–198.
956. Guralnik JM, Simonsick EM, Ferrucci L, Glynn RJ, Berkman LF, Blazer DG, et al. A short physical performance battery assessing lower extremity function: association with self-reported disability and prediction of mortality and nursing home admission. J Gerontol 1994; 49:M85–94.
957. D’Ath P, Katona P, Mullan E, Evans S, Katona C. Screening, detection and management of depression in elderly primary care attenders. I: The acceptability and performance of the 15 item Geriatric Depression Scale (GDS15) and the development of short versions. Fam Pract 1994; 11:260–266.
958. Guigoz Y, Lauque S, Vellas BJ. Identifying the elderly at risk for malnutrition. the mini nutritional assessment. Clin Geriatr Med 2002; 18:737–757.
959. O’Keeffe LM, Simpkin AJ, Tilling K, Anderson EL, Hughes AD, Lawlor DA, et al. Sex-specific trajectories of measures of cardiovascular health during childhood and adolescence: A prospective cohort study. Atherosclerosis 2018; 278:190–196.
960. Gerdts E, Sudano I, Brouwers S, Borghi C, Bruno RM, Ceconi C, et al. Sex differences in arterial hypertension. Eur Heart J 2022; 43:4777–4788.
961. Reckelhoff JF. Gender differences in the regulation of blood pressure. Hypertension (Dallas, Tex: 1979) 2001; 37:1199–1208.
962. O’Kelly AC, Michos ED, Shufelt CL, Vermunt JV, Minissian MB, Quesada O, et al. Pregnancy and reproductive risk factors for cardiovascular disease in women. Circ Res 2022; 130:652–672.
963. Boggia J, Thijs L, Hansen TW, Li Y, Kikuya M, Björklund-Bodegård K, et al. Ambulatory blood pressure monitoring in 9357 subjects from 11 populations highlights missed opportunities for cardiovascular prevention in women. Hypertension (Dallas, Tex: 1979) 2011; 57:397–405.
964. Ji H, Niiranen TJ, Rader F, Henglin M, Kim A, Ebinger JE, et al. Sex differences in blood pressure associations with cardiovascular outcomes. Circulation 2021; 143:761–763.
965. Brown RM, Tamazi S, Weinberg CR, Dwivedi A, Mieres JH. Racial disparities in cardiovascular risk and cardiovascular care in women. Curr Cardiol Rep 2022; 24:1197–1208.
966. Gorostidi M, Vinyoles E, Banegas JR, de la Sierra A. Prevalence of white-coat and masked hypertension in national and international registries. Hypertens Res 2015; 38:1–7.
967. Franklin SS, Thijs L, Hansen TW, O’Brien E, Staessen JA. White-coat hypertension: new insights from recent studies. Hypertension (Dallas, Tex: 1979) 2013; 62:982–987.
968. Banegas JR, Segura J, de la Sierra A, Gorostidi M, Rodríguez-Artalejo F, Sobrino J, et al. Gender differences in office and ambulatory control of hypertension. Am J Med 2008; 121:1078–1084.
969. Conen D, Aeschbacher S, Thijs L, Li Y, Boggia J, Asayama K, et al. Age-specific differences between conventional and ambulatory daytime blood pressure values. Hypertension (Dallas, Tex: 1979) 2014; 64:1073–1079.
970. Muiesan ML, Paini A, Aggiusti C, Bertacchini F, Rosei CA, Salvetti M. Hypertension and Organ Damage in Women. High Blood Press Cardiovasc Prev 2018; 25:245–252.
971. Okin PM, Gerdts E, Kjeldsen SE, Julius S, Edelman JM, Dahlöf B, et al. Gender differences in regression of electrocardiographic left ventricular hypertrophy during antihypertensive therapy. Hypertension (Dallas, Tex: 1979) 2008; 52:100–106.
972. Gerdts E, Regitz-Zagrosek V. Sex differences in cardiometabolic disorders. Nat Med 2019; 25:1657–1666.
973. Izzo R, Losi MA, Stabile E, Lönnebakken MT, Canciello G, Esposito G, et al. Development of Left Ventricular Hypertrophy in Treated Hypertensive Outpatients: The Campania Salute Network. Hypertension (Dallas, Tex: 1979) 2017; 69:136–142.
974. Gerdts E, Oikarinen L, Palmieri V, Otterstad JE, Wachtell K, Boman K, et al. Correlates of left atrial size in hypertensive patients with left ventricular hypertrophy: the Losartan Intervention For Endpoint Reduction in Hypertension (LIFE) Study. Hypertension (Dallas, Tex: 1979) 2002; 39:739–743.
975. Losi MA, Mancusi C, Midtbø H, Saeed S, de Simone G, Gerdts E. Impact of estimated left atrial volume on prognosis in patients with asymptomatic mild to moderate aortic valve stenosis. Int J Cardiol 2019; 297:121–125.
976. Coutinho T. Arterial stiffness and its clinical implications in women. Can J Cardiol 2014; 30:756–764.
977. Hametner B, Wassertheurer S, Mayer CC, Danninger K, Binder RK, Weber T. Aortic pulse wave velocity predicts cardiovascular events and mortality in patients undergoing coronary angiography: a comparison of invasive measurements and noninvasive estimates. Hypertension (Dallas, Tex: 1979) 2021; 77:571–581.
978. Sullivan JC, Gillis EE. Sex and gender differences in hypertensive kidney injury. Am J Physiol Renal Physiol 2017; 313:F1009–f1017.
979. Wang Q, Xie D, Xu X, Qin X, Tang G, Wang B, et al. Blood pressure and renal function decline: a 7-year prospective cohort study in middle-aged rural Chinese men and women. J Hypertens 2015; 33:136–143.
980. Neugarten J, Acharya A, Silbiger SR. Effect of gender on the progression of nondiabetic renal disease: a meta-analysis. J Am Soc Nephrol 2000; 11:319–329.
981. Albrektsen G, Heuch I, Løchen ML, Thelle DS, Wilsgaard T, Njølstad I, et al. Risk of incident myocardial infarction by gender: Interactions with serum lipids, blood pressure and smoking. The Tromsø Study 1979–2012. Atherosclerosis 2017; 261:52–59.
982. Millett ERC, Peters SAE, Woodward M. Sex differences in risk factors for myocardial infarction: cohort study of UK Biobank participants. BMJ 2018; 363:k4247.
983. Kringeland E, Tell GS, Midtbø H, Igland J, Haugsgjerd TR, Gerdts E. Stage 1 hypertension, sex, and acute coronary syndromes during midlife: the Hordaland Health Study. Eur J Prev Cardiol 2022; 29:147–154.
984. Kubicki DM, Xu M, Akwo EA, Dixon D, Muñoz D, Blot WJ, et al. Race and sex differences in modifiable risk factors and incident heart failure. JACC Heart Fail 2020; 8:122–130.
985. Levy D, Larson MG, Vasan RS, Kannel WB, Ho KK. The progression from hypertension to congestive heart failure. JAMA 1996; 275:1557–1562.
986. Stolfo D, Uijl A, Vedin O, Strömberg A, Faxén UL, Rosano GMC, et al. Sex-based differences in heart failure across the ejection fraction spectrum: phenotyping, and prognostic and therapeutic implications. JACC Heart Fail 2019; 7:505–515.
987. Gerber Y, Weston SA, Redfield MM, Chamberlain AM, Manemann SM, Jiang R, et al. A contemporary appraisal of the heart failure epidemic in Olmsted County, Minnesota, 2000 to 2010. JAMA Intern Med 2015; 175:996–1004.
988. Conrad N, Judge A, Tran J, Mohseni H, Hedgecott D, Crespillo AP, et al. Temporal trends and patterns in heart failure incidence: a population-based study of 4 million individuals. Lancet 2018; 391:572–580.
989. Sharashova E, Wilsgaard T, Ball J, Morseth B, Gerdts E, Hopstock LA, et al. Long-term blood pressure trajectories and incident atrial fibrillation in women and men: the Tromsø Study. Eur Heart J 2020; 41:1554–1562.
990. Kavousi M. Differences in epidemiology and risk factors for atrial fibrillation between women and men. Front Cardiovasc Med 2020; 7:3.
991. Westerman S, Wenger N. Gender differences in atrial fibrillation: a review of epidemiology, management, and outcomes. Curr Cardiol Rev 2019; 15:136–144.
992. Gerdts E, Okin PM, de Simone G, Cramariuc D, Wachtell K, Boman K, et al. Gender differences in left ventricular structure and function during antihypertensive treatment: the Losartan Intervention for Endpoint Reduction in Hypertension Study. Hypertension (Dallas, Tex: 1979) 2008; 51:1109–1114.
993. Anstey KJ, Peters R, Mortby ME, Kiely KM, Eramudugolla R, Cherbuin N, et al. Association of sex differences in dementia risk factors with sex differences in memory decline in a population-based cohort spanning 20–76 years. Sci Rep 2021; 11:7710.
994. Gilsanz P, Mayeda ER, Glymour MM, Quesenberry CP, Mungas DM, DeCarli C, et al. Female sex, early-onset hypertension, and risk of dementia. Neurology 2017; 89:1886–1893.
995. Kjeldsen SE, Hedner T, Syvertsen JO, Lund-Johansen P, Hansson L, Lanke J, et al. Influence of age, sex and blood pressure on the principal endpoints of the Nordic Diltiazem (NORDIL) Study. J Hypertens 2002; 20:1231–1237.
996. Os I, Franco V, Kjeldsen SE, Manhem K, Devereux RB, Gerdts E, et al. Effects of losartan in women with hypertension and left ventricular hypertrophy: results from the Losartan Intervention for Endpoint Reduction in Hypertension Study. Hypertension (Dallas, Tex: 1979) 2008; 51:1103–1108.
997. Kjeldsen SE, Warnold I, Hansson L. Influence of gender on prevention of myocardial infarction by antihypertensives and acetylsalicylic acid: the HOT study. J Gend Specif Med 2000; 3:35–38.
998. Wing LM, Reid CM, Ryan P, Beilin LJ, Brown MA, Jennings GL, et al. A comparison of outcomes with angiotensin-converting--enzyme inhibitors and diuretics for hypertension in the elderly. N Engl J Med 2003; 348:583–592.
999. Foy CG, Lovato LC, Vitolins MZ, Bates JT, Campbell R, Cushman WC, et al. Gender, blood pressure, and cardiovascular and renal outcomes in adults with hypertension from the Systolic Blood Pressure Intervention Trial. J Hypertens 2018; 36:904–915.
1000. Ochoa-Jimenez R, Viquez-Beita K, Daluwatte C, Zusterzeel R. Sex differences of patients with systemic hypertension (from the analysis of the systolic blood pressure intervention trial [SPRINT]). Am J Cardiol 2018; 122:985–993.
1001. Sacks FM, Svetkey LP, Vollmer WM, Appel LJ, Bray GA, Harsha D, et al. Effects on blood pressure of reduced dietary sodium and the Dietary Approaches to Stop Hypertension (DASH) diet. DASH-Sodium Collaborative Research Group. N Engl J Med 2001; 344:3–10.
1002. Pana TA, Luben RN, Mamas MA, Potter JF, Wareham NJ, Khaw KT, et al. Long Term Prognostic Impact of Sex-specific Longitudinal Changes in Blood Pressure. The EPIC-Norfolk Prospective Population Cohort Study. Eur J Prev Cardiol 2022; 29:180–191.
1003. Bager JE, Manhem K, Andersson T, Hjerpe P, Bengtsson-Bostrom K, Ljungman C, et al. Hypertension: sex-related differences in drug treatment, prevalence and blood pressure control in primary care. J Hum Hypertens 2023; 37:662–670.
1004. Polaczyk M, Olszanecka A, Wojciechowska W, Rajzer M, Stolarz-Skrzypek K. The occurrence of drug-induced side effects in women and men with arterial hypertension and comorbidities. Kardiol Pol 2022; 80:1094–1103.
1005. Patel H, Bell D, Molokhia M, Srishanmuganathan J, Patel M, Car J, et al. Trends in hospital admissions for adverse drug reactions in England: analysis of national hospital episode statistics 1998–2005. BMC Clin Pharmacol 2007; 7:9.
1006. Kloner RA, Sowers JR, DiBona GF, Gaffney M, Wein M. Sex- and age-related antihypertensive effects of amlodipine. The Amlodipine Cardiovascular Community Trial Study Group. Am J Cardiol 1996; 77:713–722.
1007. Igho Pemu P, Ofili E. Hypertension in women: part I. J Clin Hypertens (Greenwich) 2008; 10:406–410.
1008. Biffi A, Rea F, Iannaccone T, Filippelli A, Mancia G, Corrao G. Sex differences in the adherence of antihypertensive drugs: a systematic review with meta-analyses. BMJ Open 2020; 10:e036418.
1009. Thomopoulos C, Salamalekis G, Kintis K, Andrianopoulou I, Michalopoulou H, Skalis G, et al. Risk of hypertensive disorders in pregnancy following assisted reproductive technology: overview and meta-analysis. J Clin Hypertens (Greenwich) 2017; 19:173–183.
1010. Chih HJ, Elias FTS, Gaudet L, Velez MP. Assisted reproductive technology and hypertensive disorders of pregnancy: systematic review and meta-analyses. BMC Pregnancy Childbirth 2021; 21:449.
1011. Gillum LA, Mamidipudi SK, Johnston SC. Ischemic stroke risk with oral contraceptives: A meta-analysis. JAMA 2000; 284:72–78.
1012. Lidegaard, Løkkegaard E, Jensen A, Skovlund CW, Keiding N. Thrombotic stroke and myocardial infarction with hormonal contraception. N Engl J Med 2012; 366:2257–2266.
1013. Dinger J, Do Minh T, Heinemann K. Impact of estrogen type on cardiovascular safety of combined oral contraceptives. Contraception 2016; 94:328–339.
1014. Gerstman BB, Piper JM, Tomita DK, Ferguson WJ, Stadel BV, Lundin FE. Oral contraceptive estrogen dose and the risk of deep venous thromboembolic disease. Am J Epidemiol 1991; 133:32–37.
1015. Glisic M, Shahzad S, Tsoli S, Chadni M, Asllanaj E, Rojas LZ, et al. Association between progestin-only contraceptive use and cardiometabolic outcomes: A systematic review and meta-analysis. Eur J Prev Cardiol 2018; 25:1042–1052.
1016. Lidegaard Ø, Nielsen LH, Skovlund CW, Skjeldestad FE, Løkkegaard E. Risk of venous thromboembolism from use of oral contraceptives containing different progestogens and oestrogen doses: Danish cohort study, 2001–2009. BMJ 2011; 343:d6423.
1017. Mantha S, Karp R, Raghavan V, Terrin N, Bauer KA, Zwicker JI. Assessing the risk of venous thromboembolic events in women taking progestin-only contraception: a meta-analysis. BMJ 2012; 345:e4944.
1018. Cardoso F, Polónia J, Santos A, Silva-Carvalho J, Ferreira-de-Almeida J. Low-dose oral contraceptives and 24-hour ambulatory blood pressure. Int J Gynaecol Obstet 1997; 59:237–243.
1019. Chasan-Taber L, Willett WC, Manson JE, Spiegelman D, Hunter DJ, Curhan G, et al. Prospective study of oral contraceptives and hypertension among women in the United States. Circulation 1996; 94:483–489.
1020. Tepper NK, Curtis KM, Steenland MW, Marchbanks PA. Blood pressure measurement prior to initiating hormonal contraception: a systematic review. Contraception 2013; 87:631–638.
1021. Curtis KM, Jatlaoui TC, Tepper NK, Zapata LB, Horton LG, Jamieson DJ, et al. US Selected Practice Recommendations for Contraceptive Use, 2016. MMWR Recomm Rep 2016; 65:1–66.
1022. Lubianca JN, Moreira LB, Gus M, Fuchs FD. Stopping oral contraceptives: an effective blood pressure-lowering intervention in women with hypertension. J Hum Hypertens 2005; 19:451–455.
1023. ACOG Practice Bulletin No. 206. Use of Hormonal Contraception in Women With Coexisting Medical Conditions. Obstet Gynecol 2019; 133:e128–e150.
1024. Mosca L, Benjamin EJ, Berra K, Bezanson JL, Dolor RJ, Lloyd-Jones DM, et al. Effectiveness-based guidelines for the prevention of cardiovascular disease in women--2011 update: a guideline from the american heart association. Circulation 2011; 123:1243–1262.
1025. Issa Z, Seely EW, Rahme M, El-Hajj Fuleihan G. Effects of hormone therapy on blood pressure. Menopause 2015; 22:456–468.
1026. Shannon G, Jansen M, Williams K, Cáceres C, Motta A, Odhiambo A, et al. Gender equality in science, medicine, and global health: where are we at and why does it matter? Lancet 2019; 393:560–569.
1027. Connelly PJ, Clark A, Touyz RM, Delles C. Transgender adults, gender-affirming hormone therapy and blood pressure: a systematic review. J Hypertens 2021; 39:223–230.
1028. Hembree WC, Cohen-Kettenis PT, Gooren L, Hannema SE, Meyer WJ, Murad MH, et al. Endocrine Treatment of Gender-Dysphoric/Gender-Incongruent Persons: An Endocrine Society Clinical Practice Guideline. J Clin Endocrinol Metab 2017; 102:3869–3903.
1029. Whelton PK, Einhorn PT, Muntner P, Appel LJ, Cushman WC, Diez Roux AV, et al. Research needs to improve hypertension treatment and control in African Americans. Hypertension (Dallas, Tex: 1979) 2016; 68:1066–1072.
1030. Kaufman JS, Dolman L, Rushani D, Cooper RS. The contribution of genomic research to explaining racial disparities in cardiovascular disease: a systematic review. Am J Epidemiol 2015; 181:464–472.
1031. Mabhida SE, Mashatola L, Kaur M, Sharma JR, Apalata T, Muhamed B, et al. Hypertension in African populations: review and computational insights. Genes (Basel) 2021; 12:
1032. Diemer FS, Baldew SM, Haan YC, Karamat FA, Oehlers GP, van Montfrans GA, et al. Aortic pulse wave velocity in individuals of Asian and African ancestry: the HELISUR study. J Hum Hypertens 2020; 34:108–116.
1033. Park CM, Tillin T, March K, Jones S, Whincup PH, Mayet J, et al. Adverse effect of diabetes and hyperglycaemia on arterial stiffness in Europeans, South Asians, and African Caribbeans in the SABRE study. J Hypertens 2016; 34:282–289.
1034. Rezai MR, Wallace AM, Sattar N, Finn JD, Wu FC, Cruickshank JK. Ethnic differences in aortic pulse wave velocity occur in the descending aorta and may be related to vitamin D. Hypertension (Dallas, Tex: 1979) 2011; 58:247–253.
1035. Cruickshank JK, Silva MJ, Molaodi OR, Enayat ZE, Cassidy A, Karamanos A, et al. Ethnic differences in and childhood influences on early adult pulse wave velocity: the determinants of adolescent, now young adult, social wellbeing, and health longitudinal study. Hypertension (Dallas, Tex: 1979) 2016; 67:1133–1141.
1036. Ahmed KS, Bogdanet D, Abadi S, Dineen R, Boran G, Woods CP, et al. Rates of abnormal aldosterone/renin ratio in African-origin compared to European-origin patients: A retrospective study. Clin Endocrinol (Oxf) 2019; 90:528–533.
1037. Jones E, Rayner B. The importance of the epithelial sodium channel in determining salt sensitivity in people of African origin. Pediatr Nephrol 2021; 36:237–243.
1038. Wu X, Senanayake R, Goodchild E, Bashari WA, Salsbury J, Cabrera CP, et al. [(11)C]metomidate PET-CT versus adrenal vein sampling for diagnosing surgically curable primary aldosteronism: a prospective, within-patient trial. Nat Med 2023; 29:190–202.
1039. Ayoola OO, Omotade OO, Gemmell I, Clayton PE, Cruickshank JK. The impact of malaria in pregnancy on changes in blood pressure in children during their first year of life. Hypertension (Dallas, Tex: 1979) 2014; 63:167–172.
1040. Etyang AO, Kapesa S, Odipo E, Bauni E, Kyobutungi C, Abdalla M, et al. Effect of previous exposure to malaria on blood pressure in Kilifi, Kenya: a Mendelian randomization study. J Am Heart Assoc 2019; 8:e011771.
1041. Roberts ML, Kotchen TA, Pan X, Li Y, Yang C, Liu P, et al. Unique associations of DNA Methylation regions with 24-hour blood pressure phenotypes in black participants. Hypertension (Dallas, Tex: 1979) 2022; 79:761–772.
1042. Tennant IA, Barnett AT, Thompson DS, Kips J, Boyne MS, Chung EE, et al. Impaired cardiovascular structure and function in adult survivors of severe acute malnutrition. Hypertension (Dallas, Tex: 1979) 2014; 64:664–671.
1043. Dolezsar CM, McGrath JJ, Herzig AJM, Miller SB. Perceived racial discrimination and hypertension: a comprehensive systematic review. Health Psychol 2014; 33:20–34.
1044. Forde AT, Sims M, Muntner P, Lewis T, Onwuka A, Moore K, et al. Discrimination and Hypertension Risk Among African Americans in the Jackson Heart Study. Hypertension (Dallas, Tex: 1979) 2020; 76:715–723.
1045. Kershaw KN, Robinson WR, Gordon-Larsen P, Hicken MT, Goff DC Jr, Carnethon MR, et al. Association of Changes in Neighborhood-Level Racial Residential Segregation With Changes in Blood Pressure Among Black Adults: The CARDIA Study. JAMA Intern Med 2017; 177:996–1002.
1046. Agyemang C, van Oeffelen AA, Norredam M, Kappelle LJ, Klijn CJ, Bots ML, et al. Socioeconomic inequalities in stroke incidence among migrant groups: analysis of nationwide data. Stroke 2014; 45:2397–2403.
1047. Victora CG, Hartwig FP, Vidaletti LP, Martorell R, Osmond C, Richter LM, et al. Effects of early-life poverty on health and human capital in children and adolescents: analyses of national surveys and birth cohort studies in LMICs. Lancet 2022; 399:1741–1752.
1048. Brewster LM, Tong J, Yan LL, Moe J, Harris VC, van Montfrans GA. Health Professionals’ Perceptions of Disparities in Hypertension Control: A Mixed Methods Study. Am J Hypertens 2022; 35:955–963.
1049. Modesti PA, Reboldi G, Cappuccio FP, Agyemang C, Remuzzi G, Rapi S, et al. Panethnic Differences in Blood Pressure in Europe: A Systematic Review and Meta-Analysis. PLoS One 2016; 11:e0147601.
1050. Tillin T, Hughes AD, Mayet J, Whincup P, Sattar N, Forouhi NG, et al. The relationship between metabolic risk factors and incident cardiovascular disease in Europeans, South Asians, and African Caribbeans: SABRE (Southall and Brent Revisited) -- a prospective population-based study. J Am Coll Cardiol 2013; 61:1777–1786.
1051. Rahimi K, Emdin CA, MacMahon S. The epidemiology of blood pressure and its worldwide management. Circ Res 2015; 116:925–936.
1052. Brewster LM, Boermeester AS, Seedat YK, Van Montfrans GA. Initial combination therapy for hypertension in patients of African ancestry: a systematic review and meta-analysis. J Hypertens 2022; 40:629–640.
1053. Brewster LM, van Montfrans GA, Oehlers GP, Seedat YK. Systematic review: antihypertensive drug therapy in patients of African and South Asian ethnicity. Intern Emerg Med 2016; 11:355–374.
1054. Wright JT Jr, Dunn JK, Cutler JA, Davis BR, Cushman WC, Ford CE, et al. Outcomes in hypertensive black and nonblack patients treated with chlorthalidone, amlodipine, and lisinopril. JAMA 2005; 293:1595–1608.
1055. Ojji DB, Mayosi B, Francis V, Badri M, Cornelius V, Smythe W, et al. Comparison of Dual Therapies for Lowering Blood Pressure in Black Africans. N Engl J Med 2019; 380:2429-2439
1056. Thomopoulos C, Brguljan J, Cífková R, Persu A, Kreutz R. Mild chronic hypertension in pregnancy: to treat or wait?[Formula: see text]. Blood Press 2022; 31:121–124.
1057. Regitz-Zagrosek V, Roos-Hesselink JW, Bauersachs J, Blomstrom-Lundqvist C, Cifkova R, De Bonis M, et al. 2018 ESC Guidelines for the management of cardiovascular diseases during pregnancy. Eur Heart J 2018; 39:3165–3241.
1058. Brown MA, Magee LA, Kenny LC, Karumanchi SA, McCarthy FP, Saito S, et al. The hypertensive disorders of pregnancy: ISSHP classification, diagnosis & management recommendations for international practice. Pregnancy Hypertens 2018; 13:291–310.
1059. Lee-Ann Hawkins T, Brown MA, Mangos GJ, Davis GK. Transient gestational hypertension: Not always a benign event. Pregnancy Hypertens 2012; 2:22–27.
1060. Seely EW, Ecker J. Chronic hypertension in pregnancy. Circulation 2014; 129:1254–1261.
1061. Stergiou GS, O’Brien E, Myers M, Palatini P, Parati G, Board SBSA. STRIDE BP: an international initiative for accurate blood pressure measurement. J Hypertens 2020; 38:395–399.
1062. Penny JA, Halligan AW, Shennan AH, Lambert PC, Jones DR, de Swiet M, et al. Automated, ambulatory, or conventional blood pressure measurement in pregnancy: which is the better predictor of severe hypertension? Am J Obstet Gynecol 1998; 178:521–526.
1063. Tucker KL, Mort S, Yu LM, Campbell H, Rivero-Arias O, Wilson HM, et al. Effect of Self-monitoring of Blood Pressure on Diagnosis of Hypertension During Higher-Risk Pregnancy: The BUMP 1 Randomized Clinical Trial. JAMA 2022; 327:1656–1665.
1064. Chappell LC, Tucker KL, Galal U, Yu LM, Campbell H, Rivero-Arias O, et al. Effect of Self-monitoring of Blood Pressure on Blood Pressure Control in Pregnant Individuals With Chronic or Gestational Hypertension: The BUMP 2 Randomized Clinical Trial. JAMA 2022; 327:1666–1678.
1065. Schmella MJ, Clifton RG, Althouse AD, Roberts JM. Uric Acid Determination in Gestational Hypertension: Is it as Effective a Delineator of Risk as Proteinuria in High-Risk Women? Reprod Sci 2015; 22:1212–1219.
1066. Chappell LC, Shennan AH. Assessment of proteinuria in pregnancy. BMJ 2008; 336:968–969.
1067. Waugh J, Hooper R, Lamb E, Robson S, Shennan A, Milne F, et al. Spot protein-creatinine ratio and spot albumin-creatinine ratio in the assessment of pre-eclampsia: a diagnostic accuracy study with decision-analytic model-based economic evaluation and acceptability analysis. Health Technol Assess 2017; 21:1–90.
1068. Bartsch E, Medcalf KE, Park AL, Ray JG. High Risk of Pre-eclampsia Identification G. Clinical risk factors for pre-eclampsia determined in early pregnancy: systematic review and meta-analysis of large cohort studies. BMJ 2016; 353:i1753.
1069. Rolnik DL, Wright D, Poon LC, O’Gorman N, Syngelaki A, de Paco Matallana C, et al. Aspirin versus placebo in pregnancies at high risk for preterm preeclampsia. N Engl J Med 2017; 377:613–622.
1070. Webster K, Fishburn S, Maresh M, Findlay SC, Chappell LC, Guideline C. Diagnosis and management of hypertension in pregnancy: summary of updated NICE guidance. BMJ 2019; 366:l5119.
1071. Garovic VD, Dechend R, Easterling T, Karumanchi SA, McMurtry Baird S, Magee LA, et al. Hypertension in pregnancy: diagnosis, blood pressure goals, and pharmacotherapy: a scientific statement from the American Heart Association. Hypertension (Dallas, Tex: 1979) 2022; 79:e21–e41.
1072. O’Gorman N, Wright D, Poon LC, Rolnik DL, Syngelaki A, de Alvarado M, et al. Multicenter screening for pre-eclampsia by maternal factors and biomarkers at 11-13 weeks’ gestation: comparison with NICE guidelines and ACOG recommendations. Ultrasound Obstet Gynecol 2017; 49:756–760.
1073. Chappell LC, Duckworth S, Seed PT, Griffin M, Myers J, Mackillop L, et al. Diagnostic accuracy of placental growth factor in women with suspected preeclampsia: a prospective multicenter study. Circulation 2013; 128:2121–2131.
1074. Zeisler H, Llurba E, Chantraine F, Vatish M, Staff AC, Sennstrom M, et al. Predictive Value of the sFlt-1:PlGF Ratio in Women with Suspected Preeclampsia. N Engl J Med 2016; 374:13–22.
1075. Barakat R, Pelaez M, Cordero Y, Perales M, Lopez C, Coteron J, et al. Exercise during pregnancy protects against hypertension and macrosomia: randomized clinical trial. Am J Obstet Gynecol 2016; 214:649e641-648.
1076. Di Mascio D, Magro-Malosso ER, Saccone G, Marhefka GD, Berghella V. Exercise during pregnancy in normal-weight women and risk of preterm birth: a systematic review and meta-analysis of randomized controlled trials. Am J Obstet Gynecol 2016; 215:561–571.
1077. Santos S, Voerman E, Amiano P, Barros H, Beilin LJ, Bergstrom A, et al. Impact of maternal body mass index and gestational weight gain on pregnancy complications: an individual participant data meta-analysis of European, North American and Australian cohorts. BJOG 2019; 126:984–995.
1078. Hofmeyr GJ, Lawrie TA, Atallah AN, Torloni MR. Calcium supplementation during pregnancy for preventing hypertensive disorders and related problems. Cochrane Database Syst Rev 2018; 10:CD001059.
1079. Hoeltzenbein M, Beck E, Fietz AK, Wernicke J, Zinke S, Kayser A, et al. Pregnancy outcome after first trimester use of methyldopa: a prospective cohort study. Hypertension (Dallas, Tex: 1979) 2017; 70:201–208.
1080. Magee LA, Singer J, von Dadelszen P, Group CS. Less-tight versus tight control of hypertension in pregnancy. N Engl J Med 2015; 372:2367–2368.
1081. Tita AT, Szychowski JM, Boggess K, Dugoff L, Sibai B, Lawrence K, et al. Treatment for Mild Chronic Hypertension during Pregnancy. N Engl J Med 2022; 386:1781–1792.
1082. Whelan A, Izewski J, Berkelhammer C, Walloch J, Kay HH. Labetalol-Induced Hepatotoxicity during Pregnancy: A Case Report. AJP Rep 2020; 10:e210–e212.
1083. Vigil-DeGracia P, Ludmir J, Ng J, Reyes-Tejada O, Nova C, Beltré A, et al. Is there benefit to continue magnesium sulphate postpartum in women receiving magnesium sulphate before delivery? A randomised controlled study. Bjog 2018; 125:1304–1311.
1084. Deng NJ, Xian-Yu CY, Han RZ, Huang CY, Ma YT, Li HJ, et al. Pharmaceutical administration for severe hypertension during pregnancy: Network meta-analysis. Front Pharmacol 2022; 13:1092501.
1085. Cifkova R, Johnson MR, Kahan T, Brguljan J, Williams B, Coca A, et al. Peripartum management of hypertension: a position paper of the ESC Council on Hypertension and the European Society of Hypertension. Eur Heart J Cardiovasc Pharmacother 2020; 6:384–393.
1086. Pappaccogli M, Prejbisz A, Ciurică S, Bruno RM, Aniszczuk-Hybiak A, Bracalente I, et al. Pregnancy-Related Complications in Patients With Fibromuscular Dysplasia: A Report From the European/International Fibromuscular Dysplasia Registry. Hypertension (Dallas, Tex: 1979) 2020; 76:545–553.
1087. Morton A. Primary aldosteronism and pregnancy. Pregnancy Hypertens 2015; 5:259–262.
1088. Landau E, Amar L. Primary aldosteronism and pregnancy. Ann Endocrinol (Paris) 2016; 77:148–160.
1089. Lenders JW. Pheochromocytoma and pregnancy: a deceptive connection. Eur J Endocrinol 2012; 166:143–150.
1090. Maynard SE, Thadhani R. Pregnancy and the kidney. J Am Soc Nephrol 2009; 20:14–22.
1091. Imbasciati E, Gregorini G, Cabiddu G, Gammaro L, Ambroso G, Del Giudice A, et al. Pregnancy in CKD stages 3 to 5: fetal and maternal outcomes. Am J Kidney Dis 2007; 49:753–762.
1092. Rutgers University Press, Thomopoulos C, Makris T. Iatrogenic Aspects of Hypertension in Pregnancy. 2018.
1093. Lopes Perdigao J, Lewey J, Hirshberg A, Koelper N, Srinivas SK, Elovitz MA, et al. Furosemide for Accelerated Recovery of Blood Pressure Postpartum in women with a hypertensive disorder of pregnancy: A Randomized Controlled Trial. Hypertension (Dallas, Tex: 1979) 2021; 77:1517–1524.
1094. Lindheimer MD, Taler SJ, Cunningham FG. American Society of H. ASH position paper: hypertension in pregnancy. J Clin Hypertens (Greenwich) 2009; 11:214–225.
1095. Geller DS, Farhi A, Pinkerton N, Fradley M, Moritz M, Spitzer A, et al. Activating mineralocorticoid receptor mutation in hypertension exacerbated by pregnancy. Science 2000; 289:119–123.
1096. Podymow T, August P. Update on the use of antihypertensive drugs in pregnancy. Hypertension (Dallas, Tex: 1979) 2008; 51:960–969.
1097. Lykke JA, Langhoff-Roos J, Sibai BM, Funai EF, Triche EW, Paidas MJ. Hypertensive pregnancy disorders and subsequent cardiovascular morbidity and type 2 diabetes mellitus in the mother. Hypertension (Dallas, Tex: 1979) 2009; 53:944–951.
1098. Mannisto T, Mendola P, Vaarasmaki M, Jarvelin MR, Hartikainen AL, Pouta A, et al. Elevated blood pressure in pregnancy and subsequent chronic disease risk. Circulation 2013; 127:681–690.
1099. Parikh NI, Gonzalez JM, Anderson CAM, Judd SE, Rexrode KM, Hlatky MA, et al. Adverse Pregnancy Outcomes and Cardiovascular Disease Risk: Unique Opportunities for Cardiovascular Disease Prevention in Women: A Scientific Statement From the American Heart Association. Circulation 2021; 143:e902–e916.
1100. Brouwers L, van der Meiden-van Roest AJ, Savelkoul C, Vogelvang TE, Lely AT, Franx A, et al. Recurrence of pre-eclampsia and the risk of future hypertension and cardiovascular disease: a systematic review and meta-analysis. BJOG 2018; 125:1642–1654.
1101. Rayes B, Ardissino M, Slob EAW, Patel KHK, Girling J, Ng FS. Association of Hypertensive Disorders of Pregnancy With Future Cardiovascular Disease. JAMA Network Open 2023; 6:e230034–e1230034.
1102. Saladini F, Mancusi C, Bertacchini F, Spannella F, Maloberti A, Giavarini A, et al. Diagnosis and treatment of hypertensive emergencies and urgencies among Italian emergency and intensive care departments. Results from an Italian survey: Progetto GEAR (Gestione dell’Emergenza e urgenza in ARea critica). Eur J Intern Med 2020; 71:50–56.
1103. van den Born BH, Lip GYH, Brguljan-Hitij J, Cremer A, Segura J, Morales E, et al. ESC Council on hypertension position document on the management of hypertensive emergencies. Eur Heart J Cardiovasc Pharmacother 2019; 5:37–46.
1104. Vaughan CJ, Delanty N. Hypertensive emergencies. Lancet 2000; 356:411–417.
1105. Rubin S, Cremer A, Boulestreau R, Rigothier C, Kuntz S, Gosse P. Malignant hypertension: diagnosis, treatment and prognosis with experience from the Bordeaux cohort. J Hypertens 2019; 37:316–324.
1106. Boulestreau R, van den Born BH, Lip GYH, Gupta A. Malignant Hypertension: Current Perspectives and Challenges. J Am Heart Assoc 2022; 11:e023397.
1107. van den Born BJ, Koopmans RP, Groeneveld JO, van Montfrans GA. Ethnic disparities in the incidence, presentation and complications of malignant hypertension. J Hypertens 2006; 24:2299–2304.
1108. van den Born BJ, Lowenberg EC, van der Hoeven NV, de Laat B, Meijers JC, Levi M, et al. Endothelial dysfunction, platelet activation, thrombogenesis and fibrinolysis in patients with hypertensive crisis. J Hypertens 2011; 29:922–927.
1109. Amraoui F, Van Der Hoeven NV, Van Valkengoed IG, Vogt L, Van Den Born BJ. Mortality and cardiovascular risk in patients with a history of malignant hypertension: a case-control study. J Clin Hypertens (Greenwich) 2014; 16:122–126.
1110. Shantsila A, Lip GYH. Malignant Hypertension Revisited-Does This Still Exist? Am J Hypertens 2017; 30:543–549.
1111. Astarita A, Covella M, Vallelonga F, Cesareo M, Totaro S, Ventre L, et al. Hypertensive emergencies and urgencies in emergency departments: a systematic review and meta-analysis. J Hypertens 2020; 38:1203–1210.
1112. Salvetti M, Paini A, Colonetti E, Tarozzi L, Bertacchini F, Aggiusti C, et al. Hypertensive emergencies and urgencies: a single-centre experience in Northern Italy 2008-2015. J Hypertens 2020; 38:52–58.
1113. Perez MI, Musini VM. Pharmacological interventions for hypertensive emergencies: a Cochrane systematic review. J Hum Hypertens 2008; 22:596–607.
1114. Grassi D, O’Flaherty M, Pellizzari M, Bendersky M, Rodriguez P, Turri D, et al. Hypertensive urgencies in the emergency department: evaluating blood pressure response to rest and to antihypertensive drugs with different profiles. J Clin Hypertens (Greenwich) 2008; 10:662–667.
1115. Backer HD, Decker L, Ackerson L. Reproducibility of increased blood pressure during an emergency department or urgent care visit. Ann Emerg Med 2003; 41:507–512.
1116. Lane DA, Lip GY, Beevers DG. Improving survival of malignant hypertension patients over 40 years. Am J Hypertens 2009; 22:1199–1204.
1117. Oras P, Habel H, Skoglund PH, Svensson P. Elevated blood pressure in the emergency department: a risk factor for incident cardiovascular disease. Hypertension (Dallas, Tex: 1979) 2020; 75:229–236.
1118. Shin JH, Kim BS, Lyu M, Kim HJ, Lee JH, Park JK, et al. Clinical Characteristics and Predictors of All-Cause Mortality in Patients with Hypertensive Urgency at an Emergency Department. J Clin Med 2021; 10:
1119. Skoglund PH, Svensson P. Asking the patient or measuring blood pressure in the emergency department: which one is best? Curr Hypertens Rep 2016; 18:53.
1120. Brody AM, Miller J, Polevoy R, Nakhle A, Levy PD. Institutional Pathways to Improve Care of Patients with Elevated Blood Pressure in the Emergency Department. Curr Hypertens Rep 2018; 20:30.
1121. Goldberg EM, Wilson T, Saucier C, Brody AM, Levy PD, Eaton CB, et al. Achieving the BpTRUth: emergency department hypertension screening and the Centers for Medicare & Medicaid Services quality measure. J Am Soc Hypertens 2017; 11:290–294.
1122. Koutsaki M, Patoulias D, Tsinivizov P, Doumas M, Kallistratos M, Thomopoulos C, et al. Evaluation, risk stratification and management of hypertensive patients in the perioperative period. Eur J Intern Med 2019; 69:1–7.
1123. Hartle A, McCormack T, Carlisle J, Anderson S, Pichel A, Beckett N, et al. The measurement of adult blood pressure and management of hypertension before elective surgery: Joint Guidelines from the Association of Anaesthetists of Great Britain and Ireland and the British Hypertension Society. Anaesthesia 2016; 71:326–337.
1124. Koutsaki M, Thomopoulos C, Achimastos A, Kallistratos M, Batistaki C, Chatziagelaki E, et al. Perioperative SBP changes during orthopedic surgery in the elderly: clinical implications. J Hypertens 2019; 37:1705–1713.
1125. Group PS, Devereaux PJ, Yang H, Yusuf S, Guyatt G, Leslie K, et al. Effects of extended-release metoprolol succinate in patients undergoing non-cardiac surgery (POISE trial): a randomised controlled trial. Lancet 2008; 371:1839–1847.
1126. Walden RJ, Tomlinson B, Graham B, Liu JB, Prichard BN. Withdrawal phenomena after atenolol and bopindolol: haemodynamic responses in healthy volunteers. Br J Clin Pharmacol 1990; 30:557–565.
1127. London MJ. Perioperative beta-blockade, discontinuation, and complications: do you really know it when you see it? Anesthesiology 2009; 111:690–694.
1128. Futier E, Lefrant JY, Guinot PG, Godet T, Lorne E, Cuvillon P, et al. Effect of individualized vs standard blood pressure management strategies on postoperative organ dysfunction among high-risk patients undergoing major surgery: a randomized clinical trial. JAMA 2017; 318:1346–1357.
1129. Kozarek K, Sanders RD, Head D. Perioperative blood pressure in the elderly. Curr Opin Anaesthesiol 2020; 33:122–120.
1130. Danaei G, Ding EL, Mozaffarian D, Taylor B, Rehm J, Murray CJ, et al. The preventable causes of death in the United States: comparative risk assessment of dietary, lifestyle, and metabolic risk factors. PLoS Med 2009; 6:e1000058.
1131. MacMahon S, Peto R, Cutler J, Collins R, Sorlie P, Neaton J, et al. Blood pressure, stroke, and coronary heart disease. Part 1, Prolonged differences in blood pressure: prospective observational studies corrected for the regression dilution bias. Lancet 1990; 335:765–774.
1132. Collins R, Peto R, MacMahon S, Hebert P, Fiebach NH, Eberlein KA, et al. Blood pressure, stroke, and coronary heart disease. Part 2, Short-term reductions in blood pressure: overview of randomised drug trials in their epidemiological context. Lancet 1990; 335:827–838.
1133. Vidal-Petiot E, Ford I, Greenlaw N, Ferrari R, Fox KM, Tardif JC, et al. Cardiovascular event rates and mortality according to achieved systolic and diastolic blood pressure in patients with stable coronary artery disease: an international cohort study. Lancet 2016; 388:2142–2152.
1134. Cooper-DeHoff RM, Handberg EM, Mancia G, Zhou Q, Champion A, Legler UF, et al. INVEST revisited: review of findings from the International Verapamil SR-Trandolapril Study. Expert Rev Cardiovasc Ther 2009; 7:1329–1340.
1135. Kjeldsen SE, Berge E, Bangalore S, Messerli FH, Mancia G, Holzhauer B, et al. No evidence for a J-shaped curve in treated hypertensive patients with increased cardiovascular risk: The VALUE trial. Blood Press 2016; 25:83–92.
1136. Bohm M, Schumacher H, Teo KK, Lonn E, Mahfoud F, Mann JFE, et al. Achieved diastolic blood pressure and pulse pressure at target systolic blood pressure (120-140 mmHg) and cardiovascular outcomes in high-risk patients: results from ONTARGET and TRANSCEND trials. Eur Heart J 2018; 39:3105–3114.
1137. Böhm M, Schumacher H, Teo KK, Lonn EM, Mahfoud F, Mann JFE, et al. Achieved blood pressure and cardiovascular outcomes in high-risk patients: results from ONTARGET and TRANSCEND trials. Lancet 2017; 389:2226–2237.
1138. Messerli FH, Bangalore S, Messerli AW, Räber L. The muddy waters of the J-curve and coronary revascularization. Eur Heart J 2020; 41:1684–1686.
1139. Messerli FH, Mancia G, Conti CR, Hewkin AC, Kupfer S, Champion A, et al. Dogma disputed: can aggressively lowering blood pressure in hypertensive patients with coronary artery disease be dangerous? Ann Intern Med 2006; 144:884–893.
1140. Sleight P, Redon J, Verdecchia P, Mancia G, Gao P, Fagard R, et al. Prognostic value of blood pressure in patients with high vascular risk in the Ongoing Telmisartan Alone and in combination with Ramipril Global Endpoint Trial study. J Hypertens 2009; 27:1360–1369.
1141. Bangalore S, Messerli FH, Wun CC, Zuckerman AL, DeMicco D, Kostis JB, et al. J-curve revisited: An analysis of blood pressure and cardiovascular events in the Treating to New Targets (TNT) Trial. Eur Heart J 2010; 31:2897–2908.
1142. Mancia G, Grassi G. Blood pressure targets in type 2 diabetes. Evidence against or in favour of an aggressive approach. Diabetologia 2018; 61:517–525.
1143. Hwang D, Lee JM, Kim HK, Choi KH, Rhee TM, Park J, et al. Prognostic Impact of β-Blocker Dose After Acute Myocardial Infarction. Circ J 2019; 83:410–417.
1144. Andersson C, Shilane D, Go AS, Chang TI, Kazi D, Solomon MD, et al. β-blocker therapy and cardiac events among patients with newly diagnosed coronary heart disease. J Am Coll Cardiol 2014; 64:247–252.
1145. Bangalore S, Makani H, Radford M, Thakur K, Toklu B, Katz SD, et al. Clinical outcomes with β-blockers for myocardial infarction: a meta-analysis of randomized trials. Am J Med 2014; 127:939–953.
1146. Manolis AJ, Boden WE, Collins P, Dechend R, Kallistratos MS, Lopez Sendon J, et al. State of the art approach to managing angina and ischemia: tailoring treatment to the evidence. Eur J Intern Med 2021; 92:40–47.
1147. Puymirat E, Riant E, Aissaoui N, Soria A, Ducrocq G, Coste P, et al. β blockers and mortality after myocardial infarction in patients without heart failure: multicentre prospective cohort study. Bmj 2016; 354:i4801.
1148. Hong J, Barry AR. Long-term beta-blocker therapy after myocardial infarction in the reperfusion era: a systematic review. Pharmacotherapy 2018; 38:546–554.
1149. Ho JE, Bittner V, Demicco DA, Breazna A, Deedwania PC, Waters DD. Usefulness of heart rate at rest as a predictor of mortality, hospitalization for heart failure, myocardial infarction, and stroke in patients with stable coronary heart disease (Data from the Treating to New Targets [TNT] trial). Am J Cardiol 2010; 105:905–911.
1150. Yusuf S, Sleight P, Pogue J, Bosch J, Davies R, et al. Heart Outcomes Prevention Evaluation Study I. Effects of an angiotensin-converting-enzyme inhibitor, ramipril, on cardiovascular events in high-risk patients. N Engl J Med 2000; 342:145–153.
1151. Fox KM. Investigators EUtOrocewPiscAd Efficacy of perindopril in reduction of cardiovascular events among patients with stable coronary artery disease: randomised, double-blind, placebo-controlled, multicentre trial (the EUROPA study). Lancet 2003; 362:782–788.
1152. Moser M, Hebert PR. Prevention of disease progression, left ventricular hypertrophy and congestive heart failure in hypertension treatment trials. J Am Coll Cardiol 1996; 27:1214–1218.
1153. Angeli F, Reboldi G, Poltronieri C, Stefanetti E, Bartolini C, Verdecchia P. The prognostic legacy of left ventricular hypertrophy: cumulative evidence after the MAVI study. J Hypertens 2015; 33:2322–2330.
1154. Major cardiovascular events in hypertensive patients randomized to doxazosin vs chlorthalidone: the antihypertensive and lipid-lowering treatment to prevent heart attack trial (ALLHAT). ALLHAT Collaborative Research Group. JAMA 2000; 283:1967–1975.
1155. Diuretic versus alpha-blocker as first-step antihypertensive therapy: final results from the Antihypertensive and Lipid-Lowering Treatment to Prevent Heart Attack Trial (ALLHAT). Hypertension (Dallas, Tex: 1979) 2003; 42:239–246.
1156. Heidenreich PA, Bozkurt B, Aguilar D, Allen LA, Byun JJ, Colvin MM, et al. 2022 AHA/ACC/HFSA Guideline for the Management of Heart Failure: A Report of the American College of Cardiology/American Heart Association Joint Committee on Clinical Practice Guidelines. Circulation 2022; 145:e895–e1032.
1157. McDonagh TA, Metra M, Adamo M, Gardner RS, Baumbach A, Böhm M, et al. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J 2021; 42:3599–3726.
1158. Kjeldsen SE, von Lueder TG, Smiseth OA, Wachtell K, Mistry N, Westheim AS, et al. Medical therapies for heart failure with preserved ejection fraction. Hypertension (Dallas, Tex: 1979) 2020; 75:23–32.
1159. Solomon SD, McMurray JJV, Claggett B, de Boer RA, DeMets D, Hernandez AF, et al. Dapagliflozin in heart failure with mildly reduced or preserved ejection fraction. N Engl J Med 2022; 387:1089–1098.
1160. Anker SD, Butler J, Filippatos G, Ferreira JP, Bocchi E, Böhm M, et al. Empagliflozin in heart failure with a preserved ejection fraction. N Engl J Med 2021; 385:1451–1461.
1161. Pitt B, Pfeffer MA, Assmann SF, Boineau R, Anand IS, Claggett B, et al. Spironolactone for heart failure with preserved ejection fraction. N Engl J Med 2014; 370:1383–1392.
1162. Hjornholm U, Larstorp ACK, Andersen MH, Hoieggen A. Directly observed therapy prior to ambulatory blood pressure measurement (DOT-HTN) in uncontrolled hypertensive patients - Effect on blood pressure, safety and patient perception. Blood Press 2019; 28:327–335.
1163. January CT, Wann LS, Calkins H, Chen LY, Cigarroa JE, Cleveland JC Jr, et al. 2019 AHA/ACC/HRS Focused Update of the 2014 AHA/ACC/HRS Guideline for the Management of Patients With Atrial Fibrillation: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines and the Heart Rhythm Society in Collaboration With the Society of Thoracic Surgeons. Circulation 2019; 140:e125–e151.
1164. Hindricks G, Potpara T, Dagres N, Arbelo E, Bax JJ, Blomstrom-Lundqvist C, et al. 2020 ESC Guidelines for the diagnosis and management of atrial fibrillation developed in collaboration with the European Association for Cardio-Thoracic Surgery (EACTS): The Task Force for the diagnosis and management of atrial fibrillation of the European Society of Cardiology (ESC) Developed with the special contribution of the European Heart Rhythm Association (EHRA) of the ESC. Eur Heart J 2021; 42:373–498.
1165. Benjamin EJ, Muntner P, Alonso A, Bittencourt MS, Callaway CW, Carson AP, et al. Heart Disease and Stroke Statistics—2019 Update: A Report From the American Heart Association. Circulation 2019; 139:e56–e528.
1166. Kreutz R, Schmidt IM, Dräger D, Brüggen F, Hörter S, Zwillich C, et al. Atrial fibrillation and medication treatment among centenarians: Are all very old patients treated the same? Geriatrics & gerontology international 2018; 18:1634–1640.
1167. Kallistratos MS, Poulimenos LE, Manolis AJ. Atrial fibrillation and arterial hypertension. Pharmacological research 2018; 128:322–326.
1168. Grundvold I, Skretteberg PT, Liestol K, Erikssen G, Kjeldsen SE, Arnesen H, et al. Upper normal blood pressures predict incident atrial fibrillation in healthy middle-aged men: a 35-year follow-up study. Hypertension (Dallas, Tex: 1979) 2012; 59:198–204.
1169. Conen D, Tedrow UB, Koplan BA, Glynn RJ, Buring JE, Albert CM. Influence of systolic and diastolic blood pressure on the risk of incident atrial fibrillation in women. Circulation 2009; 119:2146–2152.
1170. Kim D, Yang PS, Kim TH, Jang E, Shin H, Kim HY, et al. Ideal blood pressure in patients with atrial fibrillation. J Am Coll Cardiol 2018; 72:1233–1245.
1171. Kollias A, Kyriakoulis KG, Stambolliu E, Stergiou GS. Prognostic value of office blood pressure measurement in patients with atrial fibrillation on anticoagulation therapy: systematic review and meta-analysis. J Hypertens 2020; 38:13–20.
1172. Ruff CT, Giugliano RP, Braunwald E, Hoffman EB, Deenadayalu N, Ezekowitz MD, et al. Comparison of the efficacy and safety of new oral anticoagulants with warfarin in patients with atrial fibrillation: a meta-analysis of randomised trials. Lancet 2014; 383:955–962.
1173. Stergiou GS, Kyriakoulis KG, Stambolliu E, Destounis A, Karpettas N, Kalogeropoulos P, et al. Blood pressure measurement in atrial fibrillation: review and meta-analysis of evidence on accuracy and clinical relevance. J Hypertens 2019; 37:2430–2441.
1174. Kollias A, Destounis A, Kalogeropoulos P, Kyriakoulis KG, Ntineri A, Stergiou GS. Atrial fibrillation detection during 24-hour ambulatory blood pressure monitoring: comparison with 24-hour electrocardiography. Hypertension (Dallas, Tex: 1979) 2018; 72:110–115.
1175. Halfon M, Wuerzner G, Marques-Vidal P, Taffe P, Vaucher J, Waeber B, et al. Use of oscillometric devices in atrial fibrillation: a comparison of three devices and invasive blood pressure measurement. Blood Press 2018; 27:48–55.
1176. Verberk WJ, Omboni S, Kollias A, Stergiou GS. Screening for atrial fibrillation with automated blood pressure measurement: Research evidence and practice recommendations. Int J Cardiol 2016; 203:465–473.
1177. Tang EWL, Yip BHK, Yu CP, Wong SYS, Lee EKP. Sensitivity and specificity of automated blood pressure devices to detect atrial fibrillation: A systematic review and meta-analysis of diagnostic accuracy. Front Cardiovasc Med 2022; 9:956542.
1178. Lowe A, Oh TH, Stewart R. Screening for Atrial Fibrillation During Automatic Blood Pressure Measurements. IEEE J Transl Eng Health Med 2018; 6:4400307.
1179. Brandes A, Stavrakis S, Freedman B, Antoniou S, Boriani G, Camm AJ, et al. Consumer-Led Screening for Atrial Fibrillation: Frontier Review of the AF-SCREEN International Collaboration. Circulation 2022; 146:1461–1474.
1180. Van Gelder IC, Rienstra M, Crijns HJ, Olshansky B. Rate control in atrial fibrillation. Lancet 2016; 388:818–828.
1181. Lafuente-Lafuente C, Valembois L, Bergmann JF, Belmin J. Antiarrhythmics for maintaining sinus rhythm after cardioversion of atrial fibrillation. Cochrane Database Syst Rev 2015; Cd005049.
1182. Wachtell K, Lehto M, Gerdts E, Olsen MH, Hornestam B, Dahlöf B, et al. Angiotensin II receptor blockade reduces new-onset atrial fibrillation and subsequent stroke compared to atenolol: the Losartan Intervention For End Point Reduction in Hypertension (LIFE) study. J Am Coll Cardiol 2005; 45:712–719.
1183. Ducharme A, Swedberg K, Pfeffer MA, Cohen-Solal A, Granger CB, Maggioni AP, et al. Prevention of atrial fibrillation in patients with symptomatic chronic heart failure by candesartan in the Candesartan in Heart failure: Assessment of Reduction in Mortality and morbidity (CHARM) program. American heart journal 2006; 152:86–92.
1184. Schneider MP, Hua TA, Böhm M, Wachtell K, Kjeldsen SE, Schmieder RE. Prevention of atrial fibrillation by Renin-Angiotensin system inhibition a meta-analysis. J Am Coll Cardiol 2010; 55:2299–2307.
1185. Okin PM, Wachtell K, Devereux RB, Harris KE, Jern S, Kjeldsen SE, et al. Regression of electrocardiographic left ventricular hypertrophy and decreased incidence of new-onset atrial fibrillation in patients with hypertension. JAMA 2006; 296:1242–1248.
1186. Cikes M, Claggett B, Shah AM, Desai AS, Lewis EF, Shah SJ, et al. Atrial Fibrillation in Heart Failure With Preserved Ejection Fraction: The TOPCAT Trial. JACC Heart Fail 2018; 6:689–697.
1187. Goette A, Schön N, Kirchhof P, Breithardt G, Fetsch T, Häusler KG, et al. Angiotensin II-antagonist in paroxysmal atrial fibrillation (ANTIPAF) trial. Circulation Arrhythmia and electrophysiology 2012; 5:43–51.
1188. Tveit A, Grundvold I, Olufsen M, Seljeflot I, Abdelnoor M, Arnesen H, et al. Candesartan in the prevention of relapsing atrial fibrillation. Int J Cardiol 2007; 120:85–91.
1189. Swedberg K, Zannad F, McMurray JJ, Krum H, van Veldhuisen DJ, Shi H, et al. Eplerenone and atrial fibrillation in mild systolic heart failure: results from the EMPHASIS-HF (Eplerenone in Mild Patients Hospitalization And SurvIval Study in Heart Failure) study. J Am Coll Cardiol 2012; 59:1598–1603.
1190. Zheng RJ, Wang Y, Tang JN, Duan JY, Yuan MY, Zhang JY. Association of SGLT2 Inhibitors With Risk of Atrial Fibrillation and Stroke in Patients With and Without Type 2 Diabetes: A Systemic Review and Meta-Analysis of Randomized Controlled Trials. J Cardiovasc Pharmacol 2022; 79:e145–e152.
1191. Pandey AK, Okaj I, Kaur H, Belley-Cote EP, Wang J, Oraii A, et al. Sodium-Glucose Co-Transporter Inhibitors and Atrial Fibrillation: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. J Am Heart Assoc 2021; 10:e022222.
1192. Zhuo M, D’Andrea E, Paik JM, Wexler DJ, Everett BM, Glynn RJ, et al. Association of Sodium-Glucose Cotransporter-2 Inhibitors With Incident Atrial Fibrillation in Older Adults With Type 2 Diabetes. JAMA Network Open 2022; 5:e2235995–e2235995.
1193. Manolis A, Doumas M, Poulimenos L, Kallistratos M, Mancia G. The unappreciated importance of blood pressure in recent and older atrial fibrillation trials. J Hypertens 2013; 31:2109–2117. discussion 2117.
1194. Antikainen RL, Peters R, Beckett NS, Rajkumar C, Bulpitt CJ. Atrial fibrillation and the risk of cardiovascular disease and mortality in the Hypertension in the Very Elderly Trial. J Hypertens 2020; 38:839–844.
1195. Ma L, Hu X, Song L, Chen X, Ouyang M, Billot L, et al. The third Intensive Care Bundle with Blood Pressure Reduction in Acute Cerebral Haemorrhage Trial (INTERACT3): an international, stepped wedge cluster randomised controlled trial. Lancet 2023; 402:27–40.
1196. Arima H, Chalmers J. PROGRESS: Prevention of Recurrent Stroke. J Clin Hypertens (Greenwich) 2011; 13:693–702.
1197. Lip GY, Frison L, Grind M. Effect of hypertension on anticoagulated patients with atrial fibrillation. Eur Heart J 2007; 28:752–759.
1198. Rao MP, Halvorsen S, Wojdyla D, Thomas L, Alexander JH, Hylek EM, et al. Blood Pressure Control and Risk of Stroke or Systemic Embolism in Patients With Atrial Fibrillation: Results From the Apixaban for Reduction in Stroke and Other Thromboembolic Events in Atrial Fibrillation (ARISTOTLE) Trial. J Am Heart Assoc 2015; 4:
1199. Pisters R, Lane DA, Nieuwlaat R, de Vos CB, Crijns HJ, Lip GY. A novel user-friendly score (HAS-BLED) to assess 1-year risk of major bleeding in patients with atrial fibrillation: the Euro Heart Survey. Chest 2010; 138:1093–1100.
1200. Badheka AO, Patel NJ, Grover PM, Shah N, Patel N, Singh V, et al. Optimal blood pressure in patients with atrial fibrillation (from the AFFIRM Trial). Am J Cardiol 2014; 114:727–736.
1201. Böhm M, Brueckmann M, Eikelboom JW, Ezekowitz M, Fräßdorf M, Hijazi Z, et al. Cardiovascular outcomes, bleeding risk, and achieved blood pressure in patients on long-term anticoagulation with the thrombin antagonist dabigatran or warfarin: data from the RE-LY trial. Eur Heart J 2020; 41:2848–2859.
1202. Kreutz R. The role of blood pressure control in hypertensive patients with atrial fibrillation. J Hypertens Res 2021; 7:1–3.
1203. Washam JB, Hellkamp AS, Lokhnygina Y, Piccini JP, Berkowitz SD, Nessel CC, et al. Efficacy and Safety of Rivaroxaban Versus Warfarin in Patients Taking Nondihydropyridine Calcium Channel Blockers for Atrial Fibrillation (from the ROCKET AF Trial). Am J Cardiol 2017; 120:588–594.
1204. Hanigan S, Das J, Pogue K, Barnes GD, Dorsch MP. The real world use of combined P-glycoprotein and moderate CYP3A4 inhibitors with rivaroxaban or apixaban increases bleeding. Journal of thrombosis and thrombolysis 2020; 49:636–643.
1205. Rossebo AB, Pedersen TR, Boman K, Brudi P, Chambers JB, Egstrup K, et al. Intensive lipid lowering with simvastatin and ezetimibe in aortic stenosis. N Engl J Med 2008; 359:1343–1356.
1206. Nielsen OW, Sajadieh A, Sabbah M, Greve AM, Olsen MH, Boman K, et al. Assessing Optimal Blood Pressure in Patients With Asymptomatic Aortic Valve Stenosis: The Simvastatin Ezetimibe in Aortic Stenosis Study (SEAS). Circulation 2016; 134:455–468.
1207. Sen J, Chung E, Neil C, Marwick T. Antihypertensive therapies in moderate or severe aortic stenosis: a systematic review and meta-analysis. BMJ Open 2020; 10:e036960.
1208. Mancusi C, de Simone G, Brguljan Hitij J, Sudano I, Mahfoud F, Parati G, et al. Management of patients with combined arterial hypertension and aortic valve stenosis: a consensus document from the Council on Hypertension and Council on Valvular Heart Disease of the European Society of Cardiology, the European Association of Cardiovascular Imaging (EACVI), and the European Association of Percutaneous Cardiovascular Interventions (EAPCI). Eur Heart J Cardiovasc Pharmacother 2021; 7:242–250.
1209. Scognamiglio R, Rahimtoola SH, Fasoli G, Nistri S, Dalla Volta S. Nifedipine in asymptomatic patients with severe aortic regurgitation and normal left ventricular function. N Engl J Med 1994; 331:689–694.
1210. Rahimi K, Mohseni H, Otto CM, Conrad N, Tran J, Nazarzadeh M, et al. Elevated blood pressure and risk of mitral regurgitation: A longitudinal cohort study of 5.5 million United Kingdom adults. PLoS Med 2017; 14:e1002404.
1211. Rodriguez-Luna D, Pineiro S, Rubiera M, Ribo M, Coscojuela P, Pagola J, et al. Impact of blood pressure changes and course on hematoma growth in acute intracerebral hemorrhage. Eur J Neurol 2013; 20:1277–1283.
1212. Sakamoto Y, Koga M, Yamagami H, Okuda S, Okada Y, Kimura K, et al. Systolic blood pressure after intravenous antihypertensive treatment and clinical outcomes in hyperacute intracerebral hemorrhage: the stroke acute management with urgent risk-factor assessment and improvement-intracerebral hemorrhage study. Stroke 2013; 44:1846–1851.
1213. Anderson CS, Heeley E, Huang Y, Wang J, Stapf C, Delcourt C, et al. Rapid blood-pressure lowering in patients with acute intracerebral hemorrhage. N Engl J Med 2013; 368:2355–2365.
1214. Qureshi AI, Palesch YY, Barsan WG, Hanley DF, Hsu CY, Martin RL, et al. Intensive Blood-Pressure Lowering in Patients with Acute Cerebral Hemorrhage. N Engl J Med 2016; 375:1033–1043.
1215. Moullaali TJ, Wang X, Martin RH, Shipes VB, Robinson TG, Chalmers J, et al. Blood pressure control and clinical outcomes in acute intracerebral haemorrhage: a preplanned pooled analysis of individual participant data. Lancet Neurol 2019; 18:857–864.
1216. Sandset EC, Anderson CS, Bath PM, Christensen H, Fischer U, Gasecki D, et al. European Stroke Organisation (ESO) guidelines on blood pressure management in acute ischaemic stroke and intracerebral haemorrhage. Eur Stroke J 2021; 6: XLVIII-LXXXIX.
1217. Wang X, Di Tanna GL, Moullaali TJ, Martin RH, Shipes VB, Robinson TG, et al. J-shape relation of blood pressure reduction and outcome in acute intracerebral hemorrhage: A pooled analysis of INTERACT2 and ATACH-II individual participant data. Int J Stroke 2022; 17:1129–1136.
1218. Moullaali TJ, Wang X, Sandset EC, Woodhouse LJ, Law ZK, Arima H, et al. Early lowering of blood pressure after acute intracerebral haemorrhage: a systematic review and meta-analysis of individual patient data. J Neurol Neurosurg Psychiatry 2022; 93:6–13.
1219. Tsivgoulis G, Katsanos AH, Butcher KS, Boviatsis E, Triantafyllou N, Rizos I, et al. Intensive blood pressure reduction in acute intracerebral hemorrhage: a meta-analysis. Neurology 2014; 83:1523–1529.
1220. Qureshi AI, Huang W, Lobanova I, Barsan WG, Hanley DF, Hsu CY, et al. Outcomes of Intensive Systolic Blood Pressure Reduction in Patients With Intracerebral Hemorrhage and Excessively High Initial Systolic Blood Pressure: Post Hoc Analysis of a Randomized Clinical Trial. JAMA Neurol 2020; 77:1355–1365.
1221. Britton M, Carlsson A, de Faire U. Blood pressure course in patients with acute stroke and matched controls. Stroke 1986; 17:861–864.
1222. Ankolekar S, Fuller M, Cross I, Renton C, Cox P, Sprigg N, et al. Feasibility of an ambulance-based stroke trial, and safety of glyceryl trinitrate in ultra-acute stroke: the rapid intervention with glyceryl trinitrate in Hypertensive Stroke Trial (RIGHT, ISRCTN66434824). Stroke 2013; 44:3120–3128.
1223. Investigators R. Prehospital transdermal glyceryl trinitrate in patients with ultra-acute presumed stroke (RIGHT-2): an ambulance-based, randomised, sham-controlled, blinded, phase 3 trial. Lancet 2019; 393:1009–1020.
1224. Bath PM, Martin RH, Palesch Y, Cotton D, Yusuf S, Sacco R, et al. Effect of telmisartan on functional outcome, recurrence, and blood pressure in patients with acute mild ischemic stroke: a PRoFESS subgroup analysis. Stroke 2009; 40:3541–3546.
1225. Sandset EC, Bath PM, Boysen G, Jatuzis D, Korv J, Luders S, et al. The angiotensin-receptor blocker candesartan for treatment of acute stroke (SCAST): a randomised, placebo-controlled, double-blind trial. Lancet 2011; 377:741–750.
1226. Zhang R, Zhong C, Zhang Y, Xie X, Zhu Z, Wang A, et al. Immediate Antihypertensive Treatment for Patients With Acute Ischemic Stroke With or Without History of Hypertension: A Secondary Analysis of the CATIS Randomized Clinical Trial. JAMA Netw Open 2019; 2:e198103.
1227. Lee M, Ovbiagele B, Hong KS, Wu YL, Lee JE, Rao NM, et al. Effect of Blood Pressure Lowering in Early Ischemic Stroke: Meta-Analysis. Stroke 2015; 46:1883–1889.
1228. Zhao R, Liu FD, Wang S, Peng JL, Tao XX, Zheng B, et al. Blood pressure reduction in the acute phase of an ischemic stroke does not improve short- or long-term dependency or mortality: a meta-analysis of current literature. Medicine (Baltimore) 2015; 94:e896.
1229. Gasecki D, Coca A, Cunha P, Hering D, Manios E, Lovic D, et al. Blood pressure in acute ischemic stroke: challenges in trial interpretation and clinical management: position of the ESH Working Group on Hypertension and the Brain. J Hypertens 2018; 36:1212–1221.
1230. Sandset EC, Murray GD, Bath PM, Kjeldsen SE, Berge E, Scandinavian Candesartan Acute Stroke Trial Study G. Relation between change in blood pressure in acute stroke and risk of early adverse events and poor outcome. Stroke 2012; 43:2108–2114.
1231. Investigators ET. Efficacy of nitric oxide, with or without continuing antihypertensive treatment, for management of high blood pressure in acute stroke (ENOS): a partial-factorial randomised controlled trial. Lancet 2015; 385:617–628.
1232. Ahmed N, Wahlgren N, Brainin M, Castillo J, Ford GA, Kaste M, et al. Relationship of blood pressure, antihypertensive therapy, and outcome in ischemic stroke treated with intravenous thrombolysis: retrospective analysis from Safe Implementation of Thrombolysis in Stroke-International Stroke Thrombolysis Register (SITS-ISTR). Stroke 2009; 40:2442–2449.
1233. Wu W, Huo X, Zhao X, Liao X, Wang C, Pan Y, et al. Relationship between Blood Pressure and Outcomes in Acute Ischemic Stroke Patients Administered Lytic Medication in the TIMS-China Study. PLoS One 2016; 11:e0144260.
1234. Berge E, Whiteley W, Audebert H, De Marchis GM, Fonseca AC, Padiglioni C, et al. European Stroke Organisation (ESO) guidelines on intravenous thrombolysis for acute ischaemic stroke. Eur Stroke J 2021; 6: I-LXII.
1235. Katsanos AH, Malhotra K, Ahmed N, Seitidis G, Mistry EA, Mavridis D, et al. Blood Pressure After Endovascular Thrombectomy and Outcomes in Patients With Acute Ischemic Stroke: An Individual Patient Data Meta-analysis. Neurology 2022; 98:e291–e301.
1236. Yang P, Song L, Zhang Y, Zhang X, Chen X, Li Y, et al. Intensive blood pressure control after endovascular thrombectomy for acute ischaemic stroke (ENCHANTED2/MT): a multicentre, open-label, blinded-endpoint, randomised controlled trial. Lancet 2022; 400:1585–1596.
1237. Mazighi M, Richard S, Lapergue B, Sibon I, Gory B, Berge J, et al. Safety and efficacy of intensive blood pressure lowering after successful endovascular therapy in acute ischaemic stroke (BP-TARGET): a multicentre, open-label, randomised controlled trial. Lancet Neurol 2021; 20:265–274.
1238. Robinson TG, Potter JF, Ford GA, Bulpitt CJ, Chernova J, Jagger C, et al. Effects of antihypertensive treatment after acute stroke in the Continue or Stop Post-Stroke Antihypertensives Collaborative Study (COSSACS): a prospective, randomised, open, blinded-endpoint trial. Lancet Neurol 2010; 9:767–775.
1239. Progress Collaborative Group. Randomised trial of a perindopril-based blood-pressure-lowering regimen among 6,105 individuals with previous stroke or transient ischaemic attack. Lancet 2001; 358:1033–1041.
1240. Pats Collaborating Group. Post-stroke antihypertensive treatment study. A preliminary result. Chin Med J (Engl) 1995; 108:710–717.
1241. Powers WJ, Rabinstein AA, Ackerson T, Adeoye OM, Bambakidis NC, Becker K, et al. Guidelines for the Early Management of Patients With Acute Ischemic Stroke: 2019 Update to the 2018 Guidelines for the Early Management of Acute Ischemic Stroke: A Guideline for Healthcare Professionals From the American Heart Association/American Stroke Association. Stroke 2019; 50:e344–e418.
1242. Mant J, McManus RJ, Roalfe A, Fletcher K, Taylor CJ, Martin U, et al. Different systolic blood pressure targets for people with history of stroke or transient ischaemic attack: PAST-BP (Prevention After Stroke--Blood Pressure) randomised controlled trial. BMJ 2016; 352:i708.
1243. Odden MC, McClure LA, Sawaya BP, White CL, Peralta CA, Field TS, et al. Achieved blood pressure and outcomes in the secondary prevention of small subcortical strokes trial. Hypertension (Dallas, Tex: 1979) 2016; 67:63–69.
1244. Kernan WN, Ovbiagele B, Black HR, Bravata DM, Chimowitz MI, Ezekowitz MD, et al. Guidelines for the prevention of stroke in patients with stroke and transient ischemic attack: a guideline for healthcare professionals from the American Heart Association/American Stroke Association. Stroke 2014; 45:2160–2236.
1245. Arima H, Chalmers J, Woodward M, Anderson C, Rodgers A, Davis S, et al. Lower target blood pressures are safe and effective for the prevention of recurrent stroke: the PROGRESS trial. J Hypertens 2006; 24:1201–1208.
1246. Group SPSS, Benavente OR, Coffey CS, Conwit R, Hart RG, McClure LA, et al. Blood-pressure targets in patients with recent lacunar stroke: the SPS3 randomised trial. Lancet 2013; 382:507–515.
1247. Bath PM, Scutt P, Blackburn DJ, Ankolekar S, Krishnan K, Ballard C, et al. Intensive versus guideline blood pressure and lipid lowering in patients with previous stroke: main results from the Pilot ‘Prevention of Decline in Cognition after Stroke Trial’ (PODCAST) Randomised Controlled Trial. PLoS One 2017; 12:e0164608.
1248. Kitagawa K, Yamamoto Y, Arima H, Maeda T, Sunami N, Kanzawa T, et al. Effect of standard vs intensive blood pressure control on the risk of recurrent stroke: a randomized clinical trial and meta-analysis. JAMA Neurol 2019; 76:1309–1318.
1249. Katsanos AH, Filippatou A, Manios E, Deftereos S, Parissis J, Frogoudaki A, et al. Blood Pressure reduction and secondary stroke prevention: a systematic review and metaregression analysis of randomized clinical trials. Hypertension (Dallas, Tex: 1979) 2017; 69:171–179.
1250. Dawson J, Bejot Y, Christensen LM, De Marchis GM, Dichgans M, Hagberg G, et al. European Stroke Organisation (ESO) guideline on pharmacological interventions for long-term secondary prevention after ischaemic stroke or transient ischaemic attack. Eur Stroke J 2022; 7:I–II.
1251. Collier DJ, Poulter NR, Dahlof B, Sever PS, Wedel H, Buch J, et al. Impact of amlodipine-based therapy among older and younger patients in the Anglo-Scandinavian Cardiac Outcomes Trial-Blood Pressure Lowering Arm (ASCOT-BPLA). J Hypertens 2011; 29:583–591.
1252. Excellence. NIfHaC. National Institute for Health and Clinical Excellence. Hypertension (CG127): clinical management of primary hypertension in adults www.nice.org.uk/guidance/CG127 (April 2018). www.nice.org.uk/guidance/CG1272018.
1253. Zanchetti A, Thomopoulos C, Parati G. Randomized controlled trials of blood pressure lowering in hypertension: a critical reappraisal. Circ Res 2015; 116:1058–1073.
1254. Thomopoulos C, Parati G, Zanchetti A. Effects of blood pressure lowering on outcome incidence in hypertension: 4. Effects of various classes of antihypertensive drugs--overview and meta-analyses. J Hypertens 2015; 33:195–211.
1255. Vickrey BG, Rector TS, Wickstrom SL, Guzy PM, Sloss EM, Gorelick PB, et al. Occurrence of secondary ischemic events among persons with atherosclerotic vascular disease. Stroke 2002; 33:901–906.
1256. Bohm M, Schumacher H, Teo KK, Lonn EM, Lauder L, Mancia G, et al. Cardiovascular outcomes in patients at high cardiovascular risk with previous myocardial infarction or stroke. J Hypertens 2021; 39:1602–1610.
1257. Wang WT, You LK, Chiang CE, Sung SH, Chuang SY, Cheng HM, et al. Comparative Effectiveness of Blood Pressure-lowering Drugs in Patients who have Already Suffered From Stroke: Traditional and Bayesian Network Meta-analysis of Randomized Trials. Medicine (Baltimore) 2016; 95:e3302.
1258. Collaborators GBDD Global, regional, and national burden of Alzheimer's disease and other dementias, 1990–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol 2019; 18:88–106.
1259. Estimation of the global prevalence of dementia in 2019 and forecasted prevalence in 2050: an analysis for the Global Burden of Disease Study 2019. Lancet Public Health 2022; 7:e105–e125.
1260. Emdin CA, Rothwell PM, Salimi-Khorshidi G, Kiran A, Conrad N, Callender T, et al. Blood Pressure and Risk of Vascular Dementia: Evidence From a Primary Care Registry and a Cohort Study of Transient Ischemic Attack and Stroke. Stroke 2016; 47:1429–1435.
1261. Siedlinski M, Carnevale L, Xu X, Carnevale D, Evangelou E, Caulfield MJ, et al. Genetic analyses identify brain structures related to cognitive impairment associated with elevated blood pressure. Eur Heart J 2023; 44:2114–2125.
1262. Sierra C, De La Sierra A, Salamero M, Sobrino J, Gomez-Angelats E, Coca A. Silent cerebral white matter lesions and cognitive function in middle-aged essential hypertensive patients. Am J Hypertens 2004; 17:529–534.
1263. Snyder HM, Corriveau RA, Craft S, Faber JE, Greenberg SM, Knopman D, et al. Vascular contributions to cognitive impairment and dementia including Alzheimer's disease. Alzheimers Dement 2015; 11:710–717.
1264. de Roos A, van der Grond J, Mitchell G, Westenberg J. Magnetic Resonance Imaging of Cardiovascular Function and the Brain: Is Dementia a Cardiovascular-Driven Disease? Circulation 2017; 135:2178–2195.
1265. Peters R, Xu Y, Fitzgerald O, Aung HL, Beckett N, Bulpitt C, et al. Blood pressure lowering and prevention of dementia: an individual patient data meta-analysis. Eur Heart J 2022; 43:4980–4990.
1266. Godin O, Tzourio C, Maillard P, Mazoyer B, Dufouil C. Antihypertensive treatment and change in blood pressure are associated with the progression of white matter lesion volumes: the Three-City (3C)-Dijon Magnetic Resonance Imaging Study. Circulation 2011; 123:266–273.
1267. Verhaaren BF, Vernooij MW, de Boer R, Hofman A, Niessen WJ, van der Lugt A, et al. High blood pressure and cerebral white matter lesion progression in the general population. Hypertension (Dallas, Tex: 1979) 2013; 61:1354–1359.
1268. Chiu WC, Ho WC, Lin MH, Lee HH, Yeh YC, Wang JD, et al. Angiotension receptor blockers reduce the risk of dementia. J Hypertens 2014; 32:938–947.
1269. van Middelaar T, van Vught LA, Moll van Charante EP, Eurelings LSM, Ligthart SA, van Dalen JW, et al. Lower dementia risk with different classes of antihypertensive medication in older patients. J Hypertens 2017; 35:2095–2101.
1270. van Dalen JW, Marcum ZA, Gray SL, Barthold D, Moll van Charante EP, van Gool WA, et al. Association of Angiotensin II-Stimulating Antihypertensive Use and Dementia Risk: Post Hoc Analysis of the PreDIVA Trial. Neurology 2021; 96:e67–e80.
1271. Marcum ZA, Cohen JB, Zhang C, Derington CG, Greene TH, Ghazi L, et al. Association of Antihypertensives That Stimulate vs Inhibit Types 2 and 4 Angiotensin II Receptors With Cognitive Impairment. JAMA Netw Open 2022; 5:e2145319.
1272. Emdin CA, Anderson SG, Callender T, Conrad N, Salimi-Khorshidi G, Mohseni H, et al. Usual blood pressure, peripheral arterial disease, and vascular risk: cohort study of 4.2 million adults. BMJ 2015; 351:h4865.
1273. Aboyans V, Ricco JB, Bartelink MEL, Bjorck M, Brodmann M, Cohnert T, et al. 2017 ESC Guidelines on the Diagnosis and Treatment of Peripheral Arterial Diseases, in collaboration with the European Society for Vascular Surgery (ESVS): Document covering atherosclerotic disease of extracranial carotid and vertebral, mesenteric, renal, upper and lower extremity arteriesEndorsed by: the European Stroke Organization (ESO)The Task Force for the Diagnosis and Treatment of Peripheral Arterial Diseases of the European Society of Cardiology (ESC) and of the European Society for Vascular Surgery (ESVS). Eur Heart J 2018; 39:763–816.
1274. Rapsomaniki E, Timmis A, George J, Pujades-Rodriguez M, Shah AD, Denaxas S, et al. Blood pressure and incidence of twelve cardiovascular diseases: lifetime risks, healthy life-years lost, and age-specific associations in 1.25 million people. Lancet 2014; 383:1899–1911.
1275. Ya’qoub L, Peri-Okonny P, Wang J, Patel KK, Stone N, Smolderen K. Blood pressure management in patients with symptomatic peripheral artery disease: insights from the PORTRAIT registry. Eur Heart J Qual Care Clin Outcomes 2019; 5:79–81.
1276. Itoga NK, Tawfik DS, Lee CK, Maruyama S, Leeper NJ, Chang TI. Association of Blood Pressure Measurements With Peripheral Artery Disease Events. Circulation 2018; 138:1805–1814.
1277. Piller LB, Simpson LM, Baraniuk S, Habib GB, Rahman M, Basile JN, et al. Characteristics and long-term follow-up of participants with peripheral arterial disease during ALLHAT. J Gen Intern Med 2014; 29:1475–1483.
1278. Paravastu SC, Mendonca DA, da Silva A. Beta blockers for peripheral arterial disease. Eur J Vasc Endovasc Surg 2009; 38:66–70.
1279. Radack K, Deck C. Beta-adrenergic blocker therapy does not worsen intermittent claudication in subjects with peripheral arterial disease. A meta-analysis of randomized controlled trials. Arch Intern Med 1991; 151:1769–1776.
1280. Espinola-Klein C, Weisser G, Jagodzinski A, Savvidis S, Warnholtz A, Ostad MA, et al. beta-Blockers in patients with intermittent claudication and arterial hypertension: results from the nebivolol or metoprolol in arterial occlusive disease trial. Hypertension (Dallas, Tex: 1979) 2011; 58:148–154.
1281. Fudim M, Hopley CW, Huang Z, Kavanagh S, Rockhold FW, Baumgartner I, et al. Association of Hypertension and Arterial Blood Pressure on Limb and Cardiovascular Outcomes in Symptomatic Peripheral Artery Disease: The EUCLID Trial. Circ Cardiovasc Qual Outcomes 2020; 13:e006512.
1282. Kaplovitch E, Eikelboom JW, Dyal L, Aboyans V, Abola MT, Verhamme P, et al. Rivaroxaban and Aspirin in Patients With Symptomatic Lower Extremity Peripheral Artery Disease: A Subanalysis of the COMPASS Randomized Clinical Trial. JAMA Cardiol 2021; 6:21–29.
1283. Erbel R, Aboyans V, Boileau C, Bossone E, Bartolomeo RD, Eggebrecht H, et al. 2014 ESC Guidelines on the diagnosis and treatment of aortic diseases: Document covering acute and chronic aortic diseases of the thoracic and abdominal aorta of the adult. The Task Force for the Diagnosis and Treatment of Aortic Diseases of the European Society of Cardiology (ESC). Eur Heart J 2014; 35:2873–2926.
1284. Groenink M, den Hartog AW, Franken R, Radonic T, de Waard V, Timmermans J, et al. Losartan reduces aortic dilatation rate in adults with Marfan syndrome: a randomized controlled trial. Eur Heart J 2013; 34:3491–3500.
1285. Shores J, Berger KR, Murphy EA, Pyeritz RE. Progression of aortic dilatation and the benefit of long-term beta-adrenergic blockade in Marfan's syndrome. N Engl J Med 1994; 330:1335–1341.
1286. Teixido-Tura G, Forteza A, Rodriguez-Palomares J, Gonzalez Mirelis J, Gutierrez L, Sanchez V, et al. Losartan Versus Atenolol for Prevention of Aortic Dilation in Patients With Marfan Syndrome. J Am Coll Cardiol 2018; 72:1613–1618.
1287. Milleron O, Arnoult F, Ropers J, Aegerter P, Detaint D, Delorme G, et al. Marfan Sartan: a randomized, double-blind, placebo-controlled trial. Eur Heart J 2015; 36:2160–2166.
1288. Pitcher A, Spata E, Emberson J, Davies K, Halls H, Holland L, et al. Angiotensin receptor blockers and beta blockers in Marfan syndrome: an individual patient data meta-analysis of randomised trials. Lancet 2022; 400:822–831.
1289. Baumgartner H, De Backer J, Babu-Narayan SV, Budts W, Chessa M, Diller GP, et al. 2020 ESC Guidelines for the management of adult congenital heart disease. Eur Heart J 2021; 42:563–645.
1290. Hibino M, Otaki Y, Kobeissi E, Pan H, Hibino H, Taddese H, et al. Blood Pressure, Hypertension, and the Risk of Aortic Dissection Incidence and Mortality: Results From the J-SCH Study, the UK Biobank Study, and a Meta-Analysis of Cohort Studies. Circulation 2022; 145:633–644.
1291. Chen SW, Chan YH, Lin CP, Wu VC, Cheng YT, Chen DY, et al. Association of Long-term Use of Antihypertensive Medications With Late Outcomes Among Patients With Aortic Dissection. JAMA Netw Open 2021; 4:e210469.
1292. Sweeting MJ, Thompson SG, Brown LC, Powell JT, collaborators R. Meta-analysis of individual patient data to examine factors affecting growth and rupture of small abdominal aortic aneurysms. Br J Surg 2012; 99:655–665.
1293. Bahia SS, Vidal-Diez A, Seshasai SR, Shpitser I, Brownrigg JR, Patterson BO, et al. Cardiovascular risk prevention and all-cause mortality in primary care patients with an abdominal aortic aneurysm. Br J Surg 2016; 103:1626–1633.
1294. Golledge J, Singh TP. Effect of blood pressure lowering drugs and antibiotics on abdominal aortic aneurysm growth: a systematic review and meta-analysis. Heart 2021; 107:1465–1471.
1295. Guessous I, Periard D, Lorenzetti D, Cornuz J, Ghali WA. The efficacy of pharmacotherapy for decreasing the expansion rate of abdominal aortic aneurysms: a systematic review and meta-analysis. PLoS One 2008; 3:e1895.
1296. Hackam DG, Thiruchelvam D, Redelmeier DA. Angiotensin-converting enzyme inhibitors and aortic rupture: a population-based case-control study. Lancet 2006; 368:659–665.
1297. Thompson A, Cooper JA, Fabricius M, Humphries SE, Ashton HA, Hafez H. An analysis of drug modulation of abdominal aortic aneurysm growth through 25 years of surveillance. J Vasc Surg 2010; 52:55–61. e52.
1298. Downie ML, Ulrich EH, Noone DG. An update on hypertension in children with type 1 diabetes. Can J Diabetes 2018; 42:199–204.
1299. Nørgaard K, Feldt-Rasmussen B, Borch-Johnsen K, Saelan H, Deckert T. Prevalence of hypertension in type 1 (insulin-dependent) diabetes mellitus. Diabetologia 1990; 33:407–410.
1300. Collado-Mesa F, Colhoun HM, Stevens LK, Boavida J, Ferriss JB, Karamanos B, et al. Prevalence and management of hypertension in type 1 diabetes mellitus in Europe: the EURODIAB IDDM Complications Study. Diabet Med 1999; 16:41–48.
1301. Whalen KL, Stewart RD. Pharmacologic management of hypertension in patients with diabetes. Am Fam Physician 2008; 78:1277–1282.
1302. Grassi G, Biffi A, Dell’Oro R, Quarti Trevano F, Seravalle G, Corrao G, et al. Sympathetic neural abnormalities in type 1 and type 2 diabetes: a systematic review and meta-analysis. J Hypertens 2020; 38:1436–1442.
1303. Jia G, Sowers JR. Hypertension in diabetes: an update of basic mechanisms and clinical disease. Hypertension (Dallas, Tex: 1979) 2021; 78:1197–1205.
1304. Mancia G, Facchetti R, Bombelli M, Polo Friz H, Grassi G, Giannattasio C, et al. Relationship of office, home, and ambulatory blood pressure to blood glucose and lipid variables in the PAMELA population. Hypertension (Dallas, Tex: 1979) 2005; 45:1072–1077.
1305. DeFronzo RA, Ferrannini E, Groop L, Henry RR, Herman WH, Holst JJ, et al. Type 2 diabetes mellitus. Nat Rev Dis Primers 2015; 1:15019.
1306. Narkiewicz K, Kjeldsen SE, Egan BM, Kreutz R, Burnier M. Masked hypertension in type 2 diabetes: never take normotension for granted and always assess out-of-office blood pressure. Blood Press 2022; 31:207–209.
1307. Brunström M, Carlberg B. Effect of antihypertensive treatment at different blood pressure levels in patients with diabetes mellitus: systematic review and meta-analyses. Bmj 2016; 352:i717.
1308. Thomopoulos C, Parati G, Zanchetti A. Effects of blood-pressure-lowering treatment on outcome incidence in hypertension: 10 - Should blood pressure management differ in hypertensive patients with and without diabetes mellitus? Overview and meta-analyses of randomized trials. J Hypertens 2017; 35:922–944.
1309. Emdin CA, Rahimi K, Neal B, Callender T, Perkovic V, Patel A. Blood pressure lowering in type 2 diabetes: a systematic review and meta-analysis. JAMA 2015; 313:603–615.
1310. Beddhu S, Chertow GM, Greene T, Whelton PK, Ambrosius WT, Cheung AK, et al. Effects of Intensive Systolic Blood Pressure Lowering on Cardiovascular Events and Mortality in Patients With Type 2 Diabetes Mellitus on Standard Glycemic Control and in Those Without Diabetes Mellitus: Reconciling Results From ACCORD BP and SPRINT. J Am Heart Assoc 2018; 7:e009326.
1311. Buckley LF, Dixon DL, Wohlford GFt, Wijesinghe DS, Baker WL, Van Tassell BW. Intensive Versus Standard Blood Pressure Control in SPRINT-Eligible Participants of ACCORD-BP. Diabetes Care 2017; 40:1733–1738.
1312. Bangalore S, Kumar S, Lobach I, Messerli FH. Blood pressure targets in subjects with type 2 diabetes mellitus/impaired fasting glucose: observations from traditional and bayesian random-effects meta-analyses of randomized trials. Circulation 2011; 123:2799–2810.
1313. Ilkun OL, Greene T, Cheung AK, Whelton PK, Wei G, Boucher RE, et al. The Influence of Baseline Diastolic Blood Pressure on the Effects of Intensive Blood Pressure Lowering on Cardiovascular Outcomes and All-Cause Mortality in Type 2 Diabetes. Diabetes Care 2020; 43:1878–1884.
1314. Olsen E, Holzhauer B, Julius S, Kjeldsen SE, Larstorp ACK, Mancia G, et al. Cardiovascular outcomes at recommended blood pressure targets in middle-aged and elderly patients with type 2 diabetes mellitus and hypertension. Blood Press 2021; 30:82–89.
1315. Joseph JJ, Deedwania P, Acharya T, Aguilar D, Bhatt DL, Chyun DA, et al. Comprehensive Management of Cardiovascular Risk Factors for Adults With Type 2 Diabetes: A Scientific Statement From the American Heart Association. Circulation 2022; 145:e722–e759.
1316. Shen J, Huang YM, Song XN, Hong XZ, Wang M, Ling W, et al. Protection against death and renal failure by renin-angiotensin system blockers in patients with diabetes and kidney disease. J Renin Angiotensin Aldosterone Syst 2016; 17.
1317. Caruso I, Cignarelli A, Sorice GP, Natalicchio A, Perrini S, Laviola L, et al. Cardiovascular and Renal Effectiveness of GLP-1 Receptor Agonists vs. Other Glucose-Lowering Drugs in Type 2 Diabetes: A Systematic Review and Meta-Analysis of Real-World Studies. Metabolites 2022; 12.
1318. Impact of diabetes on the effects of sodium glucose co-transporter-2 inhibitors on kidney outcomes: collaborative meta-analysis of large placebo-controlled trials. Lancet 2022; 400:1788–1801.
1319. Neal B, Perkovic V, Mahaffey KW, de Zeeuw D, Fulcher G, Erondu N, et al. Canagliflozin and Cardiovascular and Renal Events in Type 2 Diabetes. The New England Journal of Medicine 2017; 377:644–657.
1320. Perkovic V, Jardine MJ, Neal B, Bompoint S, Heerspink HJL, Charytan DM, et al. Canagliflozin and Renal Outcomes in Type 2 Diabetes and Nephropathy. The New England Journal of Medicine 2019; 380:2295–2306.
1321. Reifsnider OS, Kansal AR, Wanner C, Pfarr E, Koitka-Weber A, Brand SB, et al. Cost-Effectiveness of Empagliflozin in Patients With Diabetic Kidney Disease in the United States: Findings Based on the EMPA-REG OUTCOME Trial. Am J Kidney Dis 2022; 79:796–806.
1322. Cosentino F, Grant PJ, Aboyans V, Bailey CJ, Ceriello A, Delgado V, et al. 2019 ESC Guidelines on diabetes, pre-diabetes, and cardiovascular diseases developed in collaboration with the EASD. Eur Heart J 2020; 41:255–323.
1323. Jastreboff AM, Aronne LJ, Ahmad NN, Wharton S, Connery L, Alves B, et al. Tirzepatide Once Weekly for the Treatment of Obesity. N Engl J Med 2022; 387:205–216.
1324. Kosiborod MN, Bhatta M, Davies M, Deanfield JE, Garvey WT, Khalid U, et al. Semaglutide improves cardiometabolic risk factors in adults with overweight or obesity: STEP 1 and 4 exploratory analyses. Diabetes Obes Metab 2023; 25:468–478.
1325. Klag MJ, Whelton PK, Randall BL, Neaton JD, Brancati FL, Ford CE, et al. Blood pressure and end-stage renal disease in men. The New England Journal of Medicine 1996; 334:13–18.
1326. Loutradis C, Sarafidis P. Hypertension in patients with advanced chronic kidney disease. In: Mancia G, Grassi G, Tsioufis K, Dominiczak A, Rosei EA, editors. Manual of Hypertension of the European Society of Hypertension. Boca Raton: CRC Press; 2019.
1327. Boenink R, Astley ME, Huijben JA, Stel VS, Kerschbaum J, Ots-Rosenberg M, et al. The ERA Registry Annual Report 2019: summary and age comparisons. Clinical Kidney Journal 2022; 15:452–472.
1328. Johansen KL, Chertow GM, Gilbertson DT, Herzog CA, Ishani A, Israni AK, et al. US Renal Data System 2021 Annual Data Report: Epidemiology of Kidney Disease in the United States. American Journal of Kidney Diseases: The Official Journal of the National Kidney Foundation 2022; 79:A8–A12.
1329. Astor BC, Matsushita K, Gansevoort RT, van der Velde M, Woodward M, Levey AS, et al. Lower estimated glomerular filtration rate and higher albuminuria are associated with mortality and end-stage renal disease. A collaborative meta-analysis of kidney disease population cohorts. Kidney International 2011; 79:1331–1340.
1330. Fox CS, Matsushita K, Woodward M, Bilo HJG, Chalmers J, Heerspink HJL, et al. Associations of kidney disease measures with mortality and end-stage renal disease in individuals with and without diabetes: a meta-analysis. Lancet (London, England) 2012; 380:1662–1673.
1331. Levey AS, de Jong PE, Coresh J, El Nahas M, Astor BC, Matsushita K, et al. The definition, classification, and prognosis of chronic kidney disease: a KDIGO Controversies Conference report. Kidney International 2011; 80:17–28.
1332. Stevens LA, Coresh J, Greene T, Levey AS. Assessing kidney function--measured and estimated glomerular filtration rate. The New England Journal of Medicine 2006; 354:2473–2483.
1333. Drawz PE, Alper AB, Anderson AH, Brecklin CS, Charleston J, Chen J, et al. Masked Hypertension and Elevated Nighttime Blood Pressure in CKD: Prevalence and Association with Target Organ Damage. Clinical journal of the American Society of Nephrology: CJASN 2016; 11:642–652.
1334. Rossignol P, Massy ZA, Azizi M, Bakris G, Ritz E, Covic A, et al. The double challenge of resistant hypertension and chronic kidney disease. Lancet (London, England) 2015; 386:1588–1598.
1335. Ruiz-Hurtado G, Ruilope LM, de la Sierra A, Sarafidis P, de la Cruz JJ, Gorostidi M, et al. Association Between High and Very High Albuminuria and Nighttime Blood Pressure: Influence of Diabetes and Chronic Kidney Disease. Diabetes Care 2016; 39:1729–1737.
1336. Sarafidis PA, Georgianos PI, Zebekakis PE. Comparative epidemiology of resistant hypertension in chronic kidney disease and the general hypertensive population. Semin Nephrol 2014; 34:483–491.
1337. Schmieder RE. Renal denervation: where do we stand and what is the relevance to the nephrologist? Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2022; 37:638–644.
1338. Phan O, Burnier M, Wuerzner G. Hypertension in Chronic Kidney Disease - Role of Arterial Calcification and Impact on Treatment. Eur Cardiol 2014; 9:115–119.
1339. Sarafidis PA, Persu A, Agarwal R, Burnier M, de Leeuw P, Ferro C, et al. Hypertension in dialysis patients: a consensus document by the European Renal and Cardiovascular Medicine (EURECA-m) working group of the European Renal Association - European Dialysis and Transplant Association (ERA-EDTA) and the Hypertension and the Kidney working group of the European Society of Hypertension (ESH). Journal of Hypertension 2017; 35:657–676.
1340. Ninomiya T, Perkovic V, Turnbull F, Neal B, Barzi F, et al. Blood Pressure Lowering Treatment Trialists C. Blood pressure lowering and major cardiovascular events in people with and without chronic kidney disease: meta-analysis of randomised controlled trials. BMJ (Clinical research ed) 2013; 347:f5680.
1341. Sarafidis PA, Lazaridis AA, Ruiz-Hurtado G, Ruilope LM. Blood pressure reduction in diabetes: lessons from ACCORD, SPRINT and EMPA-REG OUTCOME. Nat Rev Endocrinol 2017; 13:365–374.
1342. Sarafidis PA, Ruilope LM. Aggressive blood pressure reduction and renin-angiotensin system blockade in chronic kidney disease: time for re-evaluation? Kidney International 2014; 85:536–546.
1343. Wu S, Li M, Lu J, Tang X, Wang G, Zheng R, et al. Blood Pressure Levels, Cardiovascular Events, and Renal Outcomes in Chronic Kidney Disease Without Antihypertensive Therapy: A Nationwide Population-Based Cohort Study. Hypertension (Dallas, Tex: 1979) 2023; 80:640–649.
1344. Klahr S, Levey AS, Beck GJ, Caggiula AW, Hunsicker L, Kusek JW, et al. The effects of dietary protein restriction and blood-pressure control on the progression of chronic renal disease. Modification of Diet in Renal Disease Study Group. The New England Journal of Medicine 1994; 330:877–884.
1345. Peterson JC, Adler S, Burkart JM, Greene T, Hebert LA, Hunsicker LG, et al. Blood pressure control, proteinuria, and the progression of renal disease. The Modification of Diet in Renal Disease Study. Annals of Internal Medicine 1995; 123:754–762.
1346. Wright JT, Bakris G, Greene T, Agodoa LY, Appel LJ, Charleston J, et al. Effect of blood pressure lowering and antihypertensive drug class on progression of hypertensive kidney disease: results from the AASK trial. JAMA 2002; 288:2421–2431.
1347. Lea J, Greene T, Hebert L, Lipkowitz M, Massry S, Middleton J, et al. The relationship between magnitude of proteinuria reduction and risk of end-stage renal disease: results of the African American study of kidney disease and hypertension. Archives of Internal Medicine 2005; 165:947–953.
1348. Appel LJ, Wright JT, Greene T, Agodoa LY, Astor BC, Bakris GL, et al. Intensive blood-pressure control in hypertensive chronic kidney disease. The New England Journal of Medicine 2010; 363:918–929.
1349. Cheung AK, Rahman M, Reboussin DM, Craven TE, Greene T, Kimmel PL, et al. Effects of Intensive BP Control in CKD. Journal of the American Society of Nephrology: JASN 2017; 28:2812–2823.
1350. Tight blood pressure control and risk of macrovascular and microvascular complications in type 2 diabetes: UKPDS 38. UK Prospective Diabetes Study Group. BMJ (Clinical research ed) 1998; 317:703–713.
1351. Hansson L, Zanchetti A, Carruthers SG, Dahlöf B, Elmfeldt D, Julius S, et al. Effects of intensive blood-pressure lowering and low-dose aspirin in patients with hypertension: principal results of the Hypertension Optimal Treatment (HOT) randomised trial. HOT Study Group Lancet (London, England) 1998; 351:1755–1762.
1352. Group AS, Cushman WC, Evans GW, Byington RP, Goff DC, Grimm RH, et al. Effects of intensive blood-pressure control in type 2 diabetes mellitus. The New England Journal of Medicine 2010; 362:1575–1585.
1353. Bakris GL, Weir MR, Shanifar S, Zhang Z, Douglas J, van Dijk DJ, et al. Effects of blood pressure level on progression of diabetic nephropathy: results from the RENAAL study. Archives of Internal Medicine 2003; 163:1555–1565.
1354. Pohl MA, Blumenthal S, Cordonnier DJ, De Alvaro F, Deferrari G, Eisner G, et al. Independent and additive impact of blood pressure control and angiotensin II receptor blockade on renal outcomes in the irbesartan diabetic nephropathy trial: clinical implications and limitations. Journal of the American Society of Nephrology: JASN 2005; 16:3027–3037.
1355. Malhotra R, Nguyen HA, Benavente O, Mete M, Howard BV, Mant J, et al. Association Between More Intensive vs Less Intensive Blood Pressure Lowering and Risk of Mortality in Chronic Kidney Disease Stages 3 to 5: A Systematic Review and Meta-analysis. JAMA internal medicine 2017; 177:1498–1505.
1356. Aggarwal R, Petrie B, Bala W, Chiu N. Mortality outcomes with intensive blood pressure targets in chronic kidney disease patients. Hypertension (Dallas, Tex: 1979) 2019; 73:1275–1282.
1357. Vogt L, Waanders F, Boomsma F, de Zeeuw D, Navis G. Effects of dietary sodium and hydrochlorothiazide on the antiproteinuric efficacy of losartan. Journal of the American Society of Nephrology: JASN 2008; 19:999–1007.
1358. Strippoli GF, Craig M, Deeks JJ, Schena FP, Craig JC. Effects of angiotensin converting enzyme inhibitors and angiotensin II receptor antagonists on mortality and renal outcomes in diabetic nephropathy: systematic review. BMJ 2004; 329:828.
1359. Barnett AH, Bain SC, Bouter P, Karlberg B, Madsbad S, Jervell J, et al. Angiotensin-receptor blockade versus converting-enzyme inhibition in type 2 diabetes and nephropathy. N Engl J Med 2004; 351:1952–1961.
1360. Brenner BM, Cooper ME, de Zeeuw D, Keane WF, Mitch WE, Parving HH, et al. Effects of losartan on renal and cardiovascular outcomes in patients with type 2 diabetes and nephropathy. N Engl J Med 2001; 345:861–869.
1361. Lewis EJ, Hunsicker LG, Clarke WR, Berl T, Pohl MA, Lewis JB, et al. Renoprotective effect of the angiotensin-receptor antagonist irbesartan in patients with nephropathy due to type 2 diabetes. The New England Journal of Medicine 2001; 345:851–860.
1362. Randomised placebo-controlled trial of effect of ramipril on decline in glomerular filtration rate and risk of terminal renal failure in proteinuric, non-diabetic nephropathy. The GISEN Group (Gruppo Italiano di Studi Epidemiologici in Nefrologia). Lancet (London, England) 1997; 349:1857–1863.
1363. Hou FF, Zhang X, Zhang GH, Xie D, Chen PY, Zhang WR, et al. Efficacy and safety of benazepril for advanced chronic renal insufficiency. The New England Journal of Medicine 2006; 354:131–140.
1364. Fried LF, Emanuele N, Zhang JH, Brophy M, Conner TA, Duckworth W, et al. Combined angiotensin inhibition for the treatment of diabetic nephropathy. The New England Journal of Medicine 2013; 369:1892–1903.
1365. Haller H, Ito S, Izzo JL, Januszewicz A, Katayama S, Menne J, et al. Olmesartan for the delay or prevention of microalbuminuria in type 2 diabetes. The New England Journal of Medicine 2011; 364:907–917.
1366. Rahman M, Pressel S, Davis BR, Nwachuku C, Wright JT, Whelton PK, et al. Renal outcomes in high-risk hypertensive patients treated with an angiotensin-converting enzyme inhibitor or a calcium channel blocker vs a diuretic: a report from the Antihypertensive and Lipid-Lowering Treatment to Prevent Heart Attack Trial (ALLHAT). Archives of Internal Medicine 2005; 165:936–946.
1367. Bhandari S, Mehta S, Khwaja A, Cleland JGF, Ives N, Brettell E, et al. Renin-Angiotensin System Inhibition in Advanced Chronic Kidney Disease. N Engl J Med 2022; 387:2021–2032.
1368. Sarafidis PA, Blacklock R, Wood E, Rumjon A, Simmonds S, Fletcher-Rogers J, et al. Prevalence and factors associated with hyperkalemia in predialysis patients followed in a low-clearance clinic. Clinical journal of the American Society of Nephrology: CJASN 2012; 7:1234–1241.
1369. Einhorn LM, Zhan M, Hsu VD, Walker LD, Moen MF, Seliger SL, et al. The frequency of hyperkalemia and its significance in chronic kidney disease. Archives of Internal Medicine 2009; 169:1156–1162.
1370. Epstein M, Reaven NL, Funk SE, McGaughey KJ, Oestreicher N, Knispel J. Evaluation of the treatment gap between clinical guidelines and the utilization of renin-angiotensin-aldosterone system inhibitors. Am J Manag Care 2015; 21:S212–220.
1371. Yildirim T, Arici M, Piskinpasa S, Aybal-Kutlugun A, Yilmaz R, Altun B, et al. Major barriers against renin-angiotensin-aldosterone system blocker use in chronic kidney disease stages 3-5 in clinical practice: a safety concern? Ren Fail 2012; 34:1095–1099.
1372. Fu EL, Evans M, Clase CM, Tomlinson LA, van Diepen M, Dekker FW, et al. Stopping Renin-Angiotensin System Inhibitors in Patients with Advanced CKD and Risk of Adverse Outcomes: A Nationwide Study. J Am Soc Nephrol 2021; 32:424–435.
1373. Walther CP, Winkelmayer WC, Richardson PA, Virani SS, Navaneethan SD. Renin-angiotensin system blocker discontinuation and adverse outcomes in chronic kidney disease. Nephrol Dial Transplant 2021; 36:1893–1899.
1374. Yang A, Shi M, Lau ESH, Wu H, Zhang X, Fan B, et al. Clinical outcomes following discontinuation of renin-angiotensin-system inhibitors in patients with type 2 diabetes and advanced chronic kidney disease: A prospective cohort study. EClinicalMedicine 2023; 55:101751.
1375. Bakris GL, Pitt B, Weir MR, Freeman MW, Mayo MR, Garza D, et al. Effect of Patiromer on Serum Potassium Level in Patients With Hyperkalemia and Diabetic Kidney Disease: The AMETHYST-DN Randomized Clinical Trial. JAMA 2015; 314:151–161.
1376. Roger SD, Spinowitz BS, Lerma EV, Singh B, Packham DK, Al-Shurbaji A, et al. Efficacy and Safety of Sodium Zirconium Cyclosilicate for Treatment of Hyperkalemia: An 11-Month Open-Label Extension of HARMONIZE. Am J Nephrol 2019; 50:473–480.
1377. Sarafidis PA, Georgianos PI, Bakris GL. Advances in treatment of hyperkalemia in chronic kidney disease. Expert opinion on pharmacotherapy 2015; 16:2205–2215.
1378. Valdivielso JM, Balafa O, Ekart R, Ferro CJ, Mallamaci F, Mark PB, et al. Hyperkalemia in Chronic Kidney Disease in the New Era of Kidney Protection Therapies. Drugs 2021; 81:1467–1489.
1379. Ott C, Schmieder RE. Diagnosis and treatment of arterial hypertension 2021. Kidney International 2022; 101:36–46.
1380. Sarafidis PA, Khosla N, Bakris GL. Antihypertensive therapy in the presence of proteinuria. American Journal of Kidney Diseases: The Official Journal of the National Kidney Foundation 2007; 49:12–26.
1381. Adler AI, Stratton IM, Neil HA, Yudkin JS, Matthews DR, Cull CA, et al. Association of systolic blood pressure with macrovascular and microvascular complications of type 2 diabetes (UKPDS 36): prospective observational study. BMJ (Clinical research ed) 2000; 321:412–419.
1382. Sarafidis PA, Georgianos PI, Lasaridis AN. Diuretics in clinical practice. Part I: mechanisms of action, pharmacological effects and clinical indications of diuretic compounds. Expert Opin Drug Saf 2010; 9:243–257.
1383. Bakris GL, Hart P, Ritz E. Beta blockers in the management of chronic kidney disease. Kidney International 2006; 70:1905–1913.
1384. Grassi G, Quarti-Trevano F, Seravalle G, Arenare F, Volpe M, Furiani S, et al. Early sympathetic activation in the initial clinical stages of chronic renal failure. Hypertension (Dallas, Tex: 1979) 2011; 57:846–851.
1385. Bakris GL, Copley JB, Vicknair N, Sadler R, Leurgans S. Calcium channel blockers versus other antihypertensive therapies on progression of NIDDM associated nephropathy. Kidney Int 1996; 50:1641–1650.
1386. Bakris GL, Mangrum A, Copley JB, Vicknair N, Sadler R. Effect of calcium channel or beta-blockade on the progression of diabetic nephropathy in African Americans. Hypertension (Dallas, Tex: 1979) 1997; 29:744–750.
1387. Ruggenenti P, Fassi A, Ilieva AP, Bruno S, Iliev IP, Brusegan V, et al. Preventing microalbuminuria in type 2 diabetes. The New England Journal of Medicine 2004; 351:1941–1951.
1388. Imprialos KP, Sarafidis PA, Karagiannis AI. Sodium-glucose cotransporter-2 inhibitors and blood pressure decrease: a valuable effect of a novel antidiabetic class? Journal of Hypertension 2015; 33:2185–2197.
1389. Papadopoulou E, Loutradis C, Tzatzagou G, Kotsa K, Zografou I, Minopoulou I, et al. Dapagliflozin decreases ambulatory central blood pressure and pulse wave velocity in patients with type 2 diabetes: a randomized, double-blind, placebo-controlled clinical trial. Journal of Hypertension 2021; 39:749–758.
1390. Cherney DZI, Cooper ME, Tikkanen I, Pfarr E, Johansen OE, Woerle HJ, et al. Pooled analysis of Phase III trials indicate contrasting influences of renal function on blood pressure, body weight, and HbA1c reductions with empagliflozin. Kidney International 2018; 93:231–244.
1391. Mancia G, Cannon CP, Tikkanen I, Zeller C, Ley L, Woerle HJ, et al. Impact of empagliflozin on blood pressure in patients with type 2 diabetes mellitus and hypertension by background antihypertensive medication. Hypertension (Dallas, Tex: 1979) 2016; 68:1355–1364.
1392. Piperidou A, Sarafidis P, Boutou A, Thomopoulos C, Loutradis C, Alexandrou ME, et al. The effect of SGLT-2 inhibitors on albuminuria and proteinuria in diabetes mellitus: a systematic review and meta-analysis of randomized controlled trials. Journal of Hypertension 2019; 37:1334–1343.
1393. Wiviott SD, Raz I, Bonaca MP, Mosenzon O, Kato ET, Cahn A, et al. Dapagliflozin and cardiovascular outcomes in type 2 diabetes. The New England Journal of Medicine 2019; 380:347–357.
1394. Sarafidis P, Ortiz A, Ferro CJ, Halimi J-M, Kreutz R, Mallamaci F, et al. Sodium--glucose co-transporter-2 inhibitors for patients with diabetic and nondiabetic chronic kidney disease: a new era has already begun. Journal of Hypertension 2021; 39:1090–1097.
1395. Heerspink HJL, Stefánsson BV, Correa-Rotter R, Chertow GM, Greene T, Hou F-F, et al. Dapagliflozin in patients with chronic kidney disease. The New England Journal of Medicine 2020; 383:1436–1446.
1396. Cherney DZI, Perkins BA, Soleymanlou N, Maione M, Lai V, Lee A, et al. Renal hemodynamic effect of sodium-glucose cotransporter 2 inhibition in patients with type 1 diabetes mellitus. Circulation 2014; 129:587–597.
1397. Sarafidis P, Ferro CJ, Morales E, Ortiz A, Malyszko J, Hojs R, et al. SGLT-2 inhibitors and GLP-1 receptor agonists for nephroprotection and cardioprotection in patients with diabetes mellitus and chronic kidney disease. A consensus statement by the EURECA-m and the DIABESITY working groups of the ERA-EDTA. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2019; 34:208–230.
1398. Sarafidis P, Papadopoulos CE, Kamperidis V, Giannakoulas G, Doumas M. Cardiovascular protection with sodium-glucose cotransporter-2 inhibitors and mineralocorticoid receptor antagonists in chronic kidney disease: a milestone achieved. Hypertension (Dallas, Tex: 1979) 2021; 77:1442–1455.
1399. Lin Y-C, Lin J-W, Wu M-S, Chen K-C, Peng C-C, Kang Y-N. Effects of calcium channel blockers comparing to angiotensin-converting enzyme inhibitors and angiotensin receptor blockers in patients with hypertension and chronic kidney disease stage 3 to 5 and dialysis: A systematic review and meta-analysis. PloS One 2017; 12:e0188975.
1400. Sarafidis PA, Stafylas PC, Kanaki AI, Lasaridis AN. Effects of renin-angiotensin system blockers on renal outcomes and all-cause mortality in patients with diabetic nephropathy: an updated meta-analysis. American Journal of Hypertension 2008; 21:922–929.
1401. Sharma P, Blackburn RC, Parke CL, McCullough K, Marks A, Black C. Angiotensin-converting enzyme inhibitors and angiotensin receptor blockers for adults with early (stage 1 to 3) non-diabetic chronic kidney disease. The Cochrane Database of Systematic Reviews 2011; CD007751.
1402. Alexandrou M-E, Papagianni A, Tsapas A, Loutradis C, Boutou A, Piperidou A, et al. Effects of mineralocorticoid receptor antagonists in proteinuric kidney disease: a systematic review and meta-analysis of randomized controlled trials. Journal of Hypertension 2019; 37:2307–2324.
1403. Epstein M, Williams GH, Weinberger M, Lewin A, Krause S, Mukherjee R, et al. Selective aldosterone blockade with eplerenone reduces albuminuria in patients with type 2 diabetes. Clinical journal of the American Society of Nephrology: CJASN 2006; 1:940–951.
1404. Mehdi UF, Adams-Huet B, Raskin P, Vega GL, Toto RD. Addition of angiotensin receptor blockade or mineralocorticoid antagonism to maximal angiotensin-converting enzyme inhibition in diabetic nephropathy. Journal of the American Society of Nephrology: JASN 2009; 20:2641–2650.
1405. Alexandrou M-E, Ferro CJ, Boletis I, Papagianni A, Sarafidis P. Hypertension in kidney transplant recipients. World J Transplant 2022; 12:211–222.
1406. Ortiz A, Ferro CJ, Balafa O, Burnier M, Ekart R, Halimi J-M, et al. Mineralocorticoid receptor antagonists for nephroprotection and cardioprotection in patients with diabetes mellitus and chronic kidney disease. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2021:gfab167.
1407. Bakris GL, Agarwal R, Chan JC, Cooper ME, Gansevoort RT, Haller H, et al. Effect of finerenone on albuminuria in patients with diabetic nephropathy: a randomized clinical trial. JAMA 2015; 314:884–894.
1408. Ruilope LM, Agarwal R, Anker SD, Bakris GL, Filippatos G, Nowack C, et al. Design and baseline characteristics of the finerenone in reducing cardiovascular mortality and morbidity in diabetic kidney disease trial. Am J Nephrol 2019; 50:345–356.
1409. Filippatos G, Anker SD, August P, Coats AJS, Januzzi JL, Mankovsky B, et al. Finerenone and effects on mortality in chronic kidney disease and type 2 diabetes: a FIDELITY analysis. Eur Heart J Cardiovasc Pharmacother 2023; 9:183–191.
1410. Alexandrou M-E, Theodorakopoulou MP, Sarafidis PA. Role of Mineralocorticoid Receptor Antagonists in Diabetic Kidney Disease. Kidney and Dialysis 2022; 2:163–182.
1411. Anderson GH, Blakeman N, Streeten DH. The effect of age on prevalence of secondary forms of hypertension in 4429 consecutively referred patients. Journal of Hypertension 1994; 12:609–615.
1412. Elliott WJ. Renovascular hypertension: an update. Journal of Clinical Hypertension (Greenwich Conn) 2008; 10:522–533.
1413. Safian RD, Textor SC. Renal-artery stenosis. The New England Journal of Medicine 2001; 344:431–442.
1414. Van der Niepen P, Rossignol P, Lengelé J-P, Berra E, Sarafidis P, Persu A. Renal Artery Stenosis in Patients with Resistant Hypertension: Stent It or Not? Current Hypertension Reports 2017; 19:5.
1415. Calhoun DA, Jones D, Textor S, Goff DC, Murphy TP, Toto RD, et al. Resistant hypertension: diagnosis, evaluation, and treatment: a scientific statement from the American Heart Association Professional Education Committee of the Council for High Blood Pressure Research. Circulation 2008; 117:e510–e526.
1416. Kalra PA, Guo H, Kausz AT, Gilbertson DT, Liu J, Chen S-C, et al. Atherosclerotic renovascular disease in United States patients aged 67 years or older: risk factors, revascularization, and prognosis. Kidney International 2005; 68:293–301.
1417. de Mast Q, Beutler JJ. The prevalence of atherosclerotic renal artery stenosis in risk groups: a systematic literature review. Journal of Hypertension 2009; 27:1333–1340.
1418. Hicks CW, Clark TWI, Cooper CJ, de Bhailís ÁM, De Carlo M, Green D, et al. Atherosclerotic Renovascular Disease: A KDIGO (Kidney Disease: Improving Global Outcomes) Controversies Conference. American Journal of Kidney Diseases: The Official Journal of the National Kidney Foundation 2022; 79:289–301.
1419. Persu A, Giavarini A, Touzé E, Januszewicz A, Sapoval M, Azizi M, et al. European consensus on the diagnosis and management of fibromuscular dysplasia. Journal of Hypertension 2014; 32:1367–1378.
1420. Ritchie J, Green D, Alderson HV, Chiu D, Sinha S, Kalra PA. Risks for mortality and renal replacement therapy in atherosclerotic renovascular disease compared with other causes of chronic kidney disease. Nephrology (Carlton) 2015; 20:688–696.
1421. Ritchie J, Green D, Chrysochou C, Chalmers N, Foley RN, Kalra PA. High-risk clinical presentations in atherosclerotic renovascular disease: prognosis and response to renal artery revascularization. American Journal of Kidney Diseases: The Official Journal of the National Kidney Foundation 2014; 63:186–197.
1422. Pappaccogli M, Robberechts T, Lengelé JP, Van der Niepen P, Sarafidis P, Rabbia F, et al. Endovascular versus medical management of atherosclerotic renovascular disease: update and emerging concepts. Hypertension (Dallas, Tex: 1979) 2023; 80:1150–1161.
1423. Theodorakopoulou MP, Karagiannidis AG, Ferro CJ, Ortiz A, Sarafidis PA. Renal artery stenting in the correct patients with atherosclerotic renovascular disease: time for a proper renal and cardiovascular outcome study? Clinical Kidney Journal 2022; sfac140.
1424. Bax L, Woittiez A-JJ, Kouwenberg HJ, Mali WPTM, Buskens E, Beek FJA, et al. Stent placement in patients with atherosclerotic renal artery stenosis and impaired renal function: a randomized trial. Annals of Internal Medicine 2009; 150:840–848. W150-151.
1425. Cooper CJ, Murphy TP, Cutlip DE, Jamerson K, Henrich W, Reid DM, et al. Stenting and medical therapy for atherosclerotic renal-artery stenosis. The New England Journal of Medicine 2014; 370:13–22.
1426. Investigators A, Wheatley K, Ives N, Gray R, Kalra PA, Moss JG, et al. Revascularization versus medical therapy for renal-artery stenosis. The New England Journal of Medicine 2009; 361:1953–1962.
1427. Reinhard M, Schousboe K, Andersen UB, Buus NH, Rantanen JM, Bech JN, et al. Renal Artery Stenting in Consecutive High-Risk Patients With Atherosclerotic Renovascular Disease: A Prospective 2-Center Cohort Study. Journal of the American Heart Association 2022; 11:e024421.
1428. Rangaswami J, Mathew RO, Parasuraman R, Tantisattamo E, Lubetzky M, Rao S, et al. Cardiovascular disease in the kidney transplant recipient: epidemiology, diagnosis and management strategies. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2019; 34:760–773.
1429. Ying T, Shi B, Kelly PJ, Pilmore H, Clayton PA, Chadban SJ. Death after Kidney Transplantation: An Analysis by Era and Time Post-Transplant. Journal of the American Society of Nephrology: JASN 2020; 31:2887–2899.
1430. Buus NH, Carlsen RK, Hughes AD, Skov K. Influence of Renal Transplantation and Living Kidney Donation on Large Artery Stiffness and Peripheral Vascular Resistance. American Journal of Hypertension 2020; 33:234–242.
1431. Xagas E, Sarafidis PA, Theodorakopoulou MP, Alexandrou ME, Korogiannou M, Argyris A, et al. A parallel evaluation of short- and mid-term changes of ambulatory blood pressure in kidney transplant recipients and kidney donors. Clinical Kidney Journal 2022; sfac162.
1432. Korogiannou M, Sarafidis P, Alexandrou ME, Theodorakopoulou MP, Pella E, Xagas E, et al. Ambulatory blood pressure trajectories and blood pressure variability in kidney transplant recipients: a comparative study against haemodialysis patients. Clinical Kidney Journal 2022; 15:951–960.
1433. Korogiannou M, Sarafidis P, Theodorakopoulou MP, Alexandrou ME, Xagas E, Boletis IN, et al. Diagnostic Performance of Office versus Ambulatory Blood Pressure in Kidney Transplant Recipients. Am J Nephrol 2021; 52:548–558.
1434. Opelz G, Döhler B, Collaborative Transplant S. Improved long-term outcomes after renal transplantation associated with blood pressure control. Am J Transplant 2005; 5:2725–2731.
1435. Pisano A, Mallamaci F, D’Arrigo G, Bolignano D, Wuerzner G, Ortiz A, et al. Blood pressure monitoring in kidney transplantation: a systematic review on hypertension and target organ damage. Nephrol Dial Transplant 2021; gfab076.
1436. Halimi J-M, Ortiz A, Sarafidis PA, Mallamaci F, Wuerzner G, Pisano A, et al. Hypertension in kidney transplantation: a consensus statement of the ’hypertension and the kidney’ working group of the European Society of Hypertension. Journal of Hypertension 2021; 39:1513–1521.
1437. Pisano A, Mallamaci F, D’Arrigo G, Bolignano D, Wuerzner G, Ortiz A, et al. Assessment of hypertension in kidney transplantation by ambulatory blood pressure monitoring: a systematic review and meta-analysis. Clinical Kidney Journal 2022; 15:31–42.
1438. Kooman JP, Christiaans MH, Boots JM, van Der Sande FM, Leunissen KM, van Hooff JP. A comparison between office and ambulatory blood pressure measurements in renal transplant patients with chronic transplant nephropathy. American Journal of Kidney Diseases: The Official Journal of the National Kidney Foundation 2001; 37:1170–1176.
1439. Mallamaci F, Tripepi R, D’Arrigo G, Porto G, Versace MC, Marino C, et al. Long-term blood pressure monitoring by office and 24-h ambulatory blood pressure in renal transplant patients: a longitudinal study. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2019; 34:1558–1564.
1440. Loutradis C, Sarafidis P, Marinaki S, Berry M, Borrows R, Sharif A, et al. Role of hypertension in kidney transplant recipients. Journal of Human Hypertension 2021; 35:958–969.
1441. Chatzikyrkou C, Schmieder RE, Schiffer M. Update on Treatment of Hypertension After Renal Transplantation. Current Hypertension Reports 2021; 23:25.
1442. Knight SR, Morris PJ. Steroid avoidance or withdrawal after renal transplantation increases the risk of acute rejection but decreases cardiovascular risk. A meta-analysis. Transplantation 2010; 89:1–14.
1443. Rice JB, White AG, Scarpati LM, Wan G, Nelson WW. Long-term systemic corticosteroid exposure: a systematic literature review. Clin Ther 2017; 39:2216–2229.
1444. Hausberg M, Kosch M, Harmelink P, Barenbrock M, Hohage H, Kisters K, et al. Sympathetic nerve activity in end-stage renal disease. Circulation 2002; 106:1974–1979.
1445. Pisano A, Bolignano D, Mallamaci F, D’Arrigo G, Halimi J-M, Persu A, et al. Comparative effectiveness of different antihypertensive agents in kidney transplantation: a systematic review and meta-analysis. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association 2020; 35:878–887.
1446. Ngo AT, Markar SR, De Lijster MS, Duncan N, Taube D, Hamady MS. A Systematic Review of Outcomes Following Percutaneous Transluminal Angioplasty and Stenting in the Treatment of Transplant Renal Artery Stenosis. Cardiovasc Intervent Radiol 2015; 38:1573–1588.
1447. Jordan J, Yumuk V, Schlaich M, Nilsson PM, Zahorska-Markiewicz B, Grassi G, et al. Joint statement of the European Association for the Study of Obesity and the European Society of Hypertension: obesity and difficult to treat arterial hypertension. J Hypertens 2012; 30:1047–1055.
1448. Bramlage P, Pittrow D, Wittchen HU, Kirch W, Boehler S, Lehnert H, et al. Hypertension in overweight and obese primary care patients is highly prevalent and poorly controlled. Am J Hypertens 2004; 17:904–910.
1449. Cushman WC, Ford CE, Cutler JA, Margolis KL, Davis BR, Grimm RH, et al. Success and predictors of blood pressure control in diverse North American settings: the antihypertensive and lipid-lowering treatment to prevent heart attack trial (ALLHAT). J Clin Hypertens (Greenwich) 2002; 4:393–404.
1450. Reisin E, Graves J, Yamal JM, Barzilay JI, Pressel S, Einhorn PT, et al. Blood pressure control and cardiovascular outcomes in normal, overweight, and obese hypertensives treated with three different anti-hypertensives in ALLHAT. J Hypertens 2014; 32:1503–1513.
1451. McMurray JJ, Holman RR, Haffner SM, Bethel MA, Holzhauer B, Hua TA, et al. Effect of valsartan on the incidence of diabetes and cardiovascular events. N Engl J Med 2010; 362:1477–1490.
1452. Bosch J, Yusuf S, Gerstein HC, Pogue J, Sheridan P, Dagenais G, et al. Effect of ramipril on the incidence of diabetes. N Engl J Med 2006; 355:1551–1562.
1453. Ginsberg H, Kimmerling G, Olefsky JM, Reaven GM. Demonstration of insulin resistance in untreated adult onset diabetic subjects with fasting hyperglycemia. J Clin Invest 1975; 55:454–461.
1454. Sharma AM, Pischon T, Hardt S, Kunz I, Luft FC. Hypothesis: Beta-adrenergic receptor blockers and weight gain: A systematic analysis. Hypertension (Dallas, Tex: 1979) 2001; 37:250–254.
1455. Bakris GL, Fonseca V, Katholi RE, McGill JB, Messerli FH, Phillips RA, et al. Metabolic effects of carvedilol vs metoprolol in patients with type 2 diabetes mellitus and hypertension: a randomized controlled trial. JAMA 2004; 292:2227–2236.
1456. Agabiti Rosei E, Rizzoni D. Metabolic profile of nebivolol, a beta-adrenoceptor antagonist with unique characteristics. Drugs 2007; 67:1097–1107.
1457. Messerli FH, Christie B, DeCarvalho JG, Aristimuno GG, Suarez DH, Dreslinski GR, et al. Obesity and essential hypertension. Hemodynamics, intravascular volume, sodium excretion, and plasma renin activity. Arch Intern Med 1981; 141:81–85.
1458. Rocchini AP, Key J, Bondie D, Chico R, Moorehead C, Katch V, et al. The effect of weight loss on the sensitivity of blood pressure to sodium in obese adolescents. N Engl J Med 1989; 321:580–585.
1459. Grassi G, Seravalle G, Dell’Oro R, Turri C, Bolla GB, Mancia G. Adrenergic and reflex abnormalities in obesity-related hypertension. Hypertension (Dallas, Tex: 1979) 2000; 36:538–542.
1460. Rumantir MS, Vaz M, Jennings GL, Collier G, Kaye DM, Seals DR, et al. Neural mechanisms in human obesity-related hypertension. J Hypertens 1999; 17:1125–1133.
1461. Weber MA, Jamerson K, Bakris GL, Weir MR, Zappe D, Zhang Y, et al. Effects of body size and hypertension treatments on cardiovascular event rates: subanalysis of the ACCOMPLISH randomised controlled trial. Lancet 2013; 381:537–545.
1462. Jordan J, Engeli S, Boye SW, Le Breton S, Keefe DL. Direct Renin inhibition with aliskiren in obese patients with arterial hypertension. Hypertension (Dallas, Tex: 1979) 2007; 49:1047–1055.
1463. Jordan J, Stinkens R, Jax T, Engeli S, Blaak EE, May M, et al. Improved Insulin Sensitivity With Angiotensin Receptor Neprilysin Inhibition in Individuals With Obesity and Hypertension. Clin Pharmacol Ther 2017; 101:254–263.
1464. Zomer E, Gurusamy K, Leach R, Trimmer C, Lobstein T, Morris S, et al. Interventions that cause weight loss and the impact on cardiovascular risk factors: a systematic review and meta-analysis. Obes Rev 2016; 17:1001–1011.
1465. Stamler R, Stamler J, Gosch FC, Civinelli J, Fishman J, McKeever P, et al. Primary prevention of hypertension by nutritional-hygienic means. Final report of a randomized, controlled trial. JAMA 1989; 262:1801–1807.
1466. The effects of nonpharmacologic interventions on blood pressure of persons with high normal levels. Results of the Trials of Hypertension Prevention, Phase I. JAMA 1992; 267:1213–1220.
1467. Stevens VJ, Obarzanek E, Cook NR, Lee IM, Appel LJ, Smith-West D, et al. Long-term weight loss and changes in blood pressure: results of the Trials of Hypertension Prevention, phase II. Ann Intern Med 2001; 134:1–11.
1468. Langford HG, Davis BR, Blaufox D, Oberman A, Wassertheil-Smoller S, Hawkins M, et al. Effect of drug and diet treatment of mild hypertension on diastolic blood pressure: The TAIM Research Group. Hypertension (Dallas, Tex: 1979) 1991; 17:210–217.
1469. Whelton PK, Appel LJ, Espeland MA, Applegate WB, Ettinger WH Jr, Kostis JB, et al. Sodium reduction and weight loss in the treatment of hypertension in older persons: a randomized controlled trial of nonpharmacologic interventions in the elderly (TONE). TONE Collaborative Research Group. JAMA 1998; 279:839–846.
1470. Semlitsch T, Krenn C, Jeitler K, Berghold A, Horvath K, Siebenhofer A. Long-term effects of weight-reducing diets in people with hypertension. Cochrane Database Syst Rev 2021; 2:Cd008274.
1471. Haufe S, Kaminski J, Utz W, Haas V, Mahler A, Daniels MA, et al. Differential response of the natriuretic peptide system to weight loss and exercise in overweight or obese patients. J Hypertens 2015; 33:1458–1464.
1472. Henke C, Haufe S, Ziehl D, Bornstein SR, Schulz-Menger J, Heni M, et al. Low-fat hypocaloric diet reduces neprilysin in overweight and obese human subjects. ESC Heart Fail 2021; 8:938–942.
1473. Bakris G, Calhoun D, Egan B, Hellmann C, Dolker M, Kingma I. Orlistat improves blood pressure control in obese subjects with treated but inadequately controlled hypertension. J Hypertens 2002; 20:2257–2267.
1474. Siebenhofer A, Jeitler K, Horvath K, Berghold A, Posch N, Meschik J, et al. Long-term effects of weight-reducing drugs in people with hypertension. Cochrane Database Syst Rev 2016; 3:Cd007654.
1475. Jordan J, Astrup A, Engeli S, Narkiewicz K, Day WW, Finer N. Cardiovascular effects of phentermine and topiramate: a new drug combination for the treatment of obesity. J Hypertens 2014; 32:1178–1188.
1476. Apovian CM, Aronne L, Rubino D, Still C, Wyatt H, Burns C, et al. A randomized, phase 3 trial of naltrexone SR/bupropion SR on weight and obesity-related risk factors (COR-II). Obesity (Silver Spring, Md) 2013; 21:935–943.
1477. Pi-Sunyer X, Astrup A, Fujioka K, Greenway F, Halpern A, Krempf M, et al. A randomized, controlled trial of 3.0 mg of liraglutide in weight management. N Engl J Med 2015; 373:11–22.
1478. Smith SR, Weissman NJ, Anderson CM, Sanchez M, Chuang E, Stubbe S, et al. Multicenter, placebo-controlled trial of lorcaserin for weight management. N Engl J Med 2010; 363:245–256.
1479. Rubino DM, Greenway FL, Khalid U, O’Neil PM, Rosenstock J, Sørrig R, et al. Effect of weekly subcutaneous semaglutide vs daily liraglutide on body weight in adults with overweight or obesity without diabetes: the STEP 8 randomized clinical trial. JAMA 2022; 327:138–150.
1480. Maloney A, Rosenstock J, Fonseca V. A model-based meta-analysis of 24 antihyperglycemic drugs for type 2 diabetes: comparison of treatment effects at therapeutic doses. Clin Pharmacol Ther 2019; 105:1213–1223.
1481. Knowler WC, Barrett-Connor E, Fowler SE, Hamman RF, Lachin JM, Walker EA, et al. Reduction in the incidence of type 2 diabetes with lifestyle intervention or metformin. N Engl J Med 2002; 346:393–403.
1482. Sjostrom L, Peltonen M, Jacobson P, Sjostrom CD, Karason K, Wedel H, et al. Bariatric surgery and long-term cardiovascular events. JAMA 2012; 307:56–65.
1483. Wiggins T, Guidozzi N, Welbourn R, Ahmed AR, Markar SR. Association of bariatric surgery with all-cause mortality and incidence of obesity-related disease at a population level: A systematic review and meta-analysis. PLoS Med 2020; 17:e1003206.
1484. Yoshino M, Kayser BD, Yoshino J, Stein RI, Reeds D, Eagon JC, et al. Effects of Diet versus Gastric Bypass on Metabolic Function in Diabetes. N Engl J Med 2020; 383:721–732.
1485. Adams TD, Davidson LE, Litwin SE, Kim J, Kolotkin RL, Nanjee MN, et al. Weight and Metabolic Outcomes 12 Years after Gastric Bypass. N Engl J Med 2017; 377:1143–1155.
1486. Seravalle G, Colombo M, Perego P, Giardini V, Volpe M, Dell’Oro R, et al. Long-term sympathoinhibitory effects of surgically induced weight loss in severe obese patients. Hypertension (Dallas, Tex: 1979) 2014; 64:431–437.
1487. Bonfils PK, Taskiran M, Damgaard M, Goetze JP, Floyd AK, Funch-Jensen P, et al. Roux-en-Y gastric bypass alleviates hypertension and is associated with an increase in mid-regional pro-atrial natriuretic peptide in morbid obese patients. J Hypertens 2015; 33:1215–1225.
1488. Gabrielsen AM, Omland T, Brokner M, Fredheim JM, Jordan J, Lehmann S, et al. The effect of surgical and non-surgical weight loss on N-terminal pro-B-type natriuretic peptide and its relation to obstructive sleep apnea and pulmonary function. BMC research notes 2016; 9:440.
1489. Sjöström L, Lindroos AK, Peltonen M, Torgerson J, Bouchard C, Carlsson B, et al. Lifestyle, diabetes, and cardiovascular risk factors 10 years after bariatric surgery. N Engl J Med 2004; 351:2683–2693.
1490. Schiavon CA, Bhatt DL, Ikeoka D, Santucci EV, Santos RN, Damiani LP, et al. Three-Year Outcomes of Bariatric Surgery in Patients With Obesity and Hypertension: A Randomized Clinical Trial. Ann Intern Med 2020; 173:685–693.
1491. Cappuccio FP, Cooper D, D’Elia L, Strazzullo P, Miller MA. Sleep duration predicts cardiovascular outcomes: a systematic review and meta-analysis of prospective studies. Eur Heart J 2011; 32:1484–1492.
1492. Pedrosa RP, Drager LF, de Paula LKG, Amaro ACS, Bortolotto LA, Lorenzi-Filho G. Effects of OSA treatment on BP in patients with resistant hypertension: a randomized trial. Chest 2013; 144:1487–1494.
1493. Feldstein CA. Blood pressure effects of CPAP in nonresistant and resistant hypertension associated with OSA: A systematic review of randomized clinical trials. Clin Exp Hypertens 2016; 38:337–346.
1494. Andren A, Hedberg P, Walker-Engstrom ML, Wahlen P, Tegelberg A. Effects of treatment with oral appliance on 24-h blood pressure in patients with obstructive sleep apnea and hypertension: a randomized clinical trial. Sleep Breath 2013; 17:705–712.
1495. Kou C, Zhao X, Lin X, Fan X, Wang Q, Yu J. Effect of different treatments for obstructive sleep apnoea on blood pressure. J Hypertens 2022; 40:1071–1084.
1496. Pengo MF, Soranna D, Giontella A, Perger E, Mattaliano P, Schwarz EI, et al. Obstructive sleep apnoea treatment and blood pressure: which phenotypes predict a response? A systematic review and meta-analysis. Eur Respir J 2020; 55.
1497. Kovacs DK, Gede N, Szabo L, Hegyi P, Szakacs Z, Faludi B, et al. Weight reduction added to CPAP decreases blood pressure and triglyceride level in OSA: Systematic review and meta-analysis. Clin Transl Sci 2022; 15:1238–1248.
1498. Cao MT, Sternbach JM, Guilleminault C. Continuous positive airway pressure therapy in obstuctive sleep apnea: benefits and alternatives. Expert Rev Respir Med 2017; 11:259–272.
1499. Chalegre ST, Lins-Filho OL, Lustosa TC, Franca MV, Couto TLG, Drager LF, et al. Impact of CPAP on arterial stiffness in patients with obstructive sleep apnea: a meta-analysis of randomized trials. Sleep Breath 2021; 25:1195–1202.
1500. Green M, Ken-Dror G, Fluck D, Sada C, Sharma P, Fry CH, et al. Meta-analysis of changes in the levels of catecholamines and blood pressure with continuous positive airway pressure therapy in obstructive sleep apnea. J Clin Hypertens (Greenwich) 2021; 23:12–20.
1501. Christiansen SC, Zuraw BL. Treatment of Hypertension in Patients with Asthma. N Engl J Med 2019; 381:1046–1057.
1502. Sin DD, Wu L, Man SF. The relationship between reduced lung function and cardiovascular mortality: a population-based study and a systematic review of the literature. Chest 2005; 127:1952–1959.
1503. Ferguson S, Teodorescu MC, Gangnon RE, Peterson AG, Consens FB, Chervin RD, et al. Factors associated with systemic hypertension in asthma. Lung 2014; 192:675–683.
1504. Peters U, Dixon AE, Forno E. Obesity and asthma. J Allergy Clin Immunol 2018; 141:1169–1179.
1505. Ryan S, Taylor CT, McNicholas WT. Systemic inflammation: a key factor in the pathogenesis of cardiovascular complications in obstructive sleep apnoea syndrome? Thorax 2009; 64:631–636.
1506. Chen E, Miller GE. Stress and inflammation in exacerbations of asthma. Brain Behav Immun 2007; 21:993–999.
1507. Spruill TM. Chronic psychosocial stress and hypertension. Curr Hypertens Rep 2010; 12:10–16.
1508. Global initiative for asthma. Global strategy for asthma management and prevention (http://www.ginasthma.org).
1509. Brooks TW, Creekmore FM, Young DC, Asche CV, Oberg B, Samuelson WM. Rates of hospitalizations and emergency department visits in patients with asthma and chronic obstructive pulmonary disease taking beta-blockers. Pharmacotherapy 2007; 27:684–690.
1510. Bennett M, Chang CL, Tatley M, Savage R, Hancox RJ. The safety of cardioselective β(1)-blockers in asthma: literature review and search of global pharmacovigilance safety reports. ERJ Open Res 2021; 7:
1511. Finks SW, Rumbak MJ, Self TH. Treating Hypertension in Chronic Obstructive Pulmonary Disease. N Engl J Med 2020; 382:353–363.
1512. Thomsen M, Dahl M, Lange P, Vestbo J, Nordestgaard BG. Inflammatory biomarkers and comorbidities in chronic obstructive pulmonary disease. Am J Respir Crit Care Med 2012; 186:982–988.
1513. Miller J, Edwards LD, Agusti A, Bakke P, Calverley PM, Celli B, et al. Comorbidity, systemic inflammation and outcomes in the ECLIPSE cohort. Respir Med 2013; 107:1376–1384.
1514. Yang YL, Xiang ZJ, Yang JH, Wang WJ, Xu ZC, Xiang RL. Association of β-blocker use with survival and pulmonary function in patients with chronic obstructive pulmonary and cardiovascular disease: a systematic review and meta-analysis. Eur Heart J 2020; 41:4415–4422.
1515. Hansildaar R, Vedder D, Baniaamam M, Tausche AK, Gerritsen M, Nurmohamed MT. Cardiovascular risk in inflammatory arthritis: rheumatoid arthritis and gout. Lancet Rheumatol 2021; 3:e58–e70.
1516. Krishnan E, Kwoh CK, Schumacher HR, Kuller L. Hyperuricemia and incidence of hypertension among men without metabolic syndrome. Hypertension (Dallas, Tex: 1979) 2007; 49:298–303.
1517. Bombelli M, Ronchi I, Volpe M, Facchetti R, Carugo S, Dell’oro R, et al. Prognostic value of serum uric acid: new-onset in and out-of-office hypertension and long-term mortality. J Hypertens 2014; 32:1237–1244.
1518. Cicero AF, Salvi P, D’Addato S, Rosticci M, Borghi C. Brisighella Heart Study g. Association between serum uric acid, hypertension, vascular stiffness and subclinical atherosclerosis: data from the Brisighella Heart Study. J Hypertens 2014; 32:57–64.
1519. Borghi C, Agnoletti D, Cicero AFG, Lurbe E, Virdis A. Uric acid and hypertension: a review of evidence and future perspectives for the management of cardiovascular risk. Hypertension (Dallas, Tex: 1979) 2022; 79:1927–1936.
1520. Bhole V, Krishnan E. Gout and the heart. Rheum Dis Clin North Am 2014; 40:125–143.
1521. Ndrepepa G. Uric acid and cardiovascular disease. Clin Chim Acta 2018; 484:150–163.
1522. Agarwal V, Hans N, Messerli FH. Effect of allopurinol on blood pressure: a systematic review and meta-analysis. J Clin Hypertens (Greenwich) 2013; 15:435–442.
1523. Chen JH, Lan JL, Cheng CF, Liang WM, Lin HY, Tsay GJ, et al. Effect of Urate-lowering Therapy on the Risk of Cardiovascular Disease and All-cause Mortality in Patients with Gout: A Case-matched Cohort Study. J Rheumatol 2015; 42:1694–1701.
1524. Gill D, Cameron AC, Burgess S, Li X, Doherty DJ, Karhunen V, et al. Urate, Blood Pressure, and Cardiovascular Disease: Evidence From Mendelian Randomization and Meta-Analysis of Clinical Trials. Hypertension (Dallas, Tex: 1979) 2021; 77:383–392.
1525. Drosos GC, Vedder D, Houben E, Boekel L, Atzeni F, Badreh S, et al. EULAR recommendations for cardiovascular risk management in rheumatic and musculoskeletal diseases, including systemic lupus erythematosus and antiphospholipid syndrome. Ann Rheum Dis 2022; 81:768–779.
1526. Cebollada J, Buisan C. Therapeutic role of colchicine in reducing cardiovascular risk associated with inflammation. Endocrinol Diabetes Nutr (Engl Ed) 2022; 69:551–553.
1527. White WB, Saag KG, Becker MA, Borer JS, Gorelick PB, Whelton A, et al. Cardiovascular Safety of Febuxostat or Allopurinol in Patients with Gout. N Engl J Med 2018; 378:1200–1210.
1528. Burnier M. Gout and hyperuricaemia: modifiable cardiovascular risk factors? Front Cardiovasc Med 2023; 10:
1529. Borghi C, Agabiti-Rosei E, Johnson RJ, Kielstein JT, Lurbe E, Mancia G, et al. Hyperuricaemia and gout in cardiovascular, metabolic and kidney disease. Eur J Intern Med 2020; 80:1–11.
1530. Choi HK, Soriano LC, Zhang Y, Rodriguez LA. Antihypertensive drugs and risk of incident gout among patients with hypertension: population based case-control study. BMJ 2012; 344:d8190.
1531. Hoieggen A, Alderman MH, Kjeldsen SE, Julius S, Devereux RB, De Faire U, et al. The impact of serum uric acid on cardiovascular outcomes in the LIFE study. Kidney Int 2004; 65:1041–1049.
1532. Ikdahl E, Wibetoe G, Rollefstad S, Salberg A, Bergsmark K, Kvien TK, et al. Guideline recommended treatment to targets of cardiovascular risk is inadequate in patients with inflammatory joint diseases. Int J Cardiol 2019; 274:311–318.
1533. Jagpal A, Navarro-Millan I. Cardiovascular co-morbidity in patients with rheumatoid arthritis: a narrative review of risk factors, cardiovascular risk assessment and treatment. BMC Rheumatol 2018; 2:10.
1534. Panoulas VF, Metsios GS, Pace AV, John H, Treharne GJ, Banks MJ, et al. Hypertension in rheumatoid arthritis. Rheumatology (Oxford) 2008; 47:1286–1298.
1535. Dijkshoorn B, Raadsen R, Nurmohamed MT. Cardiovascular Disease Risk in Rheumatoid Arthritis Anno 2022. J Clin Med 2022; 11:
1536. England BR, Thiele GM, Anderson DR, Mikuls TR. Increased cardiovascular risk in rheumatoid arthritis: mechanisms and implications. BMJ 2018; 361:k1036.
1537. Boyer JF, Gourraud PA, Cantagrel A, Davignon JL, Constantin A. Traditional cardiovascular risk factors in rheumatoid arthritis: a meta-analysis. Joint Bone Spine 2011; 78:179–183.
1538. McEntegart A, Capell HA, Creran D, Rumley A, Woodward M, Lowe GD. Cardiovascular risk factors, including thrombotic variables, in a population with rheumatoid arthritis. Rheumatology (Oxford) 2001; 40:640–644.
1539. Gonzalez A, Maradit Kremers H, Crowson CS, Ballman KV, Roger VL, Jacobsen SJ, et al. Do cardiovascular risk factors confer the same risk for cardiovascular outcomes in rheumatoid arthritis patients as in non-rheumatoid arthritis patients? Ann Rheum Dis 2008; 67:64–69.
1540. Kerola AM, Kauppi MJ, Kerola T, Nieminen TV. How early in the course of rheumatoid arthritis does the excess cardiovascular risk appear? Ann Rheum Dis 2012; 71:1606–1615.
1541. Kremer JM, Bloom BJ, Breedveld FC, Coombs JH, Fletcher MP, Gruben D, et al. The safety and efficacy of a JAK inhibitor in patients with active rheumatoid arthritis: Results of a double-blind, placebo-controlled phase IIa trial of three dosage levels of CP-690,550 versus placebo. Arthritis Rheum 2009; 60:1895–1905.
1542. Flammer AJ, Sudano I, Hermann F, Gay S, Forster A, Neidhart M, et al. Angiotensin-converting enzyme inhibition improves vascular function in rheumatoid arthritis. Circulation 2008; 117:2262–2269.
1543. Landgren AJ, Dehlin M, Jacobsson L, Bergsten U, Klingberg E. Cardiovascular risk factors in gout, psoriatic arthritis, rheumatoid arthritis and ankylosing spondylitis: a cross-sectional survey of patients in Western Sweden. RMD Open 2021; 7:
1544. Mehta NN, Yu Y, Pinnelas R, Krishnamoorthy P, Shin DB, Troxel AB, et al. Attributable risk estimate of severe psoriasis on major cardiovascular events. Am J Med 2011; 124:775e771–776.
1545. Azzouz B, De Guizelin A, Lambert A, Fresse A, Morel A, Trenque T. Psoriasis risk after beta-blocker exposure: Description of a pharmacovigilance signal. Br J Clin Pharmacol 2022; 88:3813–3818.
1546. Olszewski R, Cwiklinska K, Maslinska M, Kwiatkowska B. Prevention and risk assessment of cardiovascular events in a population of patients with psoriasis and psoriatic arthritis. Reumatologia 2022; 60:266–274.
1547. Frostegard J. Systemic lupus erythematosus and cardiovascular disease. J Intern Med 2023; 293:48–62.
1548. Lu X, Wang Y, Zhang J, Pu D, Hu N, Luo J, et al. Patients with systemic lupus erythematosus face a high risk of cardiovascular disease: A systematic review and Meta-analysis. Int Immunopharmacol 2021; 94:107466.
1549. Fernandez-Nebro A, Rua-Figueroa I, Lopez-Longo FJ, Galindo-Izquierdo M, Calvo-Alen J, Olive-Marques A, et al. Cardiovascular Events in Systemic Lupus Erythematosus: A Nationwide Study in Spain From the RELESSER Registry. Medicine (Baltimore) 2015; 94:e1183.
1550. Bursztyn M, Many A, Rosenthal T. Nifedipine in the treatment of hypertension in systemic lupus erythematosus. Angiology 1987; 38:359–362.
1551. Herlitz H, Edeno C, Mulec H, Westberg G, Aurell M. Captopril treatment of hypertension and renal failure in systemic lupus erythematosus. Nephron 1984; 38:253–256.
1552. Andrades C, Fuego C, Manrique-Arija S, Fernandez-Nebro A. Management of cardiovascular risk in systemic lupus erythematosus: a systematic review. Lupus 2017; 26:1407–1419.
1553. Tham YC, Li X, Wong TY, Quigley HA, Aung T, Cheng CY. Global prevalence of glaucoma and projections of glaucoma burden through 2040: a systematic review and meta-analysis. Ophthalmology 2014; 121:2081–2090.
1554. Jin J. Screening for Primary Open-Angle Glaucoma. JAMA 2022; 327:2030.
1555. Van Keer K, Breda JB, Pinto LA, Stalmans I, Vandewalle E. Estimating Mean Ocular Perfusion Pressure Using Mean Arterial Pressure and Intraocular Pressure. Invest Ophthalmol Vis Sci 2016; 57:2260.
1556. Zhao D, Cho J, Kim MH, Guallar E. The association of blood pressure and primary open-angle glaucoma: a meta-analysis. Am J Ophthalmol 2014; 158:615–627. e619.
1557. Levine RM, Yang A, Brahma V, Martone JF. Management of Blood Pressure in Patients with Glaucoma. Curr Cardiol Rep 2017; 19:109.
1558. De Moraes CG, Cioffi GA, Weinreb RN, Liebmann JM. New Recommendations for the Treatment of Systemic Hypertension and their Potential Implications for Glaucoma Management. J Glaucoma 2018; 27:567–571.
1559. Costa VP, Harris A, Anderson D, Stodtmeister R, Cremasco F, Kergoat H, et al. Ocular perfusion pressure in glaucoma. Acta Ophthalmol 2014; 92:e252–e266.
1560. Schmidl D, Garhofer G, Schmetterer L. The complex interaction between ocular perfusion pressure and ocular blood flow - relevance for glaucoma. Exp Eye Res 2011; 93:141–155.
1561. Rim TH, Lee SY, Kim SH, Kim SS, Kim CY. Increased incidence of open-angle glaucoma among hypertensive patients: an 11-year nationwide retrospective cohort study. J Hypertens 2017; 35:729–736.
1562. Horwitz A, Klemp M, Jeppesen J, Tsai JC, Torp-Pedersen C, Kolko M. Antihypertensive medication postpones the onset of glaucoma: evidence from a Nationwide study. Hypertension (Dallas, Tex: 1979) 2017; 69:202–210.
1563. Charlson ME, de Moraes CG, Link A, Wells MT, Harmon G, Peterson JC, et al. Nocturnal systemic hypotension increases the risk of glaucoma progression. Ophthalmology 2014; 121:2004–2012.
1564. Melgarejo JD, Lee JH, Petitto M, Yépez JB, Murati FA, Jin Z, et al. Glaucomatous optic neuropathy associated with nocturnal dip in blood pressure: findings from the maracaibo aging study. Ophthalmology 2018; 125:807–814.
1565. Krasińska B, Karolczak-Kulesza M, Krasiński Z, Pawlaczyk-Gabriel K, Lopatka P, Głuszek J, et al. Effects of the time of antihypertensive drugs administration on the stage of primary open-angle glaucoma in patients with arterial hypertension. Blood Press 2012; 21:240–248.
1566. Ho H, Shi Y, Chua J, Tham YC, Lim SH, Aung T, et al. Association of systemic medication use with intraocular pressure in a multiethnic asian population: the Singapore epidemiology of eye diseases study. JAMA Ophthalmol 2017; 135:196–202.
1567. Khawaja AP, Chan MP, Broadway DC, Garway-Heath DF, Luben R, Yip JL, et al. Systemic medication and intraocular pressure in a British population: the EPIC-Norfolk Eye Study. Ophthalmology 2014; 121:1501–1507.
1568. Zheng W, Dryja TP, Wei Z, Song D, Tian H, Kahler KH, et al. Systemic medication associations with presumed advanced or uncontrolled primary open-angle glaucoma. Ophthalmology 2018; 125:984–993.
1569. Wagner IV, Stewart MW, Dorairaj SK. Updates on the diagnosis and management of glaucoma. Mayo Clin Proc Innov Qual Outcomes 2022; 6:618–635.
1570. Huber M, Kölzsch M, Stahlmann R, Hofmann W, Bolbrinker J, Dräger D, et al. Ophthalmic drugs as part of polypharmacy in nursing home residents with glaucoma. Drugs Aging 2013; 30:31–38.
1571. van Dorst DCH, Dobbin SJH, Neves KB, Herrmann J, Herrmann SM, Versmissen J, et al. Hypertension and Prohypertensive Antineoplastic Therapies in Cancer Patients. Circ Res 2021; 128:1040–1061.
1572. Cohen JB, Brown NJ, Brown SA, Dent S, van Dorst DCH, Herrmann SM, et al. Cancer therapy-related hypertension: a scientific statement from the American Heart Association. Hypertension (Dallas, Tex: 1979) 2023; 80:e46–e57.
1573. Weikert S, Boeing H, Pischon T, Weikert C, Olsen A, Tjonneland A, et al. Blood pressure and risk of renal cell carcinoma in the European prospective investigation into cancer and nutrition. Am J Epidemiol 2008; 167:438–446.
1574. Kim CS, Han KD, Choi HS, Bae EH, Ma SK, Kim SW. Association of hypertension and blood pressure with kidney cancer risk: a nationwide population-based cohort study. Hypertension (Dallas, Tex: 1979) 2020; 75:1439–1446.
1575. Alcala K, Mariosa D, Smith-Byrne K, Nasrollahzadeh Nesheli D, Carreras-Torres R, Ardanaz Aicua E, et al. The relationship between blood pressure and risk of renal cell carcinoma. Int J Epidemiol 2022; 51:1317–1327.
1576. Tlemsani C, Mir O, Boudou-Rouquette P, Huillard O, Maley K, Ropert S, et al. Posterior reversible encephalopathy syndrome induced by anti-VEGF agents. Target Oncol 2011; 6:253–258.
1577. Herrmann J, Lenihan D, Armenian S, Barac A, Blaes A, Cardinale D, et al. Defining cardiovascular toxicities of cancer therapies: an International Cardio-Oncology Society (IC-OS) consensus statement. Eur Heart J 2022; 43:280–299.
1578. Ferrara N, Adamis AP. Ten years of anti-vascular endothelial growth factor therapy. Nat Rev Drug Discov 2016; 15:385–403.
1579. Versmissen J, Mirabito Colafella KM, Koolen SLW, Danser AHJ. Vascular cardio-oncology: vascular endothelial growth factor inhibitors and hypertension. Cardiovasc Res 2019; 115:904–914.
1580. Moslehi JJ. Cardiovascular toxic effects of targeted cancer therapies. N Engl J Med 2016; 375:1457–1467.
1581. Ranpura V, Pulipati B, Chu D, Zhu X, Wu S. Increased risk of high-grade hypertension with bevacizumab in cancer patients: a meta-analysis. Am J Hypertens 2010; 23:460–468.
1582. Zhu X, Stergiopoulos K, Wu S. Risk of hypertension and renal dysfunction with an angiogenesis inhibitor sunitinib: systematic review and meta-analysis. Acta Oncol 2009; 48:9–17.
1583. Wu S, Chen JJ, Kudelka A, Lu J, Zhu X. Incidence and risk of hypertension with sorafenib in patients with cancer: a systematic review and meta-analysis. The Lancet Oncology 2008; 9:117–123.
1584. Azizi M, Chedid A, Oudard S. Home blood-pressure monitoring in patients receiving sunitinib. N Engl J Med 2008; 358:95–97.
1585. Baek Moller N, Budolfsen C, Grimm D, Kruger M, Infanger M, Wehland M, et al. Drug-Induced Hypertension Caused by Multikinase Inhibitors (Sorafenib, Sunitinib, Lenvatinib and Axitinib) in Renal Cell Carcinoma Treatment. Int J Mol Sci 2019; 20.
1586. Campia U, Moslehi JJ, Amiri-Kordestani L, Barac A, Beckman JA, Chism DD, et al. Cardio-Oncology: Vascular and Metabolic Perspectives: A Scientific Statement From the American Heart Association. Circulation 2019; 139:e579–e602.
1587. Ghatalia P, Morgan CJ, Je Y, Nguyen PL, Trinh QD, Choueiri TK, et al. Congestive heart failure with vascular endothelial growth factor receptor tyrosine kinase inhibitors. Crit Rev Oncol Hematol 2015; 94:228–237.
1588. Eremina V, Jefferson JA, Kowalewska J, Hochster H, Haas M, Weisstuch J, et al. VEGF inhibition and renal thrombotic microangiopathy. N Engl J Med 2008; 358:1129–1136.
1589. Wu S, Kim C, Baer L, Zhu X. Bevacizumab increases risk for severe proteinuria in cancer patients. J Am Soc Nephrol 2010; 21:1381–1389.
1590. van Dorst DCH, van Doorn L, Mirabito Colafella KM, Manintveld OC, Hassing HC, Danser AHJ, et al. Cardiovascular toxicity of angiogenesis inhibitors and immune checkpoint inhibitors: synergistic anti-tumour effects at the cost of increased cardiovascular risk? Clin Sci (Lond) 2021; 135:1649–1668.
1591. Roviello G, Sigala S, Danesi R, Re MD, Bonetta A, Cappelletti MR, et al. Incidence and relative risk of adverse events of special interest in patients with castration resistant prostate cancer treated with CYP-17 inhibitors: A meta-analysis of published trials. Crit Rev Oncol Hematol 2016; 101:12–20.
1592. Attard G, Reid AH, Auchus RJ, Hughes BA, Cassidy AM, Thompson E, et al. Clinical and biochemical consequences of CYP17A1 inhibition with abiraterone given with and without exogenous glucocorticoids in castrate men with advanced prostate cancer. J Clin Endocrinol Metab 2012; 97:507–516.
1593. Izzedine H, Cluzel P, Deray G. Renal radiation-induced arterial stenosis. Kidney Int 2007; 71:1188.
1594. Weintraub NL, Jones WK, Manka D. Understanding radiation-induced vascular disease. J Am Coll Cardiol 2010; 55:1237–1239.
1595. Steingart RM, Bakris GL, Chen HX, Chen MH, Force T, Ivy SP, et al. Management of cardiac toxicity in patients receiving vascular endothelial growth factor signaling pathway inhibitors. American heart journal 2012; 163:156–163.
1596. Beavers CJ, Rodgers JE, Bagnola AJ, Beckie TM, Campia U, Di Palo KE, et al. Cardio-Oncology Drug Interactions: A Scientific Statement From the American Heart Association. Circulation 2022; 145:e811–e838.
1597. van Doorn L, Visser WJ, van Dorst DCH, Mirabito Colafella KM, Koolen SLW, de Mik AVE, et al. Dietary sodium restriction prevents vascular endothelial growth factor inhibitor-induced hypertension. Br J Cancer 2023; 128:354–362.
1598. van Dorst DCH, Kabadayi S, Oomen-de Hoop E, Danser AHJ, Mathijssen RHJ, Versmissen J. Treatment and Implications of Vascular Endothelial Growth Factor Inhibitor-Induced Blood Pressure Rise: A Clinical Cohort Study. J Am Heart Assoc 2023; 12:e028050.
1599. Gibson TM, Li Z, Green DM, Armstrong GT, Mulrooney DA, Srivastava D, et al. Blood pressure status in adult survivors of childhood cancer: a report from the St. Jude lifetime cohort study. Cancer Epidemiol Biomarkers Prev 2017; 26:1705–1713.
1600. Knight R, Walker V, Ip S, Cooper JA, Bolton T, Keene S, et al. Association of COVID-19 With Major Arterial and Venous Thrombotic Diseases: A Population-Wide Cohort Study of 48 Million Adults in England and Wales. Circulation 2022; 146:892–906.
1601. Xie Y, Xu E, Bowe B, Al-Aly Z. Long-term cardiovascular outcomes of COVID-19. Nat Med 2022; 28:583–590.
1602. Khairy Y, Naghibi D, Moosavi A, Sardareh M, Azami-Aghdash S. Prevalence of hypertension and associated risks in hospitalized patients with COVID-19: a meta-analysis of meta-analyses with 1468 studies and 1,281,510 patients. Syst Rev 2022; 11:242.
1603. Pavey H, Kulkarni S, Wood A, Ben-Shlomo Y, Sever P, McEniery C, et al. Primary hypertension, anti-hypertensive medications and the risk of severe COVID-19 in UK Biobank. PLoS One 2022; 17:e0276781.
1604. Corrao G, Rea F, Carle F, Scondotto S, Allotta A, Lepore V, et al. Stratification of the risk of developing severe or lethal Covid-19 using a new score from a large Italian population: a population-based cohort study. BMJ Open 2021; 11:e053281.
1605. Savoia C, Volpe M, Kreutz R. Hypertension, a Moving Target in COVID-19: Current Views and Perspectives. Circ Res 2021; 128:1062–1079.
1606. Drummond GR, Vinh A, Guzik TJ, Sobey CG. Immune mechanisms of hypertension. Nat Rev Immunol 2019; 19:517–532.
1607. Kreutz R, Algharably EAE, Azizi M, Dobrowolski P, Guzik T, Januszewicz A, et al. Hypertension, the renin-angiotensin system, and the risk of lower respiratory tract infections and lung injury: implications for COVID-19. Cardiovasc Res 2020; 116:1688–1699.
1608. Mancia G, Rea F, Ludergnani M, Apolone G, Corrao G. Renin-Angiotensin-Aldosterone System Blockers and the Risk of Covid-19. N Engl J Med 2020; 382:2431–2440.
1609. Reynolds HR, Adhikari S, Pulgarin C, Troxel AB, Iturrate E, Johnson SB, et al. Renin-Angiotensin-Aldosterone System Inhibitors and Risk of Covid-19. N Engl J Med 2020; 382:2441–2448.
1610. Zhang K, Cao L, Xuan N, Huang T, Tian B, Cui W, et al. The effect of renin-angiotensin-aldosterone system inhibitors in patients with hypertension and COVID-19: A meta-analysis of randomized controlled trials and propensity score-matched studies. J Intensive Med 2022; 2:282–290.
1611. Bavishi C, Whelton PK, Mancia G, Corrao G, Messerli FH. Renin-angiotensin-system inhibitors and all-cause mortality in patients with COVID-19: a systematic review and meta-analysis of observational studies. J Hypertens 2021; 39:784–794.
1612. Loader J, Taylor FC, Lampa E, Sundström J. Renin-Angiotensin Aldosterone System Inhibitors and COVID-19: A Systematic Review and Meta-Analysis Revealing Critical Bias Across a Body of Observational Research. J Am Heart Assoc 2022; 11:e025289.
1613. The corona-virus disease 2019 pandemic compromised routine care for hypertension: a survey conducted among excellence centers of the European Society of Hypertension. J Hypertens 2021; 39:190–195.
1614. Weber T, Amar J, de Backer T, Burkard T, van der Giet M, Gosse P, et al. Covid-19 associated reduction in hypertension-related diagnostic and therapeutic procedures in Excellence Centers of the European Society of Hypertension. Blood Press 2022; 31:71–79.
1615. Nielsen VM, Song G, Ojamaa LS, Blodgett RP, Rocchio CM, Pennock JN. The COVID-19 Pandemic and Access to Selected Ambulatory Care Services Among Populations With Severely Uncontrolled Diabetes and Hypertension in Massachusetts. Public Health Rep 2022; 137:344–351.
1616. Gotanda H, Liyanage-Don N, Moran AE, Krousel-Wood M, Green JB, Zhang Y, et al. Changes in Blood Pressure Outcomes Among Hypertensive Individuals During the COVID-19 Pandemic: A Time Series Analysis in Three US Healthcare Organizations. Hypertension (Dallas, Tex: 1979) 2022; 79:2733–2742.
1617. Steiner JF, Powers JD, Malone A, Lyons J, Olson K, Paolino AR, et al. Hypertension care during the COVID-19 pandemic in an integrated health care system. J Clin Hypertens (Greenwich) 2023; 25:315–325.
1618. Lee SG, Blood AJ, Cannon CP, Gordon WJ, Nichols H, Zelle D, et al. Remote Cardiovascular Hypertension Program Enhanced Blood Pressure Control During the COVID-19 Pandemic. J Am Heart Assoc 2023; e027296.
1619. Laffin LJ, Kaufman HW, Chen Z, Niles JK, Arellano AR, Bare LA, et al. Rise in Blood Pressure Observed Among US Adults During the COVID-19 Pandemic. Circulation 2022; 145:235–237.
1620. Shah NP, Clare RM, Chiswell K, Navar AM, Shah BR, Peterson ED. Trends of blood pressure control in the U.S. during the COVID-19 pandemic. American heart journal 2022; 247:15–23.
1621. Pengo MF, Albini F, Guglielmi G, Mollica C, Soranna D, Zambra G, et al. Home blood pressure during COVID-19-related lockdown in patients with hypertension. Eur J Prev Cardiol 2022; 29:e94–e96.
1622. Jiang J, Chan L, Kauffman J, Narula J, Charney AW, Oh W, et al. Impact of Vaccination on Major Adverse Cardiovascular Events in Patients With COVID-19 Infection. J Am Coll Cardiol 2023; 81:928–930.
1623. Kim YE, Huh K, Park YJ, Peck KR, Jung J. Association Between Vaccination and Acute Myocardial Infarction and Ischemic Stroke After COVID-19 Infection. JAMA 2022; 328:887–889.
1624. Buso G, Agabiti-Rosei C, Muiesan ML. The relationship between COVID-19 vaccines and increased blood pressure: A word of caution. Eur J Intern Med 2023.
1625. Angeli F, Reboldi G, Trapasso M, Santilli G, Zappa M, Verdecchia P. Blood pressure increase following COVID-19 vaccination: a systematic overview and meta-analysis. J Cardiovasc Dev Dis 2022; 9:
1626. Crook H, Raza S, Nowell J, Young M, Edison P. Long covid-mechanisms, risk factors, and management. BMJ 2021; 374:n1648.
1627. Thompson EJ, Williams DM, Walker AJ, Mitchell RE, Niedzwiedz CL, Yang TC, et al. Long COVID burden and risk factors in 10 UK longitudinal studies and electronic health records. Nat Commun 2022; 13:3528.
1628. Cohen K, Ren S, Heath K, Dasmariñas MC, Jubilo KG, Guo Y, et al. Risk of persistent and new clinical sequelae among adults aged 65 years and older during the post-acute phase of SARS-CoV-2 infection: retrospective cohort study. BMJ 2022; 376:e068414.
1629. Matsumoto C, Shibata S, Kishi T, Morimoto S, Mogi M, Yamamoto K, et al. Long COVID and hypertension-related disorders: a report from the Japanese Society of Hypertension Project Team on COVID-19. Hypertens Res 2023; 46:601–619.
1630. Godwin M, Birtwhistle R, Seguin R, Lam M, Casson I, Delva D, et al. Effectiveness of a protocol-based strategy for achieving better blood pressure control in general practice. Fam Pract 2010; 27:55–61.
1631. Morrison F, Shubina M, Turchin A. Encounter frequency and serum glucose level, blood pressure, and cholesterol level control in patients with diabetes mellitus. Arch Intern Med 2011; 171:1542–1550.
1632. Sherman L, Pelter MA, Deamer RL, Duan L, Batech M. Association between encounter frequency and time to blood pressure control among patients with newly diagnosed hypertension: a retrospective cohort study. The Journal of Clinical Hypertension 2018; 20:429–437.
1633. Turchin A, Goldberg SI, Shubina M, Einbinder JS, Conlin PR. Encounter frequency and blood pressure in hypertensive patients with diabetes mellitus. Hypertension (Dallas, Tex: 1979) 2010; 56:68–74.
1634. Roos NP, Carriere KC, Friesen D. Factors influencing the frequency of visits by hypertensive patients to primary care physicians in Winnipeg. CMAJ 1998; 159:777–783.
1635. Birtwhistle RV, Godwin MS, Delva MD, Casson RI, Lam M, MacDonald SE, et al. Randomised equivalence trial comparing three month and six month follow up of patients with hypertension by family practitioners. BMJ 2004; 328:204.
1636. Stergiou GS, Palatini P, Modesti PA, Asayama K, Asmar R, Bilo G, et al. Seasonal variation in blood pressure: Evidence, consensus and recommendations for clinical practice. Consensus statement by the European Society of Hypertension Working Group on Blood Pressure Monitoring and Cardiovascular Variability. J Hypertens 2020; 38:1235–1243.
1637. Sega R, Cesana G, Bombelli M, Grassi G, Stella ML, Zanchetti A, et al. Seasonal variations in home and ambulatory blood pressure in the PAMELA population. Pressione Arteriose Monitorate E Loro Associazioni. J Hypertens 1998; 16:1585–1592.
1638. Modesti PA, Morabito M, Massetti L, Rapi S, Orlandini S, Mancia G, et al. Seasonal blood pressure changes: an independent relationship with temperature and daylight hours. Hypertension (Dallas, Tex: 1979) 2013; 61:908–914.
1639. Gepts T, Nguyen AM, Cleland C, Wu W, Pham-Singer H, Shelley D. Accounting for Blood Pressure Seasonality Alters Evaluation of Practice-Level Blood Pressure Control Intervention. Am J Hypertens 2020; 33:220–222.
1640. Goyal A, Narang K, Ahluwalia G, Sohal PM, Singh B, Chhabra ST, et al. Seasonal variation in 24 h blood pressure profile in healthy adults- A prospective observational study. J Hum Hypertens 2019; 33:626–633.
1641. Parati G, Agostoni P, Basnyat B, Bilo G, Brugger H, Coca A, et al. Clinical recommendations for high altitude exposure of individuals with pre-existing cardiovascular conditions: A joint statement by the European Society of Cardiology, the Council on Hypertension of the European Society of Cardiology, the European Society of Hypertension, the International Society of Mountain Medicine, the Italian Society of Hypertension and the Italian Society of Mountain Medicine. Eur Heart J 2018; 39:1546–1554.
1642. Greer N, Bolduc J, Geurkink E, Rector T, Olson K, Koeller E, et al. Pharmacist-Led Chronic Disease Management: A Systematic Review of Effectiveness and Harms Compared With Usual Care. Ann Intern Med 2016; 165:30–40.
1643. Massimi A, De Vito C, Brufola I, Corsaro A, Marzuillo C, Migliara G, et al. Are community-based nurse-led self-management support interventions effective in chronic patients? Results of a systematic review and meta-analysis. PLoS One 2017; 12:e0173617.
1644. Mundinger MO, Kane RL, Lenz ER, Totten AM, Tsai WY, Cleary PD, et al. Primary care outcomes in patients treated by nurse practitioners or physicians: a randomized trial. JAMA 2000; 283:59–68.
1645. Tsuyuki RT, Houle SK, Charrois TL, Kolber MR, Rosenthal MM, Lewanczuk R, et al. Randomized Trial of the Effect of Pharmacist Prescribing on Improving Blood Pressure in the Community: The Alberta Clinical Trial in Optimizing Hypertension (RxACTION). Circulation 2015; 132:93–100.
1646. Vrijens B, De Geest S, Hughes DA, Przemyslaw K, Demonceau J, Ruppar T, et al. A new taxonomy for describing and defining adherence to medications. Br J Clin Pharmacol 2012; 73:691–705.
1647. Tajeu GS, Kent ST, Huang L, Bress AP, Cuffee Y, Halpern MT, et al. Antihypertensive Medication Nonpersistence and Low Adherence for Adults <65 Years Initiating Treatment in 2007–2014. Hypertension (Dallas Tex: 1979) 2019; 74:35–46.
1648. Corrao G, Zambon A, Parodi A, Merlino L, Mancia G. Incidence of cardiovascular events in Italian patients with early discontinuations of antihypertensive, lipid-lowering, and antidiabetic treatments. Am J Hypertens 2012; 25:549–555.
1649. Bruno A, Brooks DD, Abrams TA, Poorak MD, Gunio D, Kandhal PK, et al. Left ventricular hypertrophy in acute stroke patients with known hypertension. Clin Exp Hypertens 2017; 39:502–504.
1650. Kim YS, Kim HS, Oh HY, Lee MK, Kim CH, Kim YS, et al. Prevalence of microalbuminuria and associated risk factors among adult Korean hypertensive patients in a primary care setting. Hypertens Res 2013; 36:807–823.
1651. Mazzaglia G, Ambrosioni E, Alacqua M, Filippi A, Sessa E, Immordino V, et al. Adherence to antihypertensive medications and cardiovascular morbidity among newly diagnosed hypertensive patients. Circulation 2009; 120:1598–1605.
1652. Corrao G, Rea F, Ghirardi A, Soranna D, Merlino L, Mancia G. Adherence with antihypertensive drug therapy and the risk of heart failure in clinical practice. Hypertension (Dallas, Tex: 1979) 2015; 66:742–749.
1653. Cedillo-Couvert EA, Ricardo AC, Chen J, Cohan J, Fischer MJ, Krousel-Wood M, et al. Self-reported Medication Adherence and CKD Progression. Kidney Int Rep 2018; 3:645–651.
1654. Roy L, White-Guay B, Dorais M, Dragomir A, Lessard M, Perreault S. Adherence to antihypertensive agents improves risk reduction of end-stage renal disease. Kidney Int 2013; 84:570–577.
1655. Lee EKP, Poon P, Yip BHK, Bo Y, Zhu MT, Yu CP, et al. Global burden, regional differences, trends, and health consequences of medication nonadherence for hypertension during 2010 to 2020: a meta-analysis involving 27 million patients. J Am Heart Assoc 2022; 11:e026582.
1656. Whelton PK, Carey RM, Mancia G, Kreutz R, Bundy JD, Williams B. Harmonization of the American College of Cardiology/American Heart Association and European Society of Cardiology/European Society of Hypertension Blood Pressure/Hypertension Guidelines: Comparisons, Reflections, and Recommendations. J Am Coll Cardiol 2022; 80:1192–1201.
1657. Pareja-Martinez E, Esquivel-Prados E, Martinez-Martinez F, Garcia-Corpas JP. Questionnaires on adherence to antihypertensive treatment: a systematic review of published questionnaires and their psychometric properties. Int J Clin Pharm 2020; 42:355–365.
1658. Glasser SP, Vitolins M, Rocco MV, Still CH, Cofield SS, Haley WE, et al. Is Medication Adherence Predictive of Cardiovascular Outcomes and Blood Pressure Control? The Systolic Blood Pressure Intervention Trial (SPRINT). Am J Hypertens 2022; 35:182–191.
1659. Fadl Elmula FE, Hoffmann P, Larstorp AC, Fossum E, Brekke M, Kjeldsen SE, et al. Adjusted drug treatment is superior to renal sympathetic denervation in patients with true treatment-resistant hypertension. Hypertension (Dallas, Tex: 1979) 2014; 63:991–999.
1660. Gupta P, Patel P, Horne R, Buchanan H, Williams B, Tomaszewski M. How to screen for non-adherence to antihypertensive therapy. Curr Hypertens Rep 2016; 18:89.
1661. Sabaté E. Adherence to long-term therapies: evidence for action. World Health Organization 2003:ISBN: 9241545992.
1662. Simpson J, Jackson CE, Haig C, Jhund PS, Tomaszewski M, Gardner RS, et al. Adherence to prescribed medications in patients with heart failure: insights from liquid chromatography-tandem mass spectrometry-based urine analysis. Eur Heart J Cardiovasc Pharmacother 2021; 7:296–301.
1663. Corrao G, Soranna D, Merlino L, Mancia G. Similarity between generic and brand-name antihypertensive drugs for primary prevention of cardiovascular disease: evidence from a large population-based study. Eur J Clin Invest 2014; 44:933–939.
1664. Patel P, Gupta PK, White CM, Stanley AG, Williams B, Tomaszewski M. Screening for non-adherence to antihypertensive treatment as a part of the diagnostic pathway to renal denervation. J Hum Hypertens 2016; 30:368–373.
1665. van Schoonhoven AV, van Asselt ADI, Tomaszewski M, Patel P, Khunti K, Gupta P, et al. Cost-utility of an objective biochemical measure to improve adherence to antihypertensive treatment. Hypertension (Dallas, Tex: 1979) 2018; 72:1117–1124.
1666. Schneider MP, Burnier M. Partnership between patients and interprofessional healthcare providers along the multifaceted journey to medication adherence. Br J Clin Pharmacol 2022.
1667. Carter BL, Rogers M, Daly J, Zheng S, James PA. The potency of team-based care interventions for hypertension: a meta-analysis. Arch Intern Med 2009; 169:1748–1755.
1668. Carter BL, Bosworth HB, Green BB. The hypertension team: the role of the pharmacist, nurse, and teamwork in hypertension therapy. J Clin Hypertens (Greenwich) 2012; 14:51–65.
1669. Margolis KL, Asche SE, Dehmer SP, Bergdall AR, Green BB, Sperl-Hillen JM, et al. Long-term Outcomes of the Effects of Home Blood Pressure Telemonitoring and Pharmacist Management on Blood Pressure Among Adults With Uncontrolled Hypertension: Follow-up of a Cluster Randomized Clinical Trial. JAMA Netw Open 2018; 1:e181617.
1670. Phillips LS, Branch WT, Cook CB, Doyle JP, El-Kebbi IM, Gallina DL, et al. Clinical inertia. Ann Intern Med 2001; 135:825–834.
1671. Chow CK, Teo KK, Rangarajan S, Islam S, Gupta R, Avezum A, et al. Prevalence, awareness, treatment, and control of hypertension in rural and urban communities in high-, middle-, and low-income countries. JAMA 2013; 310:959–968.
1672. Wolf-Maier K, Cooper RS, Kramer H, Banegas JR, Giampaoli S, Joffres MR, et al. Hypertension treatment and control in five European countries, Canada, and the United States. Hypertension (Dallas, Tex: 1979) 2004; 43:10–17.
1673. Ali DH, Kilic B, Hart HE, Bots ML, Biermans MCJ, Spiering W, et al. Therapeutic inertia in the management of hypertension in primary care. J Hypertens 2021; 39:1238–1245.
1674. Banegas JR, Segura J, Ruilope LM, Luque M, Garcia-Robles R, Campo C, et al. Blood pressure control and physician management of hypertension in hospital hypertension units in Spain. Hypertension (Dallas, Tex: 1979) 2004; 43:1338–1344.
1675. De Backer T, Van Nieuwenhuyse B, De Bacquer D. Antihypertensive treatment in a general uncontrolled hypertensive population in Belgium and Luxembourg in primary care: Therapeutic inertia and treatment simplification. The SIMPLIFY study. PLoS One 2021; 16:e0248471.
1676. Guthrie B, Inkster M, Fahey T. Tackling therapeutic inertia: role of treatment data in quality indicators. BMJ 2007; 335:542–544.
1677. Okonofua EC, Simpson KN, Jesri A, Rehman SU, Durkalski VL, Egan BM. Therapeutic inertia is an impediment to achieving the Healthy People 2010 blood pressure control goals. Hypertension (Dallas, Tex: 1979) 2006; 47:345–351.
1678. Redon J, Coca A, Lazaro P, Aguilar MD, Cabanas M, Gil N, et al. Factors associated with therapeutic inertia in hypertension: validation of a predictive model. J Hypertens 2010; 28:1770–1777.
1679. Ferrari P, Hess L, Pechere-Bertschi A, Muggli F, Burnier M. Reasons for not intensifying antihypertensive treatment (RIAT): a primary care antihypertensive intervention study. J Hypertens 2004; 22:1221–1229.
1680. Wald DS, Law M, Morris JK, Bestwick JP, Wald NJ. Combination therapy versus monotherapy in reducing blood pressure: meta-analysis on 11,000 participants from 42 trials. Am J Med 2009; 122:290–300.
1681. Williams B, Mancia G, Spiering W, Agabiti Rosei E, Azizi M, Burnier M, et al. 2018 Practice Guidelines for the management of arterial hypertension of the European Society of Hypertension and the European Society of Cardiology: ESH/ESC Task Force for the Management of Arterial Hypertension. J Hypertens 2018; 36:2284–2309.
1682. Agarwal R, Bills JE, Hecht TJ, Light RP. Role of home blood pressure monitoring in overcoming therapeutic inertia and improving hypertension control: a systematic review and meta-analysis. Hypertension (Dallas, Tex: 1979) 2011; 57:29–38.
1683. Egan BM, Sutherland SE, Rakotz M, Yang J, Hanlin RB, Davis RA, et al. Improving hypertension control in primary care with the measure accurately, act rapidly, and partner with patients protocol. Hypertension (Dallas, Tex: 1979) 2018; 72:1320–1327.
1684. Petersen LA, Simpson K, Pietz K, Urech TH, Hysong SJ, Profit J, et al. Effects of individual physician-level and practice-level financial incentives on hypertension care: a randomized trial. JAMA 2013; 310:1042–1050.
1685. Lebeau JP, Cadwallader JS, Aubin-Auger I, Mercier A, Pasquet T, Rusch E, et al. The concept and definition of therapeutic inertia in hypertension in primary care: a qualitative systematic review. BMC Fam Pract 2014; 15:130.
1686. Zhao J, Hu Y, Zhang X, Zhang G, Lin M, Chen X, et al. Efficacy of empowerment strategies for patients with hypertension: A systematic review and meta-analysis. Patient Educ Couns 2020; 103:898–907.
1687. Johnson RA, Huntley A, Hughes RA, Cramer H, Turner KM, Perkins B, et al. Interventions to support shared decision making for hypertension: A systematic review of controlled studies. Health Expect 2018; 21:1191–1207.
1688. Pathak A, Poulter NR, Kavanagh M, Kreutz R, Burnier M. Improving the management of hypertension by tackling awareness, adherence, and clinical inertia: a symposium report. Am J Cardiovasc Drugs 2022; 22:251–261.
1689. King CC, Bartels CM, Magnan EM, Fink JT, Smith MA, Johnson HM. The importance of frequent return visits and hypertension control among US young adults: a multidisciplinary group practice observational study. J Clin Hypertens (Greenwich) 2017; 19:1288–1297.
1690. Cook NR, Cutler JA, Obarzanek E, Buring JE, Rexrode KM, Kumanyika SK, et al. Long term effects of dietary sodium reduction on cardiovascular disease outcomes: observational follow-up of the trials of hypertension prevention (TOHP). BMJ 2007; 334:885–888.
1691. Gupta AK, McGlone M, Greenway FL, Johnson WD. Prehypertension in disease-free adults: a marker for an adverse cardiometabolic risk profile. Hypertens Res 2010; 33:905–910.
1692. Julius S, Nesbitt SD, Egan BM, Weber MA, Michelson EL, Kaciroti N, et al. Feasibility of treating prehypertension with an angiotensin-receptor blocker. N Engl J Med 2006; 354:1685–1697.
1693. Thompson AM, Hu T, Eshelbrenner CL, Reynolds K, He J, Bazzano LA. Antihypertensive treatment and secondary prevention of cardiovascular disease events among persons without hypertension: a meta-analysis. JAMA 2011; 305:913–922.
1694. Viera AJ, Bangura F, Mitchell CM, Cerna A, Sloane P. Do clinicians tell patients they have prehypertension? Journal of the American Board of Family Medicine: JABFM 2011; 24:117–118.
1695. McLean G, Band R, Saunderson K, Hanlon P, Murray E, Little P, et al. Digital interventions to promote self-management in adults with hypertension systematic review and meta-analysis. J Hypertens 2016; 34:600–612.
1696. McManus RJ, Mant J, Bray EP, Holder R, Jones MI, Greenfield S, et al. Telemonitoring and self-management in the control of hypertension (TASMINH2): a randomised controlled trial. Lancet 2010; 376:163–172.
1697. Widmer RJ, Collins NM, Collins CS, West CP, Lerman LO, Lerman A. Digital health interventions for the prevention of cardiovascular disease: a systematic review and meta-analysis. Mayo Clin Proc 2015; 90:469–480.
1698. Kumar N, Khunger M, Gupta A, Garg N. A content analysis of smartphone-based applications for hypertension management. J Am Soc Hypertens 2015; 9:130–136.
1699. Parati G, Torlasco C, Omboni S, Pellegrini D. Smartphone applications for hypertension management: a potential game-changer that needs more control. Curr Hypertens Rep 2017; 19:48.
1700. Rossi GP, Bisogni V. European Society of Hypertension Working Group on Endocrine H. A useful tool to improve the case detection rate of primary aldosteronism: the aldosterone-renin ratio (ARR)-App. J Hypertens 2016; 34:1019–1021.
1701. Duan Y, Xie Z, Dong F, Wu Z, Lin Z, Sun N, et al. Effectiveness of home blood pressure telemonitoring: a systematic review and meta-analysis of randomised controlled studies. J Hum Hypertens 2017; 31:427–437.
1702. Omboni S, Padwal RS, Alessa T, Benczúr B, Green BB, Hubbard I, et al. The worldwide impact of telemedicine during COVID-19: current evidence and recommendations for the future. Connect Health 2022; 1:7–35.
1703. Johansson M, Guyatt G, Montori V. Guidelines should consider clinicians’ time needed to treat. BMJ 2023; 380:e072953.
1704. He J, Ouyang N, Guo X, Sun G, Li Z, Mu J, et al. Effectiveness of a non-physician community health-care provider-led intensive blood pressure intervention versus usual care on cardiovascular disease (CRHCP): an open-label, blinded-endpoint, cluster-randomised trial. Lancet 2023; 401:928–938.
1705. Schwalm JD, McCready T, Lopez-Jaramillo P, Yusoff K, Attaran A, Lamelas P, et al. A community-based comprehensive intervention to reduce cardiovascular risk in hypertension (HOPE 4): a cluster-randomised controlled trial. Lancet 2019; 394:1231–1242.
1706. Albasri A, Prinjha S, McManus RJ, Sheppard JP. Hypertension referrals from community pharmacy to general practice: multivariate logistic regression analysis of 131 419 patients. Br J Gen Pract 2018; 68:e541–e550.
1707. Fenna J, Chu C, Hassan R, Gomes T, Tadrous M. Extent of a valsartan drug shortage and its effect on antihypertensive drug use in the Canadian population: a national cross-sectional study. CMAJ Open 2021; 9:E1128–E1133.
1708. Pharmaceutical group of the European Union: https://www.pgeu.eu/ (visited on October 28th).
1709. Cheema E, Sutcliffe P, Singer DR. The impact of interventions by pharmacists in community pharmacies on control of hypertension: a systematic review and meta-analysis of randomized controlled trials. Br J Clin Pharmacol 2014; 78:1238–1247.
1710. Manouchehri M, Fernández-Alfonso MS, Gil-Ortega M. Impact of intervention of community pharmacists on cardiovascular outcomes in Spain: A systematic review. Journal of Pharmacy & Pharmacognosy Research 2022; 10:952–976.
1711. Milosavljevic A, Aspden T, Harrison J. Community pharmacist-led interventions and their impact on patients’ medication adherence and other health outcomes: a systematic review. Int J Pharm Pract 2018; 26:387–397.
1712. Reeves L, Robinson K, McClelland T, Adedoyin CA, Broeseker A, Adunlin G. Pharmacist Interventions in the Management of Blood Pressure Control and Adherence to Antihypertensive Medications: A Systematic Review of Randomized Controlled Trials. J Pharm Pract 2021; 34:480–492.
1713. Santschi V, Chiolero A, Colosimo AL, Platt RW, Taffé P, Burnier M, et al. Improving blood pressure control through pharmacist interventions: a meta-analysis of randomized controlled trials. J Am Heart Assoc 2014; 3:e000718.
1714. Steed L, Sohanpal R, Todd A, Madurasinghe VW, Rivas C, Edwards EA, et al. Community pharmacy interventions for health promotion: effects on professional practice and health outcomes. Cochrane Database Syst Rev 2019; 12:CD011207.
1715. Santschi V, Chiolero A, Burnand B, Colosimo AL, Paradis G. Impact of pharmacist care in the management of cardiovascular disease risk factors: a systematic review and meta-analysis of randomized trials. Arch Intern Med 2011; 171:1441–1453.
1716. So R, Al Hamarneh YN, Oleksyn C, Purschke M, Tsuyuki RT. Impact of a “Pharmacist First” innovative workflow plan in patients with hypertension and/or diabetes. Can Pharm J (Ott) 2021; 154:376–380.
1717. Viigimaa M, Farsang C, Kjeldsen SE, Narkiewicz K, Mancia G. ESH Hypertension Excellence Centres: a new strategy to combat an old foe. Journal of Hypertension 2007; 25:1744–1748.
1718. Burnier M, Prejbisz A, Weber T, Azizi M, Cunha V, Versmissen J, et al. Hypertension healthcare professional beliefs and behaviour regarding patient medication adherence: a survey conducted among European Society of Hypertension Centres of Excellence. Blood Press 2021; 30:282–290.
1719. Li Q, Li R, Zhang S, Zhang Y, He P, Zhang Z, et al. Occupational Physical Activity and New-Onset Hypertension: A Nationwide Cohort Study in China. Hypertension (Dallas, Tex: 1979) 2021; 78:220–229.
1720. Manohar S, Thongprayoon C, Cheungpasitporn W, Mao MA, Herrmann SM. Associations of rotational shift work and night shift status with hypertension: a systematic review and meta-analysis. J Hypertens 2017; 35:1929–1937.
1721. Gamboa Madeira S, Fernandes C, Paiva T, Santos Moreira C, Caldeira D. The Impact of Different Types of Shift Work on Blood Pressure and Hypertension: A Systematic Review and Meta-Analysis. Int J Environ Res Public Health 2021; 18.
1722. Badr HE, Rao S, Manee F. Gender differences in quality of life, physical activity, and risk of hypertension among sedentary occupation workers. Quality of life research: an international journal of quality of life aspects of treatment, care and rehabilitation 2021; 30:1365–1377.
1723. Faruque MO, Framke E, Sorensen JK, Madsen IEH, Rugulies R, Vonk JM, et al. Psychosocial work factors and blood pressure among 63 800 employees from The Netherlands in the Lifelines Cohort Study. J Epidemiol Community Health 2022; 76:60–66.
1724. Liu X, Matthews TA, Chen L, Li J. The associations of job strain and leisure-time physical activity with the risk of hypertension: the population-based Midlife in the United States cohort study. Epidemiol Health 2022; 44:e2022073.
1725. Fu W, Wang C, Zou L, Liu Q, Gan Y, Yan S, et al. Association between exposure to noise and risk of hypertension: a meta-analysis of observational epidemiological studies. J Hypertens 2017; 35:2358–2366.
1726. Moretti Anfossi C, Ahumada Munoz M, Tobar Fredes C, Perez Rojas F, Ross J, Head J, et al. Work Exposures and Development of Cardiovascular Diseases: A Systematic Review. Ann Work Expo Health 2022; 66:698–713.
1727. Lecours A, Major M, Lederer V, Vincent C, Lamontagne M, Drolet AA. Integrative Prevention at Work: A Concept Analysis and Meta-Narrative Review. J Occup Rehabil 2022; 1–15.
1728. Chowdhury R, Khan H, Heydon E, Shroufi A, Fahimi S, Moore C, et al. Adherence to cardiovascular therapy: a meta-analysis of prevalence and clinical consequences. Eur Heart J 2013; 34:2940–2948.
1729. James S, Rao SV, Granger CB. Registry-based randomized clinical trials--a new clinical trial paradigm. Nat Rev Cardiol 2015; 12:312–316.
1730. Zhang C, Lu Y. Study on artificial intelligence: The state of the art and future prospects. Journal of Industrial Information Integration 2021; 23:100224.
1731. Dong S, Wang P, Abbas K. A survey on deep learning and its applications. Computer Science Review 2021; 40:100379.
1732. Breiman L. Statistical Modeling: The Two Cultures (with comments and a rejoinder by the author). Statistical Science 2001; 16.
1733. Silva GFS, Fagundes TP, Teixeira BC, Chiavegatto Filho ADP. Machine Learning for Hypertension Prediction: a Systematic Review. Current Hypertension Reports 2022; 24:523–533.
1734. Niu M, Wang Y, Zhang L, Tu R, Liu X, Hou J, et al. Identifying the predictive effectiveness of a genetic risk score for incident hypertension using machine learning methods among populations in rural China. Hypertens Res 2021; 44:1483–1491.
1735. Oikonomou EK, Spatz ES, Suchard MA, Khera R. Individualising intensive systolic blood pressure reduction in hypertension using computational trial phenomaps and machine learning: a post-hoc analysis of randomised clinical trials. Lancet Digit Health 2022; 4:e796–e805.
1736. Koren G, Nordon G, Radinsky K, Shalev V. Machine learning of big data in gaining insight into successful treatment of hypertension. Pharmacol Res Perspect 2018; 6:e00396.

Giuseppe Mancia and Reinhold Kreutz contributed equally to the document.

Keywords:

antihypertensive device interventions; antihypertensive drug therapy; blood pressure; cardiac disease; cardiovascular disease; drug combinations; guidelines; heart failure; hypertension; kidney disease; lifestyle interventions; patient's follow-up; organ damage; secondary hypertension; stroke; AASK; African American study on kidney disease; ACC; American College of Cardiology; ACCOMPLISH; Avoiding Cardiovascular Events through Combination Therapy in Patients Living with Systolic Hypertension; ACCORD; Action to control cardiovascular risk in diabetes; ADVANCE; Action in Diabetes and VAscular Disease – PreterAx and DiamicroN Controlled Evaluation; AHA; American Heart Association; ALLHAT; Antihypertensive and Lipid-Lowering Treatment to Prevent Heart Attack Trial; ALTITUDE; Aliskiren Trial in Type 2 Diabetes Using Cardiorenal Endpoints; AMBER; Spironolactone With Patiromer in the Treatment of Resistant Hypertension in Chronic Kidney Disease; AMI; US Association for Advance in Medical Instrumentation; ANBP-2; Second Australian National Blood Pressure Study; ASCOT; Anglo-Scandinavian Cardiac Outcomes Trial; ATTACH-2; Antihypertensive Treatment of Acute Cerebral Hemorrhage II; BUMP-1; Blood Pressure Monitoring in High Risk Pregnancy to Improve the Detection and Monitoring of Hypertension-1; BUMP-2; Blood Pressure Monitoring in High Risk Pregnancy to Improve the Detection and Monitoring of Hypertension-2; CALM-FIM; Controlling and Lowering Blood Pressure With the MobiusHD – First in Man; CAPP; The Captopril Prevention Project Study; CARDIA; Coronary Artery Risk Development in Young Adults; CHAP; Chronic Hypertension and Pregnancy; CHIPS; Control of Hypertension in Pregnancy Study; CLICK; Chlorthalidone in Chronic Kidney disease trial; COLM; Combination of OLMesartan and a calcium channel blocker or diuretic in Japanese elderly patients trial; CONVINCE; Controlled ONset Verapamil INvestigation of Cardiovascular Endpoints; COPE; Combination Therapy of Hypertension to Prevent Cardiovascular Events; CREDENCE; Canagliflozin and Renal Events in Diabetes with Established Nephropathy Clinical Evaluation; CREOLE; Comparison of Three Combination Therapies in Lowering Blood Pressure in Black Africans; DAPA-CKD; Dapagliflozin and Prevention of Adverse Outcomes in Chronic Kidney Disease; DASH; Dietary Approaches to Stop Hypertension; DCP; Diuretic Comparison Project; EASD; European Association for the Study of Diabetes; ELSA; European Lacidipine Study on Atherosclerosis; EMPA KIDNEY TRIAL; The Study of Heart and Kidney Protection With Empagliflozin trial; ESC; European Society of Cardiology; ESH; European Society of Hypertension; EUCLID; EURODIAB Controlled Trial of Lisinopril in Insulin-dependent Diabetes Mellitus; EUROPA; Efficacy of perindopril in reduction of cardiovascular events among patients with stable coronary artery disease; FEVER; Felodipine Event Reduction; FIDELIO DKD; Finerenone in Reducing Kidney Failure and Disease Progression in Diabetic Kidney Disease; FIDELITY; The Finerenone in chronic kiDney disease and type 2 diabetes: Combined FIDELIO-DKD and FIGARO-DKD Trial programme analysis; FIGARO DKD; Finerenone in Reducing Cardiovascular Mortality and Morbidity in Diabetic Kidney Disease; GATEWAY; Gastric Bypass to Treat Obese Patients With Steady Hypertension; GRADE; Grade of Recommendation, Assessment, Development and Evaluation; HAS-BLED; Hypertension, Abnormal Renal/Liver Function, Stroke, Bleeding History or Predisposition, Labile INR, Elderly, Drugs/Alcohol Concomitantly; HOPE-3; Heart Outcomes Prevention Evaluation-3; HOT; Hypertension Optimal Treatment; HYVET; Hypertension in the Very Elderly Trial; IDACO; International Database of Ambulatory blood pressure in relation to Cardiovascular Outcome.; IDNT; Irbesartan diabetic nephropathy trial; INTERACT-2; Second Intensive Blood Pressure Reduction in Acute Cerebral Haemorrhage Trial; INVEST; INternational Verapamil SR TRandolapril STudy; ISO; International Organization for Standardization; LEGEND; Large-scale Evidence Generation and Evaluation across a Network of Databases; LIFE; Losartan Intervention For Endpoint reduction in hypertension; MATCH; Is Metomidate PET-CT superior to Adrenal vein sampling in predicting ouTCome from adrenalectomy in patients with primary Hyperaldosteronism; MDRD; Modification of diet in renal disease; MODERATO I; Moderato System in Patients With Hypertension I; MODERATO II; Moderato System in Patients With Hypertension II; NORDIL; Nordic Diltiazem; ONTARGET; Ongoing Telmisartan Alone and in Combination with Ramipril Global Endpoint Trial; OSCAR; The OlmeSartan and Calcium Antagonists Randomized study; PAMELA; Pressioni Arteriose Monitorate e Loro Associazioni; PATHWAY-1; Optimum Treatment for Drug-Resistant Hypertension -1; PATHWAY-2; Optimum Treatment for Drug-Resistant Hypertension -2; PATS; Post-stroke Antihypertensive Treatment Study; PEACE; Prevention of Events with Angiotensin Converting Enzyme Inhibition; PHYLLIS; Plaque Hypertension Lipid Lowering Italian Study; PRAISE; Prospective Randomized Amlodipine Survival Evaluation study; PREDIVA; Prevention of dementia by intensive vascular care; PROGRESS; perindopril protection against recurrent stroke study; PROSIT; Project on Stroke Services in Italy; RADIANCE HTN SOLO; A Study of the ReCor Medical Paradise System in Clinical Hypertension – SOLO; RADIANCE HTN-TRIO; A Study of the ReCor Medical Paradise System in Clinical Hypertension -TRIO; RE-HOT; Resistant Hypertension Optimal Treatment; RENAAL; Reduction of endpoints in NIDDM with angiotensin II antagonist losartan; REQUIRE; Renal Denervation on Quality of 24-hr BP Control by Ultrasound In Resistant Hypertension; ROX CONTROL HTN; ROX Coupler in Patients With Resistant Hypertension study; SCOPE; Study on Cognition and Prognosis in the Elderly; SCORE-2; Systemic coronary risk evaluation; SCORE-2-OP; Systemic coronary risk evaluation – Older Persons; SHEP; Systolic Hypertension in the Elderly Program; SOS; Swedish obese individual study; SPRINT; Systolic Blood Pressure Intervention Trial; SPRINT-MIND; Systolic Blood Pressure Intervention Trial – Memory and Cognition in Decreased Hypertension; SPYRAL HTN Off Med; catheter-based renal denervation in the absence of antihypertensive medications, multicentre, randomised, sham-controlled trial; SPYRAL HTN – On Med; catheter-based renal denervation in the presence of antihypertensive medications, multicentre, randomised, sham-controlled trial; SSa; Subsaharian Africa; STOP-1; Swedish Trial in Old Patients-1; STOP-2; Swedish Trial in Old Patients-2; STRIDE BP; Science and Technology for Regional Innovation and Development in Europe – Blood Pressure; SYMPLICITY-3 HTN; Renal Denervation in Patients With Uncontrolled Hypertension trial; SYST-CHINA; Systolic Hypertension in China; SYST-EUR; Systolic Hypertension in Europe; TAIM; Trial on antihypertensive intervention and management; TIME; Treatment in Morning vs Evening study; TOMHS; Treatment of Mild Hypertension Study; TONE; Trial on nonpharmacological intervention in the elderly; TOPCAT; Treatment of Preserved Cardiac Function Heart Failure With an Aldosterone Antagonist; TOPH; Trials of Hypertension Prevention; TRANSCEND; Telmisartan Randomized AssessmeNt Study in ACE iNtolerant subjects with cardiovascular Disease; UKPDS; United Kingdom prospective diabetes study; VALISH; Valsartan in elderly isolated systolic hypertension study; VALUE; Valsartan Antihypertensive Long-Term Use Evaluation; WHL; World Heart Federation; WHO; World Health Organization

Copyright © 2023 Wolters Kluwer Health, Inc. All rights reserved.