Research Article
21 October 2015

Cross-Immunity and Community Structure of a Multiple-Strain Pathogen in the Tick Vector

ABSTRACT

Many vector-borne pathogens consist of multiple strains that circulate in both the vertebrate host and the arthropod vector. Characterization of the community of pathogen strains in the arthropod vector is therefore important for understanding the epidemiology of mixed vector-borne infections. Borrelia afzelii and B. garinii are two species of tick-borne bacteria that cause Lyme disease in humans. These two sympatric pathogens use the same tick, Ixodes ricinus, but are adapted to different classes of vertebrate hosts. Both Borrelia species consist of multiple strains that are classified using the highly polymorphic ospC gene. Vertebrate cross-immunity against the OspC antigen is predicted to structure the community of multiple-strain Borrelia pathogens. Borrelia isolates were cultured from field-collected I. ricinus ticks over a period spanning 11 years. The Borrelia species of each isolate was identified using a reverse line blot (RLB) assay. Deep sequencing was used to characterize the ospC communities of 190 B. afzelii isolates and 193 B. garinii isolates. Infections with multiple ospC strains were common in ticks, but vertebrate cross-immunity did not influence the strain structure in the tick vector. The pattern of genetic variation at the ospC locus suggested that vertebrate cross-immunity exerts strong selection against intermediately divergent ospC alleles. Deep sequencing found that more than 50% of our isolates contained exotic ospC alleles derived from other Borrelia species. Two alternative explanations for these exotic ospC alleles are cryptic coinfections that were not detected by the RLB assay or horizontal transfer of the ospC gene between Borrelia species.

INTRODUCTION

Many vector-borne pathogens consist of multiple genetically distinct strains (14). The adaptive arm of the vertebrate immune system plays a key role in generating and maintaining this diversity of pathogen strains (57). Genetic diversity is often the highest at loci coding for surface-exposed pathogen molecules that function during the invasion and infection of host tissues (8, 9). The study of these highly polymorphic pathogen molecules is important for understanding how cross-reactive acquired immunity can mediate indirect competition and superinfection in the vertebrate host (10, 11). In addition, these pathogen outer surface proteins are often used to characterize pathogen strains because they provide an upper estimate of pathogen strain richness.
In vector-borne diseases, the community of pathogen strains can be studied in both the vertebrate host and the arthropod vector. The vertebrate immune system creates nonrandom associations between pathogen strains (1, 12) that are subsequently transmitted to the arthropod vector. Conversely, the study of mixed infections in the arthropod vector can provide information on the processes that structure the community of multiple pathogen strains in the vertebrate host (13, 14). In addition, estimates of strain richness in the arthropod vector are important for understanding the frequency with which vertebrate hosts are exposed to infections with multiple strains (13, 15). In summary, studying the diversity of pathogen strains in the arthropod vector is important for understanding the epidemiology of vector-borne diseases.
Borrelia afzelii and B. garinii are two species of tick-borne spirochete bacteria that cause Lyme borreliosis (LB) in Europe (16). These two sympatric pathogens use the same tick vector, Ixodes ricinus, but are adapted to different classes of vertebrate hosts (1719). Borrelia afzelii cycles in rodents (2025), whereas B. garinii cycles in birds (24, 2629), and this host specificity is mediated by the vertebrate complement system (30, 31). Previous work has shown that the ecological separation is not 100% complete and that double infections with these two Borrelia species do occur inside ticks (14, 3234). The tick vector, I. ricinus, has three stages, larva, nymph, and adult, that take a single blood meal to develop into the next stage. The reservoir hosts are infected by nymphal ticks, which acquired the spirochete in the previous year during the larval blood meal (vertical transmission is rare [35, 36]). In summary, B. afzelii and B. garinii occupy distinct ecological niches in the vertebrate reservoir host community, but this ecological separation is not 100% complete.
The three most-studied Borrelia species (B. burgdorferi sensu stricto, B. afzelii, and B. garinii) all contain a highly polymorphic, single-locus ospC gene that codes for outer surface protein C (OspC) (46, 47, 49, 72). The OspC protein is critical for establishing infection inside the vertebrate host (37, 38). This antigen induces a protective antibody response in the vertebrate host (3941) that is highly specific for strains carrying that particular ospC allele (4244). The ospC gene has a large amount of sequence variation that allows ospC alleles to be classified into discrete major ospC groups (45). A major ospC group is defined as a cluster of ospC alleles that is more than 8% divergent in DNA sequence from other such major groups and less than 2% divergent within the same major group (45). The absence of intermediately divergent ospC alleles (2% to 8%) suggests that cross-reactive acquired immunity has structured the community of Borrelia pathogens into discrete OspC serotypes (4547). As Borrelia strains are essentially clonal (4851), the ospC gene is a commonly used genetic marker for studying the ecology and evolution of multiple strains of Borrelia pathogens (1, 13, 45, 5255).
In the present study, we used next-generation sequencing (NGS) methods to characterize the community of genetically distinct entities characterized by a particular major ospC group (referred to here as “ospC strains”) of B. afzelii and B. garinii that were found in the same local community of I. ricinus ticks over a period of 11 years. As previous studies of the ospC polymorphism in European Borrelia species have used more traditional genotyping methods (46, 47, 49, 55), we expected that a larger sequencing effort by the NGS approach would change our understanding of the ospC gene polymorphism (52, 54). We predicted that infections with multiple ospC strains in I. ricinus nymphs would be common. We predicted that genetically similar major ospC groups would be less likely to occur in the same tick, as was previously demonstrated for B. afzelii in wild rodents (1). We predicted that our NGS approach would detect the intermediately divergent ospC alleles that must periodically appear, despite strong selection by the vertebrate immune system. Finally, we predicted that the genetic divergence between the ospC alleles of B. garinii and B. afzelii could take on intermediate divergence values (2% to 8%) because the two Borrelia species occur in different vertebrate hosts and their OspC antigens are therefore not subject to cross-immunity.
(This study is part of the Ph.D. thesis of Jonas Durand.)

MATERIALS AND METHODS

Tick collection and Borrelia detection.

Ixodes ricinus ticks were sampled monthly from the Bois de l'Hôpital site in a deciduous forest near Neuchâtel, Switzerland (47°00′55.6″N, 6°94′16.7″E; surface area of 1 ha), over a period of 11 years (2000 to 2010). We collected questing ticks by dragging a white flag (surface area of 1 m2) over the vegetation along four transects of 100 m. After collection, the species and stage of each tick were identified. In the laboratory, each tick was cut into two halves using sterile scissors. One tick half was used to determine spirochete infection via immunofluorescence microscopy (56). The other tick half was used to grow Borrelia isolates as follows. Half-ticks were placed in culture tubes containing BSK II medium (57), incubated at 34°C, and examined by dark-field microscopy every 10 days for 2 months. DNA was extracted from all spirochete-positive BSK II medium cultures. Briefly, 1 ml of culture medium was centrifuged and washed two times with phosphate-buffered saline before the pellet was suspended in 50 ml of ultrafiltered water. Pellets were incubated at 100°C for 10 min, and the thermolysates were stored at −20°C. Borrelia species were identified using a PCR-reverse line blot (RLB) assay following the method of Schouls et al. (58). The PCR protocol amplifies the 23S-5S spacer gene of B. burgdorferi sensu lato and was previously described by Alekseev et al. (59), and the RLB assay probes were described by Burri et al. (60).
As of 2012, the isolate archive at the University of Neuchâtel contained ∼3,000 tick-derived spirochete isolates for which the Borrelia species had been identified by RLB assay. In this region of Switzerland, I. ricinus ticks are commonly infected with the following Borrelia species: B. afzelii, B. bavariensis, B. burgdorferi sensu stricto, B. garinii, and B. valaisiana (14). In this study, we focused on B. afzelii and B. garinii because they are the two most common species in the isolate archive at the University of Neuchâtel and because they are the most common etiological agents of LB in Europe. We included only Borrelia isolates derived from questing nymphs because this stage takes only one blood meal during the previous larval stage. We included only those isolates that, according to the RLB assay, were infected with a single Borrelia species (either B. afzelii or B. garinii) so that we could assign major ospC groups with confidence to the correct Borrelia species. We established a database of all the nymph-derived spirochete isolates of either B. afzelii or B. garinii obtained from 2000 to 2010. For each of the two Borrelia species, we randomly sampled a maximum of 20 nymph-derived isolates for each of the 11 years. Using this sampling strategy, we obtained 193 isolates of B. afzelii and 190 isolates of B. garinii (Table 1).
TABLE 1
TABLE 1 Number of nymph-derived spirochete isolates used for each of the 22 combinations of Borrelia species and yeara
Yr No. of nymph-derived isolates tested (total no.)
B. afzelii B. garinii
Infected nymphs Uninfected nymphs Infected nymphs Uninfected nymphs
2000 20 (51)b 234 (599)c 20 (41) 297 (609)
2001 17 (37)d 295 (643) 18 (37) 312 (643)
2002 16 (69) 153 (661) 19 (39) 336 (691)
2003 12 (21) 261 (458) 14 (19) 338 (460)
2004 13 (21) 298 (448) 14 (16) 396 (453)
2005 19 (88) 104 (459) 19 (29) 339 (518)
2006 20 (57) 171 (490) 15 (33) 249 (514)
2007 20 (41) 275 (564) 20 (62) 175 (543)
2008 18 (31) 347 (598) 20 (54) 212 (575)
2009 20 (51) 211 (540) 13 (14) 535 (577)
2010 18 (58) 156 (503) 18 (32) 297 (529)
a
We randomly selected a maximum of 20 nymph-derived isolates for NGS (the total number of infected nymphs is shown in parentheses) for each combination of Borrelia species and year. To calculate the absolute frequency of each strain (F3 in Tables 2 and 3), we used a subsample of uninfected nymphs (the total number of uninfected nymphs is shown in parentheses) for each combination of Borrelia species and year. This subsample of uninfected nymphs is proportional to the subsample of infected nymphs.
b
In the year 2000, 51 nymph-derived cultures tested positive for B. afzelii. We randomly selected 20 of these 51 isolates of B. afzelii and determined the community of ospC strains using NGS.
c
In the year 2000, 599 nymph-derived cultures did not test positive for B. afzelii. We calculated the absolute frequencies of the B. afzelii ospC strains (F3 in Tables 2 and 3), using a subsample of 234 cultures [(20/51) × 599 = 234] not infected with B. afzelii.
d
In the year 2001, 37 nymph-derived cultures tested positive for B. afzelii, but only 17 of these 37 isolates of B. afzelii had sufficient material for NGS.

Deep sequencing of the ospC strain community in nymph-derived spirochete isolates.

The ospC gene was amplified using a nested PCR method consisting of two consecutive amplification steps that was described by Bunikis et al. (49). We modified the oligonucleotides for the second PCR to include 454 Life Science's A or B sequencing adapter, the key, and the template-specific primers (see the supplemental material). The forward primer was tagged for forward sequencing. The reaction was performed in an automatic thermocycler (Eppendorf, USA). We checked the success of the PCRs by visualizing the amplicons on a 1% agarose gel stained with Midori green (Labgene, Switzerland). The PCR products were purified using a DNA purification kit (Wizard SV Gel and PCR cleanup system; Promega, Switzerland). We created eight different pools of Borrelia isolates for the 454 sequencing run (40 to 50 isolates per pool). Isolates were randomly assigned to pools subject to the constraint that each pool contained a balanced number of isolates from each year. Sequencing was done on a 454 Roche GS FLX sequencing apparatus and was outsourced to Microsynth AG (Balgach, Switzerland).

Bioinformatics. (i) Cleaning of ospC gene sequences.

The 454 sequencing run produced 352,882 sequences of the ospC gene. After cleaning the data set (see the supplemental material), we retained 240,410 ospC gene sequences that were each 521 bp long.

(ii) Clustering analysis of major ospC groups.

We defined an a priori set of seed sequences prior to conducting the cluster analysis. Seed sequences were defined as those sequences that occurred at least 50 times in our cleaned data set. The seed sequences were selected using the cd-hit-dup function of CD-HIT software (61, 62) with an error rate of 1%, which produced a set of 331 seed sequences. These sequences were used as seeds to perform a cluster analysis of the data from the cleaned data set with similarity thresholds ranging from 98% to 91% using the software CD-HIT. In addition, we created a phylogenetic tree (1,000 bootstraps) of the 331 seed sequences using the PhyML program (63). This tree confirmed the reliability of the clustering of the seeds. To have a measure of genetic similarity of the major ospC groups, we calculated the mean pairwise genetic distances using the distmat program (64) with a Kimura model. Each mean pairwise genetic distance was based on 30 sequences of each major ospC group in the pair. We also calculated the focal pairwise genetic distance for each major ospC group. This coefficient measures the average genetic distance between a focal major ospC group and all other major ospC groups.

(iii) Nature of genetic variation within major ospC groups.

The nature of the genetic variation observed within each major ospC group was investigated. Only the first 350 nucleotides of the ospC gene sequences were analyzed because the end of the forward sequences contained too much sequencing error. For each major ospC group, 30 sequences were selected from the seed sequences that had been used for the clustering (a total of 690 sequences). For major ospC groups with less than 30 seed sequences, nonseed sequences (sequences that occurred less than 50 times in our data set) were added. Sequences were aligned manually, and the numbers of synonymous and nonsynonymous substitutions, insertions, and deletions were recorded.

Nomenclature of major ospC groups.

Several papers have dealt with the nomenclature of the major ospC groups in B. burgdorferi sensu lato (46, 47, 49, 54, 65, 66). In the United States, workers have developed a well-established system of names for the major ospC groups of B. burgdorferi sensu stricto (45). In Europe, in contrast, there is no universally adopted naming system for the major ospC groups of B. afzelii and B. garinii (47, 49). In the current study, we present a summary of all the different names that have been used depending on the author (see Tables S1 and S2 in the supplemental material). In addition, we suggest that future work on the major ospC groups of these two Borrelia species follow the nomenclature developed by Bunikis et al. (49), Hellgren et al. (50), and Strandh and Råberg (54) for B. afzelii and that developed by Lagal et al. (47) for B. garinii. Here, we use “ospC strain” to refer to a genetically distinct entity characterized by a particular major ospC group.

Confirmation of Borrelia species classification by RLB assay.

To confirm the reliability of the RLB assay, we performed additional sequencing of a subset of 120 isolates. According to the RLB assay, 60 of these isolates were singly infected with B. afzelii and 60 were singly infected with B. garinii. For each of these 120 samples, we used PCR to amplify the recA gene and the hbb gene according to the protocol described by Richter et al. (67). The amplicons were purified using a Qiagen purification kit. Purified amplicons were sent to Microsynth AG (Balgach, Switzerland) for Sanger sequencing.

Statistical methods.

All statistical analyses were performed using R (68).

(i) Prevalence of ospC strains inside ticks.

The absolute prevalence of a pathogen strain is its frequency in the entire population (uninfected and infected individuals). The relative prevalence of a pathogen strain is its frequency in the subpopulation of infected individuals. In the present study, the absolute prevalence of infection was determined for each Borrelia species by the RLB assay. We then used 454 sequencing to estimate the relative prevalence of each major ospC group in a randomly selected subset of Borrelia isolates. Some analyses of multiple-strain pathogen communities require estimates of the absolute prevalence. We therefore converted our relative prevalences to absolute prevalences by including the fraction of uninfected ticks that corresponded to the fraction of infected ticks that had been processed by 454 sequencing for each of the 22 combinations of Borrelia species and year (Table 1). For example, if we subsampled half of the infected ticks for a particular combination of Borrelia species and year, then we included half of the uninfected ticks for that combination of species and year. In this way, our data set was a random and representative sample of the populations of infected and uninfected ticks (Table 1). All analyses described below were conducted separately for each Borrelia species.

(ii) Aggregation of ospC strains inside ticks.

Strain richness refers to the number of different ospC strains that were found within a single tick (including uninfected ticks). To test whether strains were aggregated inside ticks, we compared the observed distribution of strain richness to the Poisson distribution, where the variance is equal to the mean (69). Aggregation occurs when the number of ticks with multiple strains is greater than the Poisson expectation (variance/mean > 1). We used a bootstrap analysis to calculate the 95% confidence limits of the variance-to-mean ratio.

(iii) Association of ospC strains inside ticks.

The strength and the direction of each pairwise association were estimated using the association coefficient described by Pielou (70). We used the absolute frequencies to calculate this pairwise association coefficient, which has a range from −1 (a complete negative association) to +1 (a complete positive association) and is undefined if all hosts are uninfected or infected with both strains. For each pair of strains, the statistical significance of the pairwise association was assessed using a two-tailed Fisher exact test. The significance level was adjusted for multiple comparisons using the Bonferroni correction (for B. afzelii, P = 0.05/171 pairs = 0.000292; for B. garinii, P = 0.05/153 pairs = 0.000327).
We also calculated the focal pairwise association coefficient for each major ospC group. This coefficient is a measure of whether a focal major ospC group is more or less likely to co-occur with other ospC strains. Strains with high focal pairwise association coefficients (+1) are more likely to be found in multiple infections than strains with low focal pairwise association coefficients (−1).
Finally, we tested whether strains with genetically dissimilar major ospC groups were more likely to occur together inside the same tick. Specifically, we tested for a correlation between the pairwise associations and the pairwise genetic distances using the Mantel test in the vegan package in R (71).

Nucleotide sequence accession numbers.

The ospC gene sequence data have been deposited in the Sequence Read Archive under BioProject PRJNA293785 with the accession number SRP063760. A type sequence of each major ospC group of B. afzelii and B. garinii is available in GenBank (see the supplemental material for the GenBank accession numbers). The GenBank accession numbers for major ospC groups V1 of B. valaisiana and Q of B. burgdorferi sensu stricto are AF093483 and JQ253799, respectively.

RESULTS

Clustering analysis.

For each of the two Borrelia species, we randomly sampled a maximum of 20 nymph-derived isolates for each of the 11 years. Using this sampling strategy, we obtained 193 isolates of B. afzelii and 190 isolates of B. garinii (Table 1). The 454 sequencing run produced 352,882 sequences of the ospC gene. After cleaning the data set (see the supplemental material), we retained 240,410 ospC gene sequences that were each 521 bp long. There were 23 different major ospC groups in our local population of I. ricinus ticks, and this number was stable across a range of similarity thresholds (93% to 98%; see Fig. S1 in the supplemental material). According to BLAST analysis of the 23 major ospC groups, 10 belonged to B. afzelii (ospC groups A1, A2, A3, A5, A7, A9, A10, A11, A12, and A14), 11 belonged to B. garinii (ospC groups G2, G4, G6, G7, G8, G9, G10, G11, G13, G14, and G15), 1 belonged to B. valaisiana (ospC group V1), and 1 belonged to B. burgdorferi sensu stricto (ospC group Q). All of the 23 major ospC groups had been previously described in the literature (1, 46, 47, 72).
There were a number of major ospC groups that were found in both Borrelia species. Five of the 10 B. afzelii major ospC groups (A1, A9, A10, A11, A14) were found in isolates that were singly infected with B. garinii according to the RLB assay. Conversely, 7 of the 11 B. garinii major ospC groups (G2, G7, G8, G9, G11, G13, G14) were found in isolates that were singly infected with B. afzelii according to the RLB assay. In what follows, we use the terms “native” and “exotic” to distinguish between these two types of major ospC groups. Thus, B. afzelii had 10 native and 9 exotic ospC groups (7 B. garinii-derived ospC groups and the V1 and Q ospC groups) for a total of 19 major ospC groups (Table 2), and B. garinii had 11 native and 7 exotic ospC groups (5 B. afzelii-derived ospC groups and the V1 and Q ospC groups) for a total of 18 major ospC groups (Table 3).
TABLE 2
TABLE 2 Frequencies of major ospC groups in the B. afzelii isolates
ospC allele ospC status Frequency (%)
F1a F2b F3c
A10 Natived 43.35 (49,601) 54.40 (105) 3.90 (105)
A9 Natived 14.31 (16,372) 31.09 (60) 2.23 (60)
A1 Natived 12.28 (14,056) 24.35 (47) 1.75 (47)
A14 Natived 11.02 (12,613) 24.87 (48) 1.78 (48)
A11 Natived 2.31 (2,642) 10.36 (20) 0.74 (20)
A12 Native 4.61 (5,270) 10.36 (20) 0.74 (20)
A3 Native 2.69 (3,080) 4.66 (9) 0.33 (9)
A5 Native 2.90 (3,318) 7.77 (15) 0.56 (15)
A2 Native 2.64 (3,023) 8.81 (17) 0.63 (17)
A7 Native 2.02 (2,307) 7.25 (14) 0.52 (14)
G9 Exotice 0.49 (560) 7.77 (15) 0.56 (15)
G11 Exotice 0.39 (449) 4.15 (8) 0.30 (8)
Q Exotic 0.36 (409) 1.04 (2) 0.07 (2)
V1 Exotic 0.27 (314) 0.52 (1) 0.04 (1)
G13 Exotice 0.26 (302) 10.88 (21) 0.78 (21)
G8 Exotice 0.04 (45) 8.81 (17) 0.63 (17)
G14 Exotice 0.02 (27) 7.77 (15) 0.56 (15)
G2 Exotice 0.02 (24) 6.74 (13) 0.48 (13)
G7 Exotice 0.02 (20) 5.70 (11) 0.41 (11)
a
Frequency 1 (F1) refers to the frequency of the major ospC group in the sample of sequences (n = 114,432 sequences), and values in parentheses are numbers of sequences.
b
Frequency 2 (F2) refers to the frequency of the major ospC group in the sample of infected nymphs (n = 193 nymphs), and values in parentheses are numbers of infected nymphs.
c
Frequency 3 (F3) refers to the frequency of the major ospC group in the sample of infected and uninfected nymphs (n = 2,500 nymphs), and values in parentheses are numbers of infected nymphs.
d
These native major ospC groups were horizontally transferred from B. afzelii to B. garinii.
e
These exotic major ospC groups were horizontally transferred from B. garinii to B. afzelii.
TABLE 3
TABLE 3 Frequencies of major ospC groups in the B. garinii isolates
ospC allele ospC status Frequency (%)
F1a F2b F3c
G8 Natived 25.19 (31,737) 42.11 (80) 2.18 (80)
G14 Natived 16.85 (21,223) 30.00 (57) 1.55 (57)
G9 Natived 11.71 (14,748) 22.63 (43) 1.17 (43)
G13 Natived 10.34 (13,030) 27.89 (53) 1.44 (53)
G2 Natived 10.68 (13,457) 21.05 (40) 1.09 (40)
G7 Natived 9.57 (12,057) 31.05 (59) 1.61 (59)
G11 Natived 4.28 (5,389) 14.74 (28) 0.76 (28)
G4 Native 3.99 (5,025) 14.74 (28) 0.76 (28)
G6 Native 3.14 (3,952) 11.58 (22) 0.60 (22)
G15 Native 0.61 (774) 3.68 (7) 0.19 (7)
G10 Native 0.55 (689) 1.58 (3) 0.08 (3)
A10 Exotice 1.75 (2,209) 31.58 (60) 1.63 (60)
A1 Exotice 0.74 (930) 12.11 (23) 0.63 (23)
V1 Exotic 0.41 (512) 1.58 (3) 0.08 (3)
Q Exotic 0.10 (128) 2.63 (5) 0.14 (5)
A9 Exotice 0.04 (50) 11.05 (21) 0.57 (21)
A14 Exotice 0.03 (36) 10.53 (20) 0.54 (20)
A11 Exotice 0.03 (32) 5.26 (10) 0.27 (10)
a
Frequency 1 (F1) refers to the frequency of the major ospC group in the sample of sequences (n = 125,978 sequences), and values in parentheses are numbers of sequences.
b
Frequency 2 (F2) refers to the frequency of the major ospC group in the sample of infected nymphs (n = 190 nymphs), and values in parentheses are numbers of infected nymphs.
c
Frequency 3 (F3) refers to the frequency of the major ospC group in the sample of infected and uninfected nymphs (n = 3,486 nymphs), and values in parentheses are numbers of infected nymphs.
d
These native major ospC groups were horizontally transferred from B. garinii to B. afzelii.
e
These exotic major ospC groups were horizontally transferred from B. afzelii to B. garinii.

Description of the Borrelia ospC strain communities in I. ricinus.

The major ospC groups were more evenly distributed in B. garinii than B. afzelii. In B. afzelii, the strain of ospC group A10 (referred to here as strain A10) was found in 54.40% of the infected nymphs (Table 2). Three strains (A1, A9, A14) were found in 24.35% to 31.09% of the infected nymphs, and the remaining 15 strains were found in less than 11% of the infected nymphs (Table 2). In B. garinii, strain G8 was found in 42.11% of the infected nymphs, and six other strains were found in 21.05% to 31.58% of the infected nymphs (Table 3). Six strains were found in 10.53% to 14.74% of the infected nymphs, whereas the five remaining strains were found in less than 6% of the infected nymphs. In B. garinii, the exotic major ospC group A10 was very common and occurred in 31.58% of the infected nymphs.

Pairwise genetic distances within the native major ospC groups.

For B. afzelii, the mean pairwise genetic distance within each of the 10 native major ospC groups was 1.06% (n = 44,250 pairwise comparisons, range = 0.42% to 2.16%). For B. garinii, the mean pairwise genetic distance within each of the 11 native major ospC groups was 1.14% (n = 54,285 pairwise comparisons, range = 0.66% to 1.63%).

Nature of genetic variation within major ospC groups.

Across the 690 sequences belonging to the 23 major ospC groups, there were 140 nucleotide substitutions, of which 58.9% were synonymous (83/141) and 41.1% were nonsynonymous (58/141). Single nucleotide insertions and deletions and codon deletions were common, whereas double nucleotide insertions or deletions were rare (see Table S3 in the supplemental material).

Pairwise genetic distances between the native major ospC groups.

We calculated the pairwise genetic distances between the 21 native major ospC groups (i.e., the exotic major ospC groups were excluded). The conspecific or allospecific pairwise genetic distance refers to whether the two native major ospC groups belonged to the same Borrelia species or to different Borrelia species, respectively. The mean conspecific pairwise genetic distances of B. afzelii (n = 45 pairwise comparisons, mean = 17.84%, range = 8.86% to 23.59%; Fig. 1) and B. garinii (n = 55 pairwise comparisons, mean = 18.54%, range = 12.77% to 26.36%; Fig. 1) were not significantly different (t = −1.13, degrees of freedom [df] = 98, P = 0.261). The mean allospecific pairwise genetic distance (n = 110 comparisons, mean = 20.94%, range = 10.71% to 25.93%; Fig. 1) was significantly larger than the mean conspecific pairwise genetic distance (t = 6.921, df = 208, P < 0.001).
FIG 1
FIG 1 Distribution of pairwise genetic distances between major ospC groups. The genetic distances for three different pairwise comparisons are shown: conspecific B. afzelii (af; n = 55), conspecific B. garinii (ga; n = 55), and allospecific (B. afzelii versus B. garinii; n = 121). Shown are the medians (heavy black lines), the 25th and 75th percentiles (edges of the boxes), the minimum and maximum values (whiskers), and the outliers (circles).

Strain richness in nymphal ticks.

The strain richness of the infected nymphal ticks was calculated using both the native and exotic major ospC groups for each Borrelia species. Most nymphal ticks were infected with multiple ospC strains. For B. afzelii (19 major ospC groups), 78.76% (152/193) of the infected nymphal ticks carried multiple ospC strains. For B. garinii (18 major ospC groups), 84.74% (161/190) of the infected nymphal ticks carried multiple ospC strains. The mean strain richness of B. garinii (2.96 strains per infected tick, range = 1 to 11 strains per tick) was 25% higher than that of B. afzelii (2.37 strains per infected tick, range = 1 to 7 strains per tick), and this difference was statistically significant (t = −3.764, df = 381, P < 0.001).

Aggregation of strains in nymphal ticks.

The variance-to-mean ratio of strain richness was calculated using the absolute frequencies of both the native and exotic major ospC groups for each Borrelia species. The variance-to-mean ratio of strain richness was significantly greater than 1.0 for both Borrelia species, indicating that the ospC strains were highly aggregated in the nymphal ticks (Fig. 2). The variance-to-mean ratio of strain richness was 47.4% higher in B. garinii (mean = 3.98, 95% confidence interval [CI] = 3.44 to 4.51) than in B. afzelii (mean = 2.70, 95% CI = 2.50 to 2.91).
FIG 2
FIG 2 Distribution of strain richness in nymphs infected with B. afzelii (A) and nymphs infected with B. garinii (B). The numbers of ticks infected with zero strains are not shown for scaling purposes (i.e., uninfected nymphs in Table 1).

Pairwise association index.

The pairwise association coefficients within each Borrelia species were calculated using the absolute frequencies of both the native and exotic major ospC groups. For B. afzelii (19 major ospC groups), 66.7% (114/171) of the pairwise association coefficients were positive (mean = 0.08, range = −0.01 to 0.35). All of the 39 statistically significant (P < 0.05/171) pairwise associations were positive (mean = 0.20, range = 0.11 to 0.35). For B. garinii (18 major ospC groups), 77.8% (119/153) of the pairwise association coefficients were positive (mean = 0.15, range = −0.004 to 0.48). Again, all of the 78 statistically significant (P < 0.05/153) pairwise associations were positive (mean = 0.25, range = 0.11 to 0.48). The mean pairwise association coefficient of B. garinii was significantly greater than that of B. afzelii (t = 6.10, df = 322, P < 0.001). The results were the same for the subset of 117 statistically significant pairwise association coefficients (t = 3.54, df = 115, P < 0.001).
The mean focal pairwise association index of B. garinii (n = 18 major ospC groups, mean = 0.35, range = 0.12 to 0.63) was 20.7% higher than that of B. afzelii (n = 19 major ospC groups, mean = 0.29, range = −0.005 to 0.62), but this difference was not significant (t = 1.35, df = 35, P = 0.187). Some ospC strains, such as B. afzelii ospC strain A10 and B. garinii ospC strains G8, G14, and A10, had very high focal pairwise association coefficients (>0.50).

Relationship between relative frequency and the focal pairwise association index.

The relative frequency (F2 in Tables 2 and 3) of each major ospC group (transformed by the square root) was a highly significant predictor of the focal pairwise association index in B. afzelii (n = 19 major ospC groups, F1, 17 = 565.7, P < 0.001, r2 = 0.971; Fig. 3) and in B. garinii (n = 18 major ospC groups, F1, 16 = 1,688.0, P < 0.001, r2 = 0.991; Fig. 3). An analysis of covariance found no significant interaction between Borrelia species and relative frequency on the pairwise association index (F1, 33 = 0.9, P = 0.354). The slope of the relationship between the focal pairwise association index and the square root-transformed relative frequency was therefore the same for each Borrelia species (0.90 ± 0.044). After controlling for the relative frequency, the focal pairwise association index remained significantly higher for B. garinii than B. afzelii (F1, 34 = 4.5, P = 0.041).
FIG 3
FIG 3 Positive relationship between the focal pairwise association index and the square root-transformed relative frequency of each major ospC group inside the tick. Results are combined for B. afzelii (circles) and B. garinii (triangles). The native major ospC groups that were horizontally transferred to the other Borrelia species (white symbols; n = 12) had much higher relative frequencies and pairwise association indices in the native Borrelia species than the native major ospC groups that were not transferred (gray symbols; n = 9). The exotic major ospC groups (n = 16) are shown in black symbols.

Relationship between pairwise genetic distance and pairwise association index.

The correlation between the pairwise genetic distance between the major ospC groups and the pairwise association index of the major ospC groups inside the ticks was not statistically significant for either B. afzelii (n = 171 pairwise elements; Mantel test, r = −0.041, P = 0.630; Fig. 4) or B. garinii (n = 153 pairwise elements; Mantel test, r = −0.053, P = 0.664; Fig. 4). Similarly, the correlation between the focal pairwise genetic distance and the focal pairwise association index was not statistically significant for either B. afzelii (n = 19 major ospC groups; Pearson correlation test, r = 0.032, t = 0.13, df = 17, P = 0.897; Fig. 5) or B. garinii (n = 18 major ospC groups; Pearson correlation test, r = −0.175, t = −0.71, df = 16, P = 0.485; Fig. 5).
FIG 4
FIG 4 Relationship between the pairwise association index of the major ospC groups inside the tick versus the pairwise genetic distance. The relationship is shown for B. afzelii (A) and B. garinii (B). The regression was done on the subset of the statistically significant pairwise associations indices but was not statistically significant for either Borrelia species.
FIG 5
FIG 5 Relationship between the focal pairwise association index and the focal pairwise genetic distance of each major ospC group inside the tick. The relationship is shown for B. afzelii (A) and B. garinii (B). All data points are labeled with the name of the major ospC group. White and gray symbols indicate native major ospC groups that had and that had not been transferred to the other Borrelia species, respectively; black symbols indicate exotic major ospC groups.

Exotic major ospC groups were widespread in both Borrelia species.

For the B. afzelii-infected nymphs (as determined by the RLB assay), 42.0% (81/193) of the infections contained exotic major ospC groups. Conversely, for the B. garinii-infected nymphs (as determined by the RLB assay), 55.3% (105/190) of the infections contained exotic major ospC groups. In B. afzelii, 1.88% of the ospC gene sequences (2,150/114,432) were of exotic origin. In B. garinii, 3.09% of the sequences (3,897/125,978) were of exotic origin.

Exotic major ospC groups were common in their native Borrelia species.

The five native B. afzelii major ospC groups that had undergone horizontal transfer (to B. garinii) accounted for 84.62% of the native ospC gene sequences of B. afzelii (95,598/112,596). Similarly, the seven native B. garinii major ospC groups that had undergone horizontal transfer (to B. afzelii) accounted for 91.45% of the native ospC gene sequences of B. garinii (111,641/122,081). For the two Borrelia species combined, we compared the mean relative frequency between native major ospC groups that had experienced horizontal transfer or not (Tables 2 and 3). The 12 native major ospC groups that had experienced horizontal transfer had a significantly higher relative frequency (t = 4.89, df = 19, P < 0.001; Fig. 3) than the 9 native major ospC groups that had not been transferred. For the 12 major ospC groups that had experienced horizontal transfer, the relative frequency in the native species was always higher than the frequency in the exotic species (paired t test, t = 8.13, df = 11, P < 0.001). There was a significant correlation between the relative frequencies of these 12 major ospC groups in their native species and in their exotic species (Pearson correlation test, r = 0.783, t = 3.99, df = 10, P = 0.002).

Sanger sequencing of Borrelia genes to confirm Borrelia species classification by RLB assay.

For a subsample of 120 Borrelia isolates, we obtained high-quality sequences for both the recA gene and the hbb gene for 110 isolates. This genetic analysis confirmed the Borrelia species classification of the RLB assay in 95.5% (105/110) of our samples. In general, this additional analysis showed that the RLB assay provided a reliable classification of the Borrelia species.

DISCUSSION

Cross-immunity and co-occurrence of Borrelia ospC strains.

Cross-reactive acquired immunity should reduce the co-occurrence of antigenically similar strains in the same reservoir host (5, 73, 74). Assuming that the genetic similarity between major ospC groups predicts the probability of cross-reactive acquired immunity between their respective OspC antigens, we expected a positive relationship between the pairwise genetic distance and the pairwise association index. A study of multiple B. afzelii infections in a population of wild voles found that dissimilar major ospC groups were more likely to be found together in the same host than similar major ospC groups (1). In contrast, we found no relationship between the pairwise genetic distance and the pairwise association index inside the ticks. A simple explanation for the difference between these two studies is that one looked at the reservoir host, whereas the other looked at the tick vector. The vertebrate immune system structures the community of ospC strains inside the reservoir hosts, and the tick is expected to acquire a subsample of this strain community (13, 75). However, the community structure of the colonizing strains might be modified by subsequent events inside the tick. The spirochete load can change dramatically following the blood meal and during the larva-to-nymph molt, and these changes in abundance may alter the community of strains inside the tick (14, 76). There is currently very little information as to how Borrelia strains interact inside the tick over time (75).

Strong selection against ospC alleles with intermediate divergence.

Despite a large sequencing effort (240,410 sequences from 389 ticks over a period of 11 years), our study did not find any ospC alleles that were intermediately divergent (2% to 8%) from the community of major ospC groups. The average pairwise genetic distance between conspecific major ospC groups was always greater than 8%. Thus, our study confirmed the highly discontinuous pattern of genetic variation at the ospC locus, as was reported in previous studies of B. burgdorferi sensu lato (4547, 49). The original studies demonstrating the 8% divergence cutoff between the major ospC groups in B. afzelii and B. garinii were based on a total of 140 sequences (46, 47, 49). Thus, it was reasonable to expect that a larger sequencing effort would discover the missing intermediate major ospC groups. The present study suggests that strains carrying ospC alleles that are 2% to 8% divergent rarely coexist in the same Borrelia species. Interestingly, our results also suggest that such intermediately divergent ospC alleles rarely coexist in Borrelia species occupying different niches in the vertebrate host community. The allospecific pairwise genetic distances between B. afzelii and B. garinii had the same discontinuous pattern, suggesting that the evolution of the ospC allele is not independent in each Borrelia species. One explanation is that frequent horizontal transfer of the ospC gene (see below) essentially synchronizes the community of major ospC groups between these two Borrelia species. Thus, a new ospC mutant can rise to a detectable frequency only if it is at least 8% divergent from all the major ospC groups of all the Borrelia pathogens in the local community.
Purifying selection by cross-reactive antibodies is the most plausible explanation for the missing intermediate major ospC groups. Mutant major ospC groups that are genetically similar and thus antigenically similar to their conspecific counterparts are vulnerable to cross-reactive antibodies, which prevent these mutants from invading the local population (11, 74, 77). Indirect evidence for this hypothesis comes from studies of B. burgdorferi sensu stricto that have shown that antibodies against a given OspC antigen provide protection against strains belonging to the same major ospC group but not against strains belonging to a different major ospC group (42, 44). Thus, in B. burgdorferi sensu stricto, we see the end result of divergent selection on the ospC antigen to escape from the cross-reactive antibodies directed against its neighbors. However, immune escape between strains carrying different major ospC group alleles is not always perfect. A recent study of B. afzelii found that OspC antigens belonging to different major ospC groups (that were 20.68% divergent in DNA sequence) could still interfere with one another via the cross-reactive antibodies of the rodent host (43). In summary, cross-reactive antibodies are the most plausible source of purifying selection that removes intermediately divergent major ospC groups from the local population of Borrelia pathogens.
Each major ospC group contains a cluster of ospC alleles that are less than 2% divergent from each other with respect to their DNA sequence. An earlier study of the ospC polymorphism found that there was very little sequence variation within the major ospC groups (9 of the 11 major ospC groups were isosequential) and concluded that the effective population size of each group was very small (45). The present study found that strain richness within the major ospC groups was high (331 unique seed sequences clustered into 23 major ospC groups, giving a mean richness of 14.4 seed sequences per group). Examination of the sequence variation within the major ospC groups found that the majority (∼60%) of the nucleotide substitutions were synonymous. Sequencing error was the likely cause of all insertions and deletions that resulted in a frameshift, as the downstream parts of the OspC protein were highly conserved. The present study suggests that most of the sequence variation observed within a major ospC group does not affect the amino acid sequence and/or the function of the OspC protein. Under this scenario, natural selection cannot discriminate among the different allelic variants, and the frequencies of these silent alleles depend on genetic drift alone (78). Thus, the observation that the major ospC groups contain allelic variants does not contradict the hypothesis that cross-immunity selects against insufficiently divergent ospC alleles.

Richness of the major ospC groups.

We found that the B. afzelii and B. garinii isolates in our local area contained 10 and 11 native major ospC groups, respectively (strain richness increased to 19 and 18, respectively, after the exotic major ospC groups were included). These findings were similar to those of previous studies of the ospC polymorphism in local tick populations (45, 53, 55, 79). In the United States, local populations of B. burgdorferi sensu stricto have between 13 and 16 different major ospC groups (13, 45, 53, 79). In Europe, local populations of B. afzelii have 10 different major ospC groups (55). Why do populations of B. burgdorferi sensu lato contain 12 and not 120 different major ospC groups? The relatively small length of the ospC gene (∼600 bp) in combination with the constraints of remaining functional and of avoiding cross-reactive acquired immunity with all of its neighbors must set an upper limit to the number of major ospC groups that can exist in a single, local population.

Aggregation of Borrelia ospC strains in the tick vector.

Ixodes ricinus nymphs infected with B. afzelii or B. garinii contained a mean richness of 2.37 and 2.97 different ospC strains per nymph, respectively. Our estimates of the mean ospC strain richness in questing I. ricinus nymphs were considerably higher than those from previous studies (1.07 to 1.44 strains per nymph) of B. burgdorferi sensu lato infections of questing nymphs in I. scapularis, I. pacificus, and I. ricinus (53, 55, 80). However, our estimates of the mean ospC strain richness are similar to those from previous studies (1.0 to 3.7 strains per tick) that sampled larval ticks feeding on infected hosts (13, 55, 81, 82) or questing adults (45, 53). We found a very high prevalence of multiple ospC strains in both B. afzelii (78.76%) and B. garinii (84.74%). Our estimates of the prevalence of multiple ospC strain infections were much higher than those in previous studies (0.0% to 50.0%) (1, 45, 55, 8185). These earlier studies estimated the mean ospC strain richness and the prevalence of multiple strain infections by using molecular methods, such as the RLB assay, cold single-strand conformation polymorphism (SSCP) analysis, and Sanger sequencing. A previous study showed that the RLB assay was more sensitive at detecting infections with multiple strains than SSCP analysis (45, 53). Thus, differences in strain richness among studies are partially confounded by differences in the sensitivity of the molecular methods. In the present study, we used next-generation sequencing (NGS) to analyze an average of 618 ospC gene sequences per tick (240,410 sequences, 389 ticks). This high level of coverage for each tick allowed us to detect major ospC groups that represented a tiny fraction of that particular amplicon (<0.2%). Thus, NGS methods estimate higher levels of strain diversity than the older molecular methods because these approaches are more sensitive at detecting rare strains.
The ospC strains were highly aggregated in I. ricinus ticks. There are a number of explanations for this pattern of aggregation. First, vertebrate hosts vary in their competence in acquiring infections and transmitting spirochetes to vector ticks (18, 30, 86, 87). If a substantial proportion of ticks feed on incompetent hosts and the remaining ticks feed on infected reservoir hosts, the Borrelia strains will appear to be aggregated inside the ticks when these two subpopulations of ticks are analyzed together. Second, the aggregation of ticks on reservoir hosts can also cause the aggregation of Borrelia strains in the tick vector. Numerous studies have observed that a minority of wild rodent hosts feeds a majority of the I. ricinus ticks (88, 89). There are a variety of ecological factors that cause aggregation of ticks in wild rodents (90). Third, the immune system of the vertebrate host might have more difficulty controlling and/or clearing infections with multiple strains than infections with a single strain (74, 91). Under this hypothesis, infections with multiple strains have a higher probability of establishing infection in the reservoir host and/or have a higher host-to-tick transmission success than infections with a single strain.

Differences between B. garinii and B. afzelii.

The mean ospC strain richness of B. garinii (2.96 strains per nymph) was 25% higher than that of B. afzelii (2.37 strains per nymph). Borrelia garinii ospC strains were more often found in multiple infections and had higher pairwise association indices than their B. afzelii counterparts. The higher mean ospC strain richness for B. garinii was not caused by differences in total strain richness, which was the same between B. afzelii (n = 19 ospC strains) and B. garinii (n = 18 ospC strains). Borrelia garinii is mostly found in avian hosts, whereas B. afzelii is mostly found in rodent hosts (16, 2022, 24, 25). Thus, one explanation is that the richness of B. garinii ospC strains in birds is higher than the richness of B. afzelii ospC strains in rodents. To date, most of the studies investigating ospC strain richness in vertebrate reservoir hosts have focused on mammals (1, 13, 50, 52, 55, 66, 82, 83). The two studies that investigated birds (81, 85) found a mean ospC strain richness (1.00 to 1.76 strains per tick) that was low compared to that found in the studies with mammals, but this could be an artifact of the relatively inefficient molecular methods that were used in the bird studies (RLB assay and Sanger sequencing). A recent study characterized the spirochete loads of different Borrelia species inside questing I. ricinus nymphs (14). This study found that the median spirochete load of B. garinii (5,080 spirochetes per nymph) was 1.6 times higher than that of B. afzelii (3,140 spirochetes per nymph) (14). Thus, another explanation for the discrepancy in strain richness between the two species is that the tick vector has a higher carrying capacity for B. garinii than for B. afzelii.

Presence of exotic major ospC group alleles in nymphs that, according to the RLB assay, were infected with a single Borrelia species.

In the present study, we used only nymph-derived isolates that, according to the RLB assay (which targeted the 23S-5S spacer gene), were singly infected with either B. afzelii or B. garinii, and we therefore excluded those isolates that, according to the RLB assay, were coinfected with multiple Borrelia species. A surprising result was therefore that 42% of the B. afzelii-infected ticks and 55% of the B. garinii-infected ticks contained exotic major ospC groups from other Borrelia species. While the exotic major ospC groups were widespread, they were not abundant. For B. afzelii and B. garinii, the exotic major ospC groups accounted for 1.88% and 3.09% of all the ospC sequences, respectively. Two alternative explanations for these results are the coinfection explanation and the horizontal gene transfer explanation.

Coinfection explanation for the presence of exotic major ospC group alleles.

In the coinfection explanation, the presence of native and exotic ospC groups in the same isolate represents a true coinfection with B. afzelii and B. garinii that was not detected by the RLB assay. In this scenario, the RLB assay detected the highly abundant Borrelia species (corresponding to the abundant native ospC alleles from the 454 sequencing) but not the Borrelia species of low abundance (corresponding to the low-abundance exotic ospC alleles from the 454 sequencing). Although we have recently shown that the sensitivity of our RLB assay is similar to that of a widely used quantitative PCR assay for detecting Borrelia infections in ticks (14, 92), we cannot exclude this possibility. Previous studies have shown that nymphs and adult ticks carry coinfections with B. afzelii and B. garinii (14, 3234). In our local I. ricinus population, the RLB assay found that 0.75% (13/1,731) of the nymphs are coinfected with B. afzelii and B. garinii (14). If the coinfection explanation is true, the implication is that the true frequency of coinfections with B. afzelii and B. garinii has been underestimated by a factor of 56 to 73 (42% to 55% in the present study versus 0.75% in the study by Herrmann et al. [14]). Interestingly, we recently showed that the spirochete load of B. afzelii and B. garinii coinfections was much lower than the additive expectation (14). We suggested that the complement system of the vertebrate host reduced the spirochete load of the maladapted Borrelia species inside the nymphal tick (14). Thus, the observation that the exotic ospC groups had low abundances (1.88% and 3.09%) is consistent with our understanding of how the vertebrate complement system would interact with coinfections containing bird- and rodent-adapted Borrelia species inside the same tick.
If the coinfection hypothesis is true, the observation that >50% of our nymphs had coinfections with B. afzelii and B. garinii requires an additional explanation, given that these two Borrelia species are adapted to different vertebrate hosts (24, 32). We recently presented four mechanisms by which B. afzelii and B. garinii could coinfect the same tick (14). The specificity of B. afzelii for rodents and of B. garinii for birds is not 100% complete. Studies in France (93) and England (94) found that introduced species of rodents were infected with the bird-adapted B. garinii. Similarly, other studies found that birds can transmit the rodent-adapted B. afzelii (95, 96). Rodents or birds that are coinfected with B. afzelii and B. garinii could transmit both species to feeding ticks. Larval ticks could also acquire coinfections of B. afzelii and B. garinii by taking multiple blood meals from different hosts (43, 44), by a combination of cofeeding and systemic transmission, or by a combination of vertical transmission and systemic transmission (14). For example, a larval tick cofeeding next to a B. afzelii-infected nymph on a B. garinii-infected bird would acquire both Borrelia species. Cofeeding transmission has been observed in both B. afzelii and B. garinii (90). In contrast, vertical transmission of Borrelia pathogens is believed to be rare (36).

Horizontal gene transfer explanation for the presence of exotic major ospC group alleles.

In the horizontal gene transfer explanation, the RLB assay result is correct (all nymphs were infected with a single Borrelia species) and the exotic major ospC group alleles were horizontally transferred into the recipient Borrelia species. Previous studies have found that major ospC group alleles can be transferred between Borrelia species (46, 97, 98). In the present study, we found that 12 of the 21 native ospC major groups occurred as exotic major ospC groups in the other Borrelia species. If the horizontal gene transfer explanation is true, we make the following three observations. First, the present study suggests that horizontal transfer of the ospC gene is much more common than previous reports in the literature would suggest (45, 72, 97100). Second, horizontal gene transfer was more likely for major ospC groups that were common in the donor Borrelia species. Third, some of the exotic ospC alleles increased to an appreciable frequency in the recipient Borrelia species. For example, the major ospC group A10, which is native to B. afzelii, had a relative frequency of 31.58% (60/190 infected nymphs) in B. garinii (Table 3). If the horizontal gene transfer explanation is true, one puzzling observation is why the exotic major ospC groups are so much less abundant (1.88% to 3.09%) than the native ospC groups (98.12% and 96.91%) in the nymph-derived isolates. This observation suggests that the exotic strains are underperforming relative to the native strains. Future studies should perform deep sequencing of other Borrelia genes that differ between B. afzelii and B. garinii (e.g., recA or hbb) in order to differentiate between the coinfection hypothesis and the horizontal gene transfer hypothesis.
In conclusion, despite a large sequencing effort (240,410 sequences), our study was unable to find any ospC alleles that were intermediately divergent (>2% to <8%) from the community of major ospC groups. Our study suggests that cross-reactive antibodies prevent the invasion of intermediately divergent ospC alleles. In contrast, we found no evidence that cross-immunity in the vertebrate host was structuring the community of ospC strains in the tick vector. Deep sequencing found that more than 50% of our Borrelia isolates contained exotic major ospC group alleles that had a low abundance (1.88% to 3.09% of all ospC gene sequences) and that had been derived from other Borrelia species. Two alternative explanations for these exotic major ospC group alleles are coinfections, where only the numerically dominant Borrelia species was detected by the RLB assay, or horizontal transfer of the ospC gene between Borrelia species.

ACKNOWLEDGMENTS

This work was supported by a grant from the Swiss National Science Foundation (SNSF) to Maarten J. Voordouw (FN 31003A_141153). The University of Neuchâtel supported this work by an SNSF overhead grant to Maarten J. Voordouw (U.01851.01 project 4.5).
The members of the working group Tiques et Maladies à Tiques (GDR REID) provided insightful discussions. Thanks go to Marco Pagni from the Swiss Institute of Bioinformatics for helping with the clustering analysis. Thanks go to Jacob Koella and two anonymous reviewers for comments on the manuscript.

Supplemental Material

File (zam999116712so1.pdf)
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.

REFERENCES

1.
Andersson M, Scherman K, Raberg L. 2013. Multiple-strain infections of Borrelia afzelii: a role for within-host interactions in the maintenance of antigenic diversity? Am Nat 181:545–554.
2.
Futse JE, Brayton KA, Dark MJ, Knowles DP, Palmer GH. 2008. Superinfection as a driver of genomic diversification in antigenically variant pathogens. Proc Natl Acad Sci U S A 105:2123–2127.
3.
Juliano JJ, Porter K, Mwapasa V, Sem R, Rogers WO, Ariey F, Wongsrichanalai C, Read A, Meshnick SR. 2010. Exposing malaria in-host diversity and estimating population diversity by capture-recapture using massively parallel pyrosequencing. Proc Natl Acad Sci U S A 107:20138–20143.
4.
Lorono-Pino M, Cropp C, Farfan J, Vorndam A, Rodriguez-Angulo E, Rosado-Paredes E, Flores-Flores L, Beaty B, Gubler D. 1999. Common occurrence of concurrent infections by multiple dengue virus serotypes. Am J Trop Med Hyg 61:725–730.
5.
Frank SA. 2002. Immunology and evolution of infectious disease. Princeton University Press, Princeton, NJ.
6.
Gog JR, Grenfell BT. 2002. Dynamics and selection of many-strain pathogens. Proc Natl Acad Sci U S A 99:17209–17214.
7.
Gupta S, Anderson R. 1999. Population structure of pathogens: the role of immune selection. Parasitol Today 15:497–501.
8.
Brunham RC, Plummer FA, Stephens RS. 1993. Bacterial antigenic variation, host immune-response, and pathogen-host coevolution. Infect Immun 61:2273–2276.
9.
Conway D, Polley S. 2002. Measuring immune selection. Parasitology 125:S3–S16.
10.
Fairlie-Clarke KJ, Shuker DM, Graham AL. 2009. Why do adaptive immune responses cross-react? Evol Appl 2:122–131.
11.
Read AF, Taylor LH. 2001. The ecology of genetically diverse infections. Science 292:1099–1102.
12.
Gilbert SC, Plebanski M, Gupta S, Morris J, Cox M, Aidoo M, Kwiatkowski D, Greenwood BM, Whittle HC, Hill AV. 1998. Association of malaria parasite population structure, HLA, and immunological antagonism. Science 279:1173–1177.
13.
Brisson D, Dykhuizen DE. 2004. ospC diversity in Borrelia burgdorferi: different hosts are different niches. Genetics 168:713–722.
14.
Herrmann C, Gern L, Voordouw MJ. 2013. Species co-occurrence patterns among Lyme borreliosis pathogens in the tick vector Ixodes ricinus. Appl Environ Microbiol 79:7273–7280.
15.
Imwong M, Nakeesathit S, Day NP, White NJ. 2011. A review of mixed malaria species infections in anopheline mosquitoes. Malar J 10:253.
16.
Piesman J, Gern L. 2004. Lyme borreliosis in Europe and North America. Parasitology 129:S191.
17.
Gern L, Humair P. 2002. Ecology of Borrelia burgdorferi sensu lato in Europe, p 149–174. In Lyme borreliosis biology, epidemiology and control. Centre for Agriculture and Biosciences International (CABI), Wallingford, United Kingdom.
18.
Gern L, Estrada-Peña A, Frandsen F, Gray JS, Jaenson TGT, Jongejan F, Kahl O, Korenberg E, Mehl R, Nuttall PA. 1998. European reservoir hosts of Borrelia burgdorferi sensu lato. Zentralbl Bakteriol 287:196–204.
19.
Humair P-F, Gern L. 2000. The wild hidden face of Lyme borreliosis in Europe. Microbes Infect 2:915–922.
20.
Hanincova K, Schäfer S, Etti S, Sewell H-S, Taragelova V, Ziak D, Labuda M, Kurtenbach K. 2003. Association of Borrelia afzelii with rodents in Europe. Parasitology 126:11–20.
21.
Huegli D, Hu C, Humair P-F, Wilske B, Gern L. 2002. Apodemus species mice are reservoir hosts of Borrelia garinii OspA serotype 4 in Switzerland. J Clin Microbiol 40:4735–4737.
22.
Humair P, Gern L. 1998. Relationship between Borrelia burgdorferi sensu lato species, red squirrels (Sciurus vulgaris) and Ixodes ricinus in enzootic areas in Switzerland. Acta Trop 69:213–227.
23.
Humair P-F, Peter O, Wallich R, Gern L. 1995. Strain variation of Lyme disease spirochetes isolated from Ixodes ricinus ticks and rodents collected in two endemic areas in Switzerland. J Med Entomol 32:433–438.
24.
Kurtenbach K, Peacey M, Rijpkema SG, Hoodless AN, Nuttall PA, Randolph SE. 1998. Differential transmission of the genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Appl Environ Microbiol 64:1169–1174.
25.
Humair P, Rais O, Gern L. 1999. Transmission of Borrelia afzelii from Apodemus mice and Clethrionomys voles to Ixodes ricinus ticks: differential transmission pattern and overwintering maintenance. Parasitology 118:33–42.
26.
Hanincová K, Taragelová V, Koci J, Schäfer SM, Hails R, Ullmann AJ, Piesman J, Labuda M, Kurtenbach K. 2003. Association of Borrelia garinii and B. valaisiana with songbirds in Slovakia. Appl Environ Microbiol 69:2825–2830.
27.
Humair P, Postic D, Wallich R, Gern L. 1998. An avian reservoir (Turdus merula) of the Lyme borreliosis spirochetes. Zentralbl Bakteriol 287:521–538.
28.
Taragel'ová V, Koči J, Hanincová K, Kurtenbach K, Derdáková M, Ogden NH, Literák I, Kocianová E, Labuda M. 2008. Blackbirds and song thrushes constitute a key reservoir of Borrelia garinii, the causative agent of borreliosis in Central Europe. Appl Environ Microbiol 74:1289–1293.
29.
Olsén B, Jaenson TG, Noppa L, Bunikis J, Bergström S. 1993. A Lyme borreliosis cycle in seabirds and Ixodes uriae ticks. Nature 362:340–342.
30.
Kurtenbach K, De Michelis S, Etti S, Schäfer SM, Sewell H-S, Brade V, Kraiczy P. 2002. Host association of Borrelia burgdorferi sensu lato—the key role of host complement. Trends Microbiol 10:74–79.
31.
Kurtenbach K, Sewell H-S, Ogden NH, Randolph SE, Nuttall PA. 1998. Serum complement sensitivity as a key factor in Lyme disease ecology. Infect Immun 66:1248–1251.
32.
Kurtenbach K, De Michelis S, Sewell H-S, Etti S, Schäfer SM, Hails R, Collares-Pereira M, Santos-Reis M, Haninçová K, Labuda M. 2001. Distinct combinations of Borrelia burgdorferi sensu lato genospecies found in individual questing ticks from Europe. Appl Environ Microbiol 67:4926–4929.
33.
Gern L, Douet V, López Z, Rais O, Cadenas FM. 2010. Diversity of Borrelia genospecies in Ixodes ricinus ticks in a Lyme borreliosis endemic area in Switzerland identified by using new probes for reverse line blotting. Ticks Tick-Borne Dis 1:23–29.
34.
Rauter C, Hartung T. 2005. Prevalence of Borrelia burgdorferi sensu lato genospecies in Ixodes ricinus ticks in Europe: a metaanalysis. Appl Environ Microbiol 71:7203–7216.
35.
Richter D, Debski A, Hubalek Z, Matuschka F-R. 2012. Absence of Lyme disease spirochetes in larval Ixodes ricinus ticks. Vector-Borne Zoonotic Dis 12:21–27.
36.
Rollend L, Fish D, Childs JE. 2013. Transovarial transmission of Borrelia spirochetes by Ixodes scapularis: a summary of the literature and recent observations. Ticks Tick-Borne Dis 4:46–51.
37.
Grimm D, Tilly K, Byram R, Stewart PE, Krum JG, Bueschel DM, Schwan TG, Policastro PF, Elias AF, Rosa PA. 2004. Outer-surface protein C of the Lyme disease spirochete: a protein induced in ticks for infection of mammals. Proc Natl Acad Sci U S A 101:3142–3147.
38.
Tilly K, Krum JG, Bestor A, Jewett MW, Grimm D, Bueschel D, Byram R, Dorward D, VanRaden MJ, Stewart P. 2006. Borrelia burgdorferi OspC protein required exclusively in a crucial early stage of mammalian infection. Infect Immun 74:3554–3564.
39.
Gilmore RD, Kappel KJ, Dolan MC, Burkot TR, Johnson B. 1996. Outer surface protein C (OspC), but not P39, is a protective immunogen against a tick-transmitted Borrelia burgdorferi challenge: evidence for a conformational protective epitope in OspC. Infect Immun 64:2234–2239.
40.
Preac-Mursic V, Wilske B, Jauris S, Will G, Reinhardt S, Lehnert G, Patsouris E, Mehraein P, Soutschek E, Klockmann U. 1992. Active immunization with OspC protein of Borrelia burgdorferi protects gerbils against B. burgdorferi infection. Infection 20:342–349.
41.
Probert WS, LeFebvre RB. 1994. Protection of C3H/HeN mice from challenge with Borrelia burgdorferi through active immunization with OspA, OspB, or OspC, but not with OspD or the 83-kilodalton antigen. Infect Immun 62:1920–1926.
42.
Probert WS, Crawford M, Cadiz RB, LeFebvre RB. 1997. Immunization with outer surface protein (Osp) A, but not OspC, provides cross-protection of mice challenged with North American isolates of Borrelia burgdorferi. J Infect Dis 175:400–405.
43.
Jacquet M, Durand J, Rais O, Voordouw MJ. 2015. Cross-reactive acquired immunity influences transmission success of the Lyme disease pathogen, Borrelia afzelii. Infect Genet Evol 36:131–140.
44.
Earnhart CG, Buckles EL, Dumler JS, Marconi RT. 2005. Demonstration of OspC type diversity in invasive human Lyme disease isolates and identification of previously uncharacterized epitopes that define the specificity of the OspC murine antibody response. Infect Immun 73:7869–7877.
45.
Wang IN, Dykhuizen DE, Qin WG, Dunn JJ, Bosler EM, Luft BJ. 1999. Genetic diversity of ospC in a local population of Borrelia burgdorferi sensu stricto. Genetics 151:15–30.
46.
Baranton G, Seinost G, Theodore G, Postic D, Dykhuizen D. 2001. Distinct levels of genetic diversity of Borrelia burgdorferi are associated with different aspects of pathogenicity. Res Microbiol 152:149–156.
47.
Lagal V, Postic D, Ruzic-Sabljic E, Baranton G. 2003. Genetic diversity among Borrelia strains determined by single-strand conformation polymorphism analysis of the ospC gene and its association with invasiveness. J Clin Microbiol 41:5059–5065.
48.
Brisson D, Drecktrah D, Eggers CH, Samuels DS. 2012. Genetics of Borrelia burgdorferi. Annu Rev Genet 46:515–536.
49.
Bunikis J, Garpmo U, Tsao J, Berglund J, Fish D, Barbour AG. 2004. Sequence typing reveals extensive strain diversity of the Lyme borreliosis agents Borrelia burgdorferi in North America and Borrelia afzelii in Europe. Microbiology 150:1741–1755.
50.
Hellgren O, Andersson M, Råberg L. 2011. The genetic structure of Borrelia afzelii varies with geographic but not ecological sampling scale. J Evol Biol 24:159–167.
51.
Qiu W-G, Schutzer SE, Bruno JF, Attie O, Xu Y, Dunn JJ, Fraser CM, Casjens SR, Luft BJ. 2004. Genetic exchange and plasmid transfers in Borrelia burgdorferi sensu stricto revealed by three-way genome comparisons and multilocus sequence typing. Proc Natl Acad Sci U S A 101:14150–14155.
52.
Jacquot M, Bisseux M, Abrial D, Marsot M, Ferquel E, Chapuis J-L, Vourc'h G, Bailly X. 2014. High-throughput sequence typing reveals genetic differentiation and host specialization among populations of the Borrelia burgdorferi species complex that infect rodents. PLoS One 9:e88581.
53.
Qiu W-G, Dykhuizen DE, Acosta MS, Luft BJ. 2002. Geographic uniformity of the Lyme disease spirochete (Borrelia burgdorferi) and its shared history with tick vector (Ixodes scapularis) in the northeastern United States. Genetics 160:833–849.
54.
Strandh M, Råberg L. 2015. Within-host competition between Borrelia afzelii ospC strains in wild hosts as revealed by massively parallel amplicon sequencing. Philos Trans R Soc Lond B Biol Sci 370:20140293.
55.
Pérez D, Kneubühler Y, Rais O, Jouda F, Gern L. 2011. Borrelia afzelii ospC genotype diversity in Ixodes ricinus questing ticks and ticks from rodents in two Lyme borreliosis endemic areas: contribution of co-feeding ticks. Ticks Tick-Borne Dis 2:137–142.
56.
Gern L, Hu CM, Kocianova E, Vyrostekova V, Rehacek J. 1999. Genetic diversity of Borrelia burgdorferi sensu lato isolates obtained from Ixodes ricinus ticks collected in Slovakia. Eur J Epidemiol 15:665–669.
57.
Sinsky RJ, Piesman J. 1989. Ear punch biopsy method for detection and isolation of Borrelia burgdorferi from rodents. J Clin Microbiol 27:1723–1727.
58.
Schouls LM, Van De Pol I, Rijpkema SG, Schot CS. 1999. Detection and identification of Ehrlichia, Borrelia burgdorferi sensu lato, and Bartonella species in Dutch Ixodes ricinus ticks. J Clin Microbiol 37:2215–2222.
59.
Alekseev AN, Dubinina HV, Van De Pol I, Schouls LM. 2001. Identification of Ehrlichia spp. and Borrelia burgdorferi in Ixodes ticks in the Baltic regions of Russia. J Clin Microbiol 39:2237–2242.
60.
Burri C, Cadenas FM, Douet V, Moret J, Gern L. 2007. Ixodes ricinus density and infection prevalence of Borrelia burgdorferi sensu lato along a north-facing altitudinal gradient in the Rhône Valley (Switzerland). Vector-Borne Zoonotic Dis 7:50–58.
61.
Fu L, Niu B, Zhu Z, Wu S, Li W. 2012. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics 28:3150–3152.
62.
Li W, Godzik A. 2006. Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics 22:1658–1659.
63.
Guindon S, Dufayard J-F, Lefort V, Anisimova M, Hordijk W, Gascuel O. 2010. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol 59:307–321.
64.
Rice P, Longden I, Bleasby A. 2000. EMBOSS: the European Molecular Biology Open Software Suite. Trends Genet 16:276–277.
65.
Theisen M, Borre M, Mathiesen MJ, Mikkelsen B, Lebech A-M, Hansen K. 1995. Evolution of the Borrelia burgdorferi outer surface protein OspC. J Bacteriol 177:3036–3044.
66.
Anderson JM, Norris DE. 2006. Genetic diversity of Borrelia burgdorferi sensu stricto in Peromyscus leucopus, the primary reservoir of Lyme disease in a region of endemicity in southern Maryland. Appl Environ Microbiol 72:5331–5341.
67.
Richter D, Postic D, Sertour N, Livey I, Matuschka FR, Baranton G. 2006. Delineation of Borrelia burgdorferi sensu lato species by multilocus sequence analysis and confirmation of the delineation of Borrelia spielmanii sp. nov. Int J Syst Evol Microbiol 56:873–881.
68.
Core Team R. 2013. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/.
69.
Lord C, Barnard B, Day K, Hargrove J, McNamara J, Paul R, Trenholme K, Woolhouse M. 1999. Aggregation and distribution of strains in microparasites. Philos Trans R Soc Lond B Biol Sci 354:799–807.
70.
Pielou EC. 1977. Mathematical ecology. John Wiley & Sons, New York, NY.
71.
Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O'Hara RB, Simpson GL, Solymos P, Stevens MHH, Wagner H. 2013. Package ‘vegan.’ Community ecology package, version 2.0-10. https://cran.r-project.org/web/packages/vegan/vegan.pdf.
72.
Wang G, Dam AP, Dankert J. 1999. Evidence for frequent OspC gene transfer between Borrelia valaisiana sp. nov. and other Lyme disease spirochetes. FEMS Microbiol Lett 177:289–296.
73.
Gupta S, Ferguson N, Anderson R. 1998. Chaos, persistence, and evolution of strain structure in antigenically diverse infectious agents. Science 280:912–915.
74.
Balmer O, Tanner M. 2011. Prevalence and implications of multiple-strain infections. Lancet Infect Dis 11:868–878.
75.
Rego RO, Bestor A, Stefka J, Rosa PA. 2014. Population bottlenecks during the infectious cycle of the Lyme disease spirochete Borrelia burgdorferi. PLoS One 9:e101009.
76.
Piesman J, Oliver JR, Sinsky RJ. 1990. Growth kinetics of the Lyme disease spirochete (Borrelia burgdorferi) in vector ticks (Ixodes dammini). Am J Trop Med Hyg 42:352–357.
77.
Gupta S, Maiden MCJ, Feavers IM, Nee S, May RM, Anderson RM. 1996. The maintenance of strain structure in populations of recombining infectious agents. Nat Med 2:437–442.
78.
Grenfell BT, Pybus OG, Gog JR, Wood JL, Daly JM, Mumford JA, Holmes EC. 2004. Unifying the epidemiological and evolutionary dynamics of pathogens. Science 303:327–332.
79.
States SL, Brinkerhoff RJ, Carpi G, Steeves TK, Folsom-O'Keefe C, DeVeaux M, Diuk-Wasser MA. 2014. Lyme disease risk not amplified in a species-poor vertebrate community: similar Borrelia burgdorferi tick infection prevalence and OspC genotype frequencies. Infect Genet Evol 27:566–575.
80.
Girard YA, Travinsky B, Schotthoefer A, Fedorova N, Eisen RJ, Eisen L, Barbour AG, Lane RS. 2009. Population structure of the Lyme borreliosis spirochete Borrelia burgdorferi in the western black-legged tick (Ixodes pacificus) in Northern California. Appl Environ Microbiol 75:7243–7252.
81.
Ogden NH, Lindsay LR, Hanincova K, Barker IK, Bigras-Poulin M, Charron DF, Heagy A, Francis CM, O'Callaghan CJ, Schwartz I, Thompson RA. 2008. Role of migratory birds in introduction and range expansion of Ixodes scapularis ticks and of Borrelia burgdorferi and Anaplasma phagocytophilum in Canada. Appl Environ Microbiol 74:1780–1790.
82.
Ogden NH, Margos G, Aanensen DM, Drebot MA, Feil EJ, Hanincova K, Schwartz I, Tyler S, Lindsay LR. 2011. Investigation of genotypes of Borrelia burgdorferi in Ixodes scapularis ticks collected during surveillance in Canada. Appl Environ Microbiol 77:3244–3254.
83.
Fedorova N, Kleinjan JE, James D, Hui LT, Peeters H, Lane RS. 2014. Remarkable diversity of tick or mammalian-associated borreliae in the metropolitan San Francisco Bay Area, California. Ticks Tick-Borne Dis 5:951–961.
84.
Hamer SA, Hickling GJ, Walker ED, Tsao JI. 2014. Increased diversity of zoonotic pathogens and Borrelia burgdorferi strains in established versus incipient Ixodes scapularis populations across the midwestern United States. Infect Genet Evol 27:531–542.
85.
Mathers A, Smith RP, Cahill B, Lubelczyk C, Elias SP, Lacombe E, Morris SR, Vary CP, Parent CE, Rand PW. 2011. Strain diversity of Borrelia burgdorferi in ticks dispersed in North America by migratory birds. J Vector Ecol 36:24–29.
86.
Ginsberg HS, Buckley P, Balmforth MG, Zhioua E, Mitra S, Buckley FG. 2005. Reservoir competence of native North American birds for the Lyme disease spirochete, Borrelia burgdorferi. J Med Entomol 42:445–449.
87.
LoGiudice K, Ostfeld RS, Schmidt KA, Keesing F. 2003. The ecology of infectious disease: effects of host diversity and community composition on Lyme disease risk. Proc Natl Acad Sci U S A 100:567–571.
88.
Perkins SE, Cattadori IM, Tagliapietra V, Rizzoli AP, Hudson PJ. 2003. Empirical evidence for key hosts in persistence of a tick-borne disease. Int J Parasitol 33:909–917.
89.
Randolph S, Miklisova D, Lysy J, Rogers D, Labuda M. 1999. Incidence from coincidence: patterns of tick infestations on rodents facilitate transmission of tick-borne encephalitis virus. Parasitology 118:177–186.
90.
Voordouw MJ. 2015. Co-feeding transmission in Lyme disease pathogens. Parasitology 142:290–302.
91.
Read A, Taylor L. 2000. Within-host ecology of infectious diseases: patterns and consequences, p 59–75. In Molecular epidemiology of infectious diseases. Arnold, London, United Kingdom.
92.
Berret J, Voordouw MJ. 2015. Lyme disease bacterium does not affect attraction to rodent odour in the tick vector. Parasites Vectors 8:249.
93.
Marsot M, Sigaud M, Chapuis J, Ferquel E, Cornet M, Vourc'h G. 2011. Introduced Siberian chipmunks (Tamias sibiricus barberi) harbor more-diverse Borrelia burgdorferi sensu lato genospecies than native bank voles (Myodes glareolus). Appl Environ Microbiol 77:5716–5721.
94.
Millins C, Magierecka A, Gilbert L, Edoff A, Brereton A, Kilbride E, Denwood M, Birtles R, Biek R. 2015. An invasive mammal (the gray squirrel, Sciurus carolinensis) commonly hosts diverse and atypical genotypes of the zoonotic pathogen Borrelia burgdorferi sensu lato. Appl Environ Microbiol 81:4236–4245.
95.
Dubska L, Literak I, Kocianova E, Taragelova V, Sychra O. 2009. Differential role of passerine birds in distribution of Borrelia spirochetes, based on data from ticks collected from birds during the postbreeding migration period in Central Europe. Appl Environ Microbiol 75:596–602.
96.
Franke J, Moldenhauer A, Hildebrandt A, Dorn W. 2010. Are birds reservoir hosts for Borrelia afzelii? Ticks Tick-Borne Dis 1:109–112.
97.
Livey I, Gibbs C, Schuster R, Dorner F. 1995. Evidence for lateral transfer and recombination in OspC variation in Lyme disease Borrelia. Mol Microbiol 18:257–269.
98.
Barbour AG, Travinsky B. 2010. Evolution and distribution of the ospC gene, a transferable serotype determinant of Borrelia burgdorferi. mBio 1:e00153-10.
99.
Dykhuizen DE, Baranton G. 2001. The implications of a low rate of horizontal transfer in Borrelia. Trends Microbiol 9:344–350.
100.
Jauris-Heipke S, Liegl G, Preac-Mursic V, Rössler D, Schwab E, Soutschek E, Will G, Wilske B. 1995. Molecular analysis of genes encoding outer surface protein C (OspC) of Borrelia burgdorferi sensu lato: relationship to ospA genotype and evidence of lateral gene exchange of ospC. J Clin Microbiol 33:1860–1866.

Information & Contributors

Information

Published In

cover image Applied and Environmental Microbiology
Applied and Environmental Microbiology
Volume 81Number 2215 November 2015
Pages: 7740 - 7752
Editor: H. L. Drake
PubMed: 26319876

History

Received: 16 July 2015
Accepted: 25 August 2015
Published online: 21 October 2015

Permissions

Request permissions for this article.

Contributors

Authors

Laboratory of Ecology and Evolution of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland
Maxime Jacquet
Laboratory of Ecology and Evolution of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland
Lye Paillard
Laboratory of Ecology and Evolution of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland
Olivier Rais
Laboratory of Eco-Epidemiology of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland
Lise Gern
Laboratory of Eco-Epidemiology of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland
Maarten J. Voordouw
Laboratory of Ecology and Evolution of Parasites, Institute of Biology, University of Neuchâtel, Neuchâtel, Switzerland

Editor

H. L. Drake
Editor

Notes

Address correspondence to Jonas Durand, [email protected].

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy