ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

A Mechanistic Basis for Phosphoethanolamine Modification of the Cellulose Biofilm Matrix in Escherichia coli

Cite this: Biochemistry 2021, 60, 47, 3659–3669
Publication Date (Web):November 11, 2021
https://doi.org/10.1021/acs.biochem.1c00605

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

1462

Altmetric

-

Citations

-
LEARN ABOUT THESE METRICS
PDF (3 MB)
Supporting Info (1)»

Abstract

Biofilms are communities of self-enmeshed bacteria in a matrix of exopolysaccharides. The widely distributed human pathogen and commensal Escherichia coli produces a biofilm matrix composed of phosphoethanolamine (pEtN)-modified cellulose and amyloid protein fibers, termed curli. The addition of pEtN to the cellulose exopolysaccharide is accomplished by the action of the pEtN transferase, BcsG, and is essential for the overall integrity of the biofilm. Here, using the synthetic co-substrates p-nitrophenyl phosphoethanolamine and β-d-cellopentaose, we demonstrate using an in vitro pEtN transferase assay that full activity of the pEtN transferase domain of BcsG from E. coli (EcBcsGΔN) requires Zn2+ binding, a catalytic nucleophile/acid-base arrangement (Ser278/Cys243/His396), disulfide bond formation, and other newly uncovered essential residues. We further confirm that EcBcsGΔN catalysis proceeds by a ping-pong bisubstrate–biproduct reaction mechanism and displays inefficient kinetic behavior (kcat/KM = 1.81 × 10–4 ± 2.81 × 10–5 M–1 s–1), which is typical of exopolysaccharide-modifying enzymes in bacteria. Thus, the results presented, especially with respect to donor binding (as reflected by KM), have importantly broadened our understanding of the substrate profile and catalytic mechanism of this class of enzymes, which may aid in the development of inhibitors targeting BcsG or other characterized members of the pEtN transferase family, including the intrinsic and mobile colistin resistance factors.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence
Biofilms are dynamic communities of microorganisms fixed in a synthesized matrix of secreted polysaccharide materials. (1) This biofilm matrix serves as the interface between its constituents and their outside environment, but also plays many other roles to the benefit of the enclosed microorganisms that may justify the large resource cost of producing copious quantities of this extracellular matrix material. (2) For example, biofilms have been observed to serve in resource capture, (3) accelerate cell growth, (4) and improve tolerance to stressors or disinfectants, such as shearing forces, (2) desiccation, (2) antimicrobial compounds, (2,5) extreme temperature, (2,5,6) or sanitizing agents (2,3,5) for the enclosed microorganisms. The persistent colonization of surfaces by the widely distributed human and animal pathogens Escherichia coli and Salmonella spp. has been linked to the production of biofilms composed of the polysaccharide cellulose and amyloid protein fibers termed curli. (7) In addition to these species, biofilms containing cellulose have been reported in numerous other species, including other pathogens of plants and humans, like Agrobacterium tumefaciens, Cronobacter sakazakii, Acetobacter xylinum, Clostridium (previously Sarcina) spp., Rhizobium spp., and some Pseudomonads. (8−12) Although the biofilms of these bacteria were long thought to contain a similar linear β-(1–4)-glucan, the diversity of these biofilm polysaccharides is beginning to be understood. (13) For example, the biofilm of Pseudomonas fluorescens SBW25 was observed to contain an acetylated form of cellulose, (8) and more recently, the discovery of phosphoethanolamine (pEtN) cellulose was reported in E. coli and Salmonella enterica (Figure 1). (14) Cellulose has also been recently proposed in the Gram-positive human pathogen Clostridioides difficile, (12) and an operon for the production of Pel exopolysaccharide was reported in Bacillus cereus, (15) thereby expanding the repertoire of bacteria known to produce these exopolysaccharides beyond Gram-negative organisms.

Figure 1

Figure 1. Schematic of the reaction catalyzed by BcsG. Phosphoethanolamine is added to approximately 50% of glucose units of bacterial cellulose, exclusively at the C6 position with an unknown pattern across the polymer. (14,25)

At present, all Gram-negative bacteria known to produce cellulose as a component of their biofilm possess an operon encoding the essential and conserved protein complex responsible for cellulose biosynthesis and export. (16) Structural and functional studies, informed by research on other polysaccharide secretion systems, have provided evidence for the roles of each gene product in this operon, often annotated bcsABCZ. (17) Biosynthesis of cellulose occurs by the action of the processive glycosyltransferase BcsA at the cytoplasmic face of the inner membrane, using UDP-glucose as a substrate. (18) Regulation of cellulose biosynthesis occurs through the binding of the second messenger molecule cyclic dimeric guanosine monophosphate (c-di-GMP) to the PilZ domain of BcsA, allosterically regulating the activity of the glycosyltransferase domain. (19) Translocation of the growing chain from the cytoplasm to the periplasm occurs co-synthetically, through a pore in BcsA, and is aided by a carbohydrate-binding domain found in the periplasmic protein BcsB. (18,20) Similar to structural and functional studies of other polysaccharide systems, cellulose export across the outer membrane was recently shown to require an outer membrane β-barrel and periplasmic tetratricopeptide repeat (TPR)-containing protein, encoded by BcsC. (21−24) Finally, a periplasmic polysaccharide lyase or glycoside hydrolase is found in many known polysaccharide biosynthesis operons. (22) With respect to microbial cellulose, structural and functional studies of BcsZ from E. coli and CcsZ from C. difficile demonstrated that these terminal enzymes were from glycoside hydrolase families 8 and 5, respectively, and both possess endo-β-(1,4)-glucanase activity for modification of the polymer. (12,26)
Previously, the role of accessory operons or genes within these operons remained unknown until the discovery of modified polysaccharide materials within biofilms. (17) For example, in cellulose-producing organisms, the role of a type II operon, containing bcsEFG, was unknown, and its gene products showed little similarity to other characterized proteins based on sequence. (17) These genes were annotated as necessary for maximal cellulose production, (17) but more recently, structural and functional characterization has shed light on their true roles in processing of the cellulose polysaccharide. (14,25,27−29) The BcsE protein was shown to comprise both a degenerate receiver domain and a GGDEF domain responsible for sensing c-di-GMP. (28) BcsE has also recently been proposed to be responsible for the recruitment of BcsQ and BcsR to the membrane during early Bcs macrocomplex assembly. (29) Other recent efforts with cryo-electron microscopy have resolved the native Bcs macrocomplex in several states and demonstrate its piecewise assembly. (30,31) In light of the discovery of pEtN cellulose, bcsG was also shown to be necessary for the architecture and assembly of the biofilm matrix. (14) BcsG was further proposed to be the pEtN transferase responsible for the postsynthetic modification of cellulose in the periplasm of organisms producing a pEtN cellulose polysaccharide. (14,27) We previously reported the structure and activity of BcsG from E. coli (EcBcsG), demonstrating that it acts as a pEtN transferase with substrate specificity for cellulose. (25) Additionally, the structure and function of BcsG from Salmonella typhimurium (StBcsG) was independently reported, and the results support the conclusion that EcBcsG and StBcsG are functionally equivalent pEtN transferases acting on cellulose polysaccharides. (27)
Both EcBcsG and StBcsG studies reported the functional complementation of bcsG using a library of BcsG amino acid variants. (25,27) These combined results demonstrated the essential catalytic residues of BcsG, which are partially conserved among other characterized pEtN transferases, including the intrinsic and mobile colistin resistance factors. Disruption of these essential residues in EcBcsG resulted in a fragile pellicle biofilm phenotype in E. coli AR3110 that was indistinguishable from a strain bearing a chromosomal deletion of bcsG. (25) However, the enzymatic consequences of amino acid substitutions that resulted in these phenotypes remain unexplored. Using an in vitro pEtN transferase assay, we show here that Zn2+ binding, disulfide bond formation, and the catalytic nucleophile/acid-base arrangement (Ser278/Cys243/His396) are essential for full transferase activity of the recombinant catalytic domain of EcBcsG (herein EcBcsGΔN). Furthermore, we expand the donor substrate profile to confirm EcBcsGΔN is specific for pEtN transfer and report the Michaelis–Menten kinetics of EcBcsG using the co-substrate analogues p-nitrophenyl phosphoethanolamine (p-NPPE) and cello-oligosaccharides. We also demonstrate that the kinetic behavior of EcBcsG further supports a ping-pong bisubstrate–biproduct (bibi) catalytic mechanism and expand the active site to include new residues proposed by examination of the structural model to be involved in catalysis.

Experimental Procedures

ARTICLE SECTIONS
Jump To

Materials

Culture media, isopropyl β-d-thiogalactopyranoside (IPTG), and kanamycin sulfate were purchased from BioBasic (Markham, ON). Cello-oligomers were purchased from Megazyme (Dublin, Ireland). Nickel-nitrilotriacetic acid (Ni-NTA) agarose was purchased from Qiagen (Valencia, CA). Unless otherwise specified, all other materials were purchased from Sigma-Aldrich Canada Ltd. (Oakville, ON).

Cloning, Expression, and Purification of EcBcsGΔN

The catalytic domain of the bcsG gene was cloned into the pET28a expression vector, as described previously. (25) Briefly, bcsGΔN was amplified from E. coli K12 chromosomal DNA and cloned into the pET28a expression vector using NdeI and XhoI restriction sites. The polymerase chain reaction (PCR) products were digested by the appropriate enzymes for 1 h, followed by purification using a GeneJET PCR purification kit (Thermo Fisher Scientific). The purified PCR product was ligated into an NdeI/XhoI doubly digested vector using T4 DNA ligase (Thermo Fisher Scientific) at 22 °C for 4 h. The ligation mixture was used to transform E. coli TOP10 (Thermo Fisher Scientific), and pET28a-bcsGΔN was isolated from 16–18 h grown cultures using an EZ-10 plasmid prep kit (Bio Basic). The resulting recombinant BcsG C-terminal catalytic domain, EcBcsGΔN, was comprised of residues Ala164–Gln559 from UniProt entry P37659 and included the addition of an N-terminal hexahistidine tag conferred by the pET28a expression plasmid.

Site-Directed Mutagenesis

Site-directed mutagenesis was performed using the PCR primer method with pET28a-bcsGΔN as a template. Amino acid variants of Arg458 and Ser244 were generated from site-directed mutagenesis of the wild-type pET28a-bcsGΔN plasmid as a service provided by BioBasic. The primers used to generate the alanine EcBcsGΔN amino acid variants of Cys243, Tyr277, Ser278, His396, Glu442, and His443 are available in Table S1. Following amplification of amino acid variant plasmids, DpnI (1 unit) was added to reaction mixtures to digest methylated template DNA. The reaction mixture was transformed into E. coli TOP10, and the resulting clones were sequenced to confirm the presence of the desired mutations (The Centre for Applied Genomics).

Expression and Purification of EcBcsGΔN and Derivatives

pET28a-bcsGΔN and derivative plasmids thereof were transformed into E. coli BL21 (DE3), and cells were grown in 1 L volumes of rich medium (32 g of tryptone, 20 g of yeast extract, and 10 g of NaCl) containing kanamycin (50 μg mL–1). Cultures were incubated at 37 °C while being shaken, until they reached an optical density at 600 nm of 0.6–0.8. Expression of EcBcsGΔN and amino acid variants thereof was induced by the addition of IPTG (1 mM), and growth was allowed to continue for 16–20 h at 23 °C. The cells were collected by centrifugation (4300 × g for 15 min at 4 °C) and stored at −20 °C until use. Cell pellets were resuspended in lysis buffer [50 mM Tris (pH 8.0) and 300 mM NaCl] and supplemented with 0.5 mg of RNase A (BioBasic), 0.25 mg of DNase I (BioBasic), and for some variants one EDTA-free mini protease inhibitor tablet (Roche). The resulting cell suspensions were lysed using a cell disruptor (Constant Systems) operating at 17000 psi (117211 kPa) and 4 °C. The lysate was separated by centrifugation (28000 × g for 45 min at 4 °C), and hexahistidine-tagged EcBcsGΔN (and amino acid variants thereof) were purified from the soluble fraction using nickel affinity chromatography. Ni-NTA resin (Thermo Fisher Scientific) pre-equilibrated in lysis buffer (2 mL) was suspended in the cell lysate and incubated for 1 h at 4 °C to facilitate binding. This solution was filtered by being passed over a chromatography column and washing of the resin in at least three 10 mL volumes of lysis buffer, followed by three 10 mL volumes of lysis buffer with the addition of 25 mM imidazole. Elution was achieved by addition of 10 mL of elution buffer (lysis buffer with the addition of 250 mM imidazole) to the chromatography resin. Buffer exchange was performed by a two-step dialysis against 50 mM Tris (pH 8.0) and 150 mM NaCl using two 2 L volumes for at least 1 h per volume at 4 °C. Secondary purification was performed by anion exchange chromatography when deemed necessary as assessed by sodium dodecyl sulfate–polyacrylamide gel electrophoresis. For secondary purification, proteins were buffer exchanged by dialysis against 50 mM Tris (pH 8.0) using two 2 L volumes for at least 1 h per volume at 4 °C. The resulting solutions were loaded onto an ÄKTA Start fast-protein liquid chromatography instrument (Cytiva) equipped with a HiTrap Q FF cartridge (1 mL, Cytiva). The solvents were (A) 50 mM Tris (pH 8.0) and (B) 50 mM Tris and (pH 8.0) and 1 M NaCl. The protein was bound by passing sample five times over the column at a flow rate of 1 mL/min in 100% A. The column was washed for 20 min with 100% A, followed by elution in a linear gradient from 0% to 100% B over 30 min. Peaks of >50 mAU at 280 nm were collected using a Frac30 fraction collector (Cytiva). All purification buffers were supplemented with 5 mM MgCl2, as we observed this to improve the stability of recombinant EcBcsGΔN and derivatives. Recombinant protein eluted from purification columns was pooled and concentrated to between 30 and 45 mg mL–1 in a 30 kDa molecular weight cutoff centrifugal filter unit (Cytiva) and stored at 4 or −20 °C for short- or long-term storage, respectively.

Measurement of the Enzymatic Rate

The enzymatic rate of EcBcsGΔN and derivatives was measured with the chromogenic assay using the mock co-substrates p-nitrophenyl phosphoethanolamine (p-NPPE), p-nitrophenyl phosphopropanolamine (p-NPPP), p-nitrophenyl phosphate (p-NPP) (Sigma), and cellooligosaccharides, as described previously. (25) The synthesis of p-NPPE was performed as described previously. (25) The detailed synthesis of p-NPPP is described in the Supporting Information. The esterase activity of EcBcsGΔN and amino acid variants thereof was measured using 7 mM p-NPPE and 50 μM enzyme. The substrates p-NPPP and p-NPP were tested at 2 mM with and without 3 mM cellopentaose. Transferase activity was measured under the same conditions, but only in the presence of cellooligosaccharides, as described previously. (25) All reactions were carried out in 50 mM HEPES buffer (pH 7.5) containing 50 mM NaCl at 37 °C using 100 μL volumes. Concentrations of 1 mM ethylenediamine tetraacetic acid (EDTA) or dithiothreitol (DTT) were exposed to EcBcsGΔN by buffer exchange for no less than 24 h prior to the assay. Reactions were monitored in a Cytation5 imaging plate reader (Biotek) at 37 °C for 30 min at 414 nm. The specific activity and velocity of the reactions were calculated from absorbance data using a standard curve prepared with p-nitrophenol analytical reference material (Sigma, catalog no. 241326). All rates are reported as the mean ± the standard deviation (SD) of at least three replicates. Figure preparation and statistical analysis were performed using GraphPad Prism version 9.0.1.

Steady-State Kinetics

Kinetic assays were performed using the same assay design described above but with p-NPPE concentrations as specified that ranged from 0 to 14.5 mM. Kinetics with the cellooligosaccharide acceptor were also conducted with concentrations ranging from 0 to 6 mM with a fixed p-NPPE concentration of 7 mM. Michaelis–Menten parameters (kcat and KM) were determined by nonlinear regression analysis of plots of initial velocity (nanomoles per second) as a function of p-NPPE concentration. All model fitting, analysis, and figure preparation were performed using GraphPad Prism version 9.0.1.

Mass Spectrometry

Liquid chromatography–mass spectrometry analyses were performed at the Mass Spectrometry Facility of the Advanced Analysis Centre (University of Guelph, Guelph, ON) on an Agilent 1200 HPLC liquid chromatograph interfaced with an Agilent UHD 6530 Q-TOF mass spectrometer. A C18 column (Agilent Extend-C18, 50 mm × 2.1 mm, 1.8 μm) was used for chromatographic separation. The following solvents were used for separation: water with 0.1% (v/v) formic acid (A) and acetonitrile with 0.1% (v/v) formic acid (B). The initial mobile phase conditions were 10% B, hold for 1 min, and then increase to 100% B over 29 min.

Results and Discussion

ARTICLE SECTIONS
Jump To

Amino Acid Replacements

To assess the catalytic consequences of amino acid substitutions within the EcBcsG active site, we constructed a library of amino acid EcBcsGΔN variants with alanine substitutions of active site residues observed to be essential for catalytic activity in vivo (Cys243, Tyr277, Ser278, His396, Glu442, and His443). These residues represent the structural equivalent to the active site of known pEtN transferases. Interestingly, only His443 and His396 are strictly conserved, although our previous work demonstrated that the remainder are catalytically essential in vivo and probably represent the functional equivalents to the conserved residues of other pEtN transferases. (25) However, the consequences of these amino acid replacements remained untested on an enzymatic level, so herein we explored these residues, along with two new residues we suspected may be catalytically important (Ser244 and Arg458) based on inspection of the BcsG structure and that of its homologues. All of the constructs used in this study (C243A, S244A, Y277F, S278A, H396A, E442A, H443A, R458A, and R458H) were generated using site-directed mutagenesis of pET28a-bcsGΔN. Each of these recombinant proteins were expressed in E. coli BL21 transformants with observed yields similar to that of wild-type EcBcsGΔN. These variant proteins were purified to apparent homogeneity using the protocol described previously, (25) with additional washing steps as deemed necessary for purification to homogeneity. Care was taken to use new chromatography media for the purification of each respective enzyme to prevent enzyme carryover and cross-contamination of assay data with multiple enzyme forms.
As expected, replacement of the nucleophilic Ser278 (S278A) resulted in a loss of any detectable activity. The observed rates of both esterase (i.e., bulk solvent serves as the pEtN acceptor) and transferase activity (i.e., in the presence of a defined carbohydrate acceptor) of the S278A replacement were not different from an enzyme-free control (Figure 2; t tests, t(4) = 0.3586 and p = 0.7830 for esterase; t(4) = 0.4935 and p = 0.6475 for transferase). This lack of activity supports the role of this nucleophilic serine in the catalytic mechanism. His396 is highly conserved and has also been shown to be catalytically important in other pEtN transferases. (32) Replacement of this histidine in BcsG (H396A) also led to reductions in activity with a residual esterase activity of 16.6% and a residual transferase activity of 7.0% (Figure 2), which is consistent with results previously noted for this mutation in the in vivo biofilm assays, where alanine substitution of this residue resulted in the fragile pellicle phenotype. (25)

Figure 2

Figure 2. Amino acid replacements of essential catalytic residues abrogate EcBcsGΔN activity. (A) The structure of the EcBcsGΔN active site (PDB entry 6PD0) is depicted here and was previously mapped by functional complementation in vivo. (25) (B) The measured rates of esterase (i.e., carbohydrate acceptor-free conditions) and transferase (i.e., in the presence of a cellopentaose acceptor) activities of EcBcsGΔN amino acid variants within the catalytic site demonstrate these residues are essential in vitro with the exception of Y277. Asterisks denote statistical significance [Tukey’s multiple comparisons, q(33) > 8.40 and p < 0.0001 for all comparisons].

Substantial losses of activity were also observed with alanine replacement of the metal-binding triad residues of the enzyme (Figure 2). The Glu442 replacement (E442A) resulted in an esterase activity indistinguishable from an enzyme-free control (e.g., 1.96 ± 0.07 nmol min–1 mg–1), and only 2.3% residual transferase activity was detected in the presence of β-d-cellopentaose. Replacement of Cys243 with alanine (C243A) also resulted in low residual specific activities of 4.7% (esterase) and 2.4% (transferase). The His443 substitution (H443A) also demonstrated reduced specific activities, although to a lesser degree, representing 38.5% residual esterase and 10.4% residual transferase activity. The retention of some activity by His443 may be explained by the observation of Ser278 also serving as a ligand for Zn2+ binding in some of the structures when His443 is not. Completion of the Zn2+ coordination sphere in at least one of the structural models shows the Zn2+ also bound to a water molecule, which is consistent with the Zn2+ playing a catalytic and not a structural role. (33) Combined, these observations and the low residual specific activities are consistent with the essential Zn2+ binding roles of these residues that had previously been noted in the structural model and assayed in vivo. (25) These findings further support the observation that Zn2+ binding is indispensable for the catalytic activity of BcsG in vivo, suggesting the residues we replaced here are truly equivalent to the catalytic residues of other polymyxin resistance factors. (25)
Previous mass spectrometry and structural data we collected suggested that an additional residue may contribute to the catalytic fold and belonged to the loop that presents Cys243 to the active site. (25) Although we did not test a full-length BcsG amino acid replacement in vivo, we propose that the identity of this residue is Ser244 (Figure 2). The side chain hydroxyl of Ser244 is oriented toward the Zn2+ at an atomic distance of 4.5 Å, and this residue is the only one that could plausibly serve an essential function given both our structural and mass spectrometry data sets. In support of our hypothesis, we found that replacement of Ser244 with alanine (S244A) resulted in 7.2% residual esterase and 5.8% residual transferase activity, demonstrating that Ser244 is indeed part of the active site and is required for full activity.
To our surprise, replacement of Tyr277 with Phe (Y277F) produced an enzyme variant with greater esterase activity and 84% residual transferase activity compared to wild-type EcBcsGΔN (Figure 2). This observation did not support our earlier proposal that Tyr277 may function in cellulose accommodation in the proximity of the Zn2+ ion at the active center. (25) An increase in solvent accessibility to the active site may account for the improved rate of esterase activity and an increased level of competition for transfer to oligosaccharides that is reflected by the lower transferase activity. This hypothesis is further supported by the fact that an equivalent tyrosine residue is not observed in available structures of other pEtN transferases, (32,34,35) nor was a tyrosine residue observed to participate in contact with the lipid-binding site in the structure of the full-length pEtN transferase EptA from Neisseria meningitidis [NmEptA, Protein Data Bank (PDB) entry 5FGN]. (36) Instead, the residue equivalent to Tyr277 in NmEptA, Ser279, appears to shape the lipid-binding pocket at the interface of the catalytic and membrane domains. (36) Consequently, we rationalized that this disparity in our data may indicate that the role of the Tyr277 hydroxyl may also include the stability of the local folding and interaction of the EcBcsG C-terminal catalytic domain relative to the N-terminal transmembrane region and, thus, allow EcBcsG to sample productive conformations during the catalytic cycle. This hypothesis would account for the importance of Tyr277 in EcBcsG activity observed in vivo, but was not needed for EcBcsGΔNin vitro activity. (25)
The MCR-1 pEtN transferase has been resolved by X-ray crystallography with both one Zn2+ ion and two Zn2+ ions at the active site (32) (PDB entries 5LRN and 5LRM, respectively). Although the Zn2+ ion resolved in EcBcsGΔN is found at the conserved location found in all pEtN transferases, it remains unclear if the second Zn2+ site reported for MCR-1 is catalytically important and whether this site is found in other pEtN transferases. Two His residues in MCR-1 coordinate this second Zn2+ ion and are equivalent to His396 and Arg458 in EcBcsG. To assess if Arg458 plays a role in secondary Zn2+ binding or catalysis, we measured the enzymatic rates of alanine (R458A) and histidine (R458H) replacements (Figure 2). We found that the R458A variant displayed 25.6% esterase and no detectable transferase activity, while the R458H variant displayed 11.4% esterase and also did not display detectable transferase activity compared to that of the wild type. As the histidine imidazole group would be expected to functionally complement the arginine guanidino group in metal binding, these results suggest that Arg458 does not participate in secondary metal binding. Instead, because Arg458 contributes a guanidino group at a similar distance and across the face of the active site from His396 (i.e., 5 Å above the face of the Zn2+ ion), this configuration is consistent with the possibility that Arg458 stabilizes the charged pEtN–enzyme intermediate formed during the first mechanistic step that results in reduced esterase activity in both mutants. Because replacement of Arg458 with either Ala or His also abrogated transferase activity, this residue likely also plays an essential role in accommodation of the acceptor co-substrate during the second mechanistic step. An in situ assay for biofilm phenotype in S. typhimurium using BcsG variants suggested that both His and Ala replacements of the equivalent to Arg458 resulted in reduced cellulose production by S. typhimurium, (27) corroborating the role of this conserved Arg residue in substrate binding and catalysis rather than Zn2+ binding.

Molecular Determinants

To assess if the losses of activity we observed for substitutions of the Zn2+-binding residues were in fact due to the loss of Zn2+ in the EcBcsGΔN active site or instead due to structural changes in the enzyme, we supplied the metal chelating agent ethylenediaminetetraacetic acid (EDTA) to EcBcsGΔN prior to assaying. We observed that treatment with 1 mM EDTA significantly reduced the specific esterase activity of EcBcsGΔN to 10.6% of the untreated activity [Figure 3; Tukey’s test, q(14) = 12.12, and p < 0.0001]. A matched enzyme-free control containing 1 mM EDTA displayed no difference in the rate of p-NPPE turnover, suggesting the effect seen was due to loss of Zn2+ in the active site. This significant loss of esterase activity agreed with the large reduction in esterase activity observed for the alanine replacements of Cys243, Glu442, and His443. To our surprise, however, the loss of transferase activity observed for EDTA treatment was markedly smaller than that observed for single-amino acid substitutions of the Zn2+-binding residues (i.e., 43% residual activity of the wild type for the EDTA-treated form, compared to 2.4%, 2.3%, and 10.4% residual transferase activity for alanine replacements of Cys243, Glu442, and His443, respectively). Our data suggest either that BcsG has an exceptional affinity for Zn2+, given that 1 mM EDTA is a standard concentration used to assess the metal dependency of enzymes, or that in addition to Zn2+ binding, Cys243, Glu442, and His443 are also catalytically important residues, particularly during the latter step of catalysis when pEtN is transferred to cellulose. In support of this theory, any general mechanism involving Ser278 as the catalytic nucleophile would also require a general acid–base residue to abstract the serine hydroxyl proton during the formation of the covalent enzyme intermediate and then subsequently replace it following transfer of the pEtN group to its ultimate acceptor. Our results also mirror the findings that treatment of MCR-1-expressing cells with 1 mM EDTA reduces but does not abolish the colistin MIC to MCR-1 negative levels. Similarly, replacement of the residue equivalent to Cys243 in MCR-1 (Glu246) reduces the colistin MIC to the same extent as the vector control, which is beyond what is observed for EDTA treatment. In the literature, there are examples of cysteine residues acting as a catalytic base when proximal to a histidine, (37) but this would be unprecedented among pEtN family members. Indeed, cysteine is an objectively poorer acid–base catalyst, and the Glu442 residue seems intuitively better suited for this function. However, given that alanine replacements of both Glu442 and Cys243 are significantly less active than either a single replacement of His443 or EDTA-treated enzyme, these data suggest that these two amino acids and/or His396 (as noted below) are the leading candidates to participate as acid–base catalytic residues.

Figure 3

Figure 3. Zn2+ binding and disulfide bond formation are essential catalytic features of EcBcsGΔN. Treatment of EcBcsGΔN with the metal chelating agent EDTA or the reducing agent DTT results in impaired enzymatic activities, consistent with impaired biofilm-forming phenotypes observed in vivo. Asterisks denote significant differences from respective controls [Tukey’s multiple comparisons, q(14) > 6.5 and p < 0.004 for all comparisons].

A disulfide bond that fixes the helical element presenting the catalytic nucleophile is a conserved and essential feature of pEtN transferases. (25,38) We previously showed that BcsG contains a disulfide bond between Cys290 and Cys306, and that replacement of the Cys pair with Ala results in a loss of catalytic activity in vivo. (25) Surprisingly, this disulfide bond in BcsG is not at the conserved location but serves to fix the same helical element presenting the nucleophilic Ser278. We were unable to isolate a C290A/C306A variant of EcBcsGΔN, further suggesting that the disulfide bond is a critical structural feature of BcsG. To assess the importance of disulfide bond formation to enzymatic activity in vitro, we introduced the reducing agent dithiothreitol (DTT) to EcBcsGΔN prior to measurement of the enzymatic rate. An observed decrease in enzymatic activity, representing 31.9% of residual esterase activity, was observed (Figure 3). Surprisingly, however, the measured residual transferase activity was 72.8% of that of wild-type EcBcsGΔN, while a matched enzyme-free control showed no differences in the rate of p-NPPE hydrolysis. These findings are in apparent disagreement with the loss of function of the disulfide-deficient EcBcsG variant we reported in vivo. These data may in part be rationalized by the specific requirements of exopolysaccharide modification for biofilm formation. For example, the biofilm polysaccharide poly-β-(1,6)-N-acetyl-d-glucosamine (PNAG) requires only partial deacetylation by the enzyme PgaB or its orthologues for successful biofilm formation in various bacteria. (39−41) PNAG found in biofilms has been observed to have approximately 15–20% of the N-acetyl-d-glucosamine saccharide units deacetylated, with either more or less extensive deacetylation causing disruption in biofilm formation, probably due to the loss of a distributed cationic charge on the polymer. (40,42,43) Similarly, a loss of EcBcsGΔN activity of approximately 30% measured in the presence of DTT might be considered trivial for the biological function of some enzymes; however, this loss may not be enough to maintain the 50% level of pEtN modification of d-glucose saccharide units that has been observed for pEtN cellulose biofilms isolated from E. coli or S. enterica. (14) Thus, even subtle reductions in the enzymatic rate, such as those observed here, could still plausibly alter the degree of pEtN substitution so that it is not matched properly to the rate of cellulose synthesis, thereby disrupting the otherwise uniform charge and chemistry required for proper biofilm architecture and assembly in vivo.

Donor Co-substrate Preference

Although it has been established that the natural co-substrates are phosphatidylethanolamine (14,27) and bacterial cellulose, (14) we demonstrated that BcsG was not capable of transfer to equivalent linear β-1,4-aminosugar polymers (i.e., chitin). (25) However, the sufficiency of BcsG to use alternative phosphatidylethanolamine mimetics as phospho-donors remains unexplored, which may be of interest for materials science and development. To investigate the phospho-donor preference of EcBcsG, we assayed the enzyme in vitro using the commercially available substrate analogue p-nitrophenyl phosphate (p-NPP). Additionally, using a synthetic approach similar to that of p-NPPE, (25) the related compound p-nitrophenyl phosphopropanolamine (p-NPPP) was synthesized in two steps (see the Supporting Information). BcsG demonstrated detectable esterase activity on both the minimal substrate p-NPP and the extended substrate p-NPPP, although to a lesser extent than the preferred p-NPPE (Figure S1). We repeated these experiments in the presence of cellopentaose as an acceptor substrate and observed no significant increase in the rate of turnover of p-NPP or p-NPPP that would suggest catalytic transfer of the ester-linked phosphate or phosphopropanolamine. Analysis of the enzymatic products by LC-MS corroborated these results, as no evidence of phospho- or phosphopropanolamine cellulose was detected, thereby indicating that BcsG is exclusively capable of transferring the pEtN functional group under these conditions but also further confirming it as the true biological substrate.

Steady-State Esterase Kinetics

To investigate the kinetic parameters of EcBcsGΔN, we measured the steady-state esterase rate and calculated the specific activity at varied concentrations of p-NPPE (0–14.5 mM) without the addition of an acceptor co-substrate (Figure 4A). The resulting data were fit in agreement with the Michaelis–Menten model, with an R2 value of 0.9624 (Table 1). We calculated from our data an apparent KM of 2.40 ± 0.28 mM, a maximal rate (Vmax) of 2.20 × 10–2 nmol s–1, and a kcat of 4.34 × 10–7 ± 1.30 × 10–8 s–1. The derived catalytic efficiency kcat/KM was 1.81 × 10–4 ± 2.81 × 10–5 M–1 s–1.

Figure 4

Figure 4. Steady-state kinetics of EcBcsGΔN. (A) Steady-state esterase (WT) and transferase (mutant) kinetics using p-NPPE as the phosphoethanolamine donor and 3 mM cellopentaose as the acceptor co-substrate. (B) Steady-state transferase kinetics using 7 mM p-NPPE as the donor and cellooligosaccharides with DP values of 4–6 as the acceptor co-substrates.

Ser244 and His396 Participate in Both Mechanistic Steps but Not in Zn2+ Binding

While the roles of the other active site residues can be implied from the crystal structure and from trapping of the covalent enzyme intermediate, the catalytic importance of Ser244 and His396 was not obviated from prior data, notably for His396 that may plausibly serve as the complementary base for the reaction that has not otherwise been identified. (25,27) Although our specific activity data (noted above) suggest that these variants of EcBcsGΔN display impaired ability to transfer pEtN to cellulosic substrates, it remained unclear if these residues are exclusively involved in the latter half of the mechanism whereby pEtN is transferred to cellulose, or if their roles might extend to cleavage of the pEtN substrate. Therefore, we assessed the steady-state kinetics of the S244A and H396A EcBcsGΔN variants in the presence of 3 mM cellopentaose.
The alanine replacement of Ser244 demonstrated a decrease in turnover number (kcat) and a greatly reduced affinity for p-NPPE (KM increases >3-fold) in the presence of acceptor, suggesting this residue plays a role in both steps of the proposed mechanism (Figure 4A and Table 1). While other biochemically characterized pEtN transferases possess a conserved Thr residue that is equivalent to Ser244 in BcsG, none of the previous work has kinetically explored the role of this residue in catalysis. Thus, the results presented herein, especially with respect to donor binding (as reflected by KM), have importantly broadened our understanding of the catalytic mechanism of this class of enzymes, which may aid in the development of inhibitors. Other biochemically characterized pEtN transferases, including MCR-1, NmEptA, and CjEptC, each possess a conserved histidine equivalent to the BcsG His396 residue. (32,34,35) Although this His residue has been shown to be essential for at least MCR-1, (32) it has also been proposed to coordinate a second Zn2+ ion at the active site, which is apparently required for catalysis (Figure 5C). (32,44) In addition to our specific activity results (Figure 2), our kinetic analysis of His396 in EcBcsGΔN confirmed the importance of this residue in catalysis and expanded our knowledge of its role in the reaction mechanism. Specifically, H396A demonstrated a lower turnover number (kcat decreases 2.6-fold) and a modest reduction in affinity for p-NPPE in the presence of cellopentaose, thereby indicating that this residue is important in both steps of the proposed mechanism (Figure 4A and Table 1). However, our group and others (27) were not able to support the existence of a second Zn2+ ion in the active site of BcsG, suggesting the role of His396 is truly in catalysis rather than secondary metal binding among the pEtN transferases. In the structural models of BcsG, His396 is positioned above the zinc ion (approximately 5 Å) and in some models is within hydrogen bonding distance of the main chain carbonyl of Asp397, as well as a water molecule when present (Figure 5A). These results are consistent with the distances noted for the equivalents to His396 and Asp397 in ICRMc (His429 and Gly430, respectively). However, in ICRMc, the water molecule is instead occupied by the phosphoryl group of the bound phosphoethanolamine (PDB entry 6BND), which is further coordinated from the opposite side by the zinc ion (Figure 5B). These findings suggest that His396 is also positioned to be part of a similar proton relay that is important for catalysis in BcsG. While further work to explore this role is warranted, the essential nature of His396 in the previously proposed ping-pong bibi mechanism (25) has been further evidenced by our kinetic and structural results.
Table 1. Measured Michaelis–Menten Parameters for EcBcsG and Its Variants
parameter WT S244Aa H396Aa
kcat (s–1) 4.34 × 10–7 ± 1.30 × 10–8 2.98 × 10–7 ± 2.45 × 10–8 1.64 × 10–7 ± 9.58 × 10–9
KM (mM) 2.40 ± 0.28 7.85 ± 1.30 3.20 ± 0.56
Vmax (nmol s–1) 0.022 0.015 0.008
kcat/KM (M–1 s–1) 1.81 × 10–4 ± 2.81 × 10–5 3.80 × 10–5 ± 9.42 × 10–6 5.13 × 10–5 ± 8.98 × 10–6
a

With 3 mM cellopenatose co-substrate.

Figure 5

Figure 5. Conserved His396 functions in catalysis and not secondary metal binding. The active sites of (A) BcsG (PDB entry 6PD0), (B) IcrMc (PDB entry 6BND), and (C) MCR-1 (PDB entry 5LRM) display a conserved His residue in the proximity of the active site. Although involved in binding a second Zn2+ ion in some structures of MCR-1, this His residue is within H-bonding distance of the adjacent main chain carbonyl in each of the three structures and likely functions in proton relay, supported by kinetic studies of the BcsG H396A variant.

Steady-State Transferase Kinetics

We previously reported that introducing high relative concentrations of cellooligosaccharides with degrees of polymerization (DP) from 4 to 6 significantly increased the specific activity of EcBcsGΔN in a length-dependent manner. (25) Under these conditions, pEtN-modified cellooligosaccharides accumulated over time, demonstrating EcBcsGΔN is capable of transferase activity in vitro. However, the substrate preference of EcBcsGΔN using these cellooligosaccharides remains to be seen. To that end, we performed steady-state kinetics using fixed and saturating concentrations of p-NPPE and varied the concentration of cellotetraose, cellopentaose, and cellohexaose to understand EcBcsGΔN acceptor preference (Table 2).
Table 2. Measured Michaelis–Menten Parameters for EcBcsG Acceptor Co-substrates
parameter cellotetraose cellopentaose cellohexaose
kcat (s–1) 2.00 × 10–6 ± 1.42 × 10–7 3.80 × 10–6 ± 2.99 × 10–7 3.27 × 10–6 ± 1.97 × 10–7
KM (mM) 3.20 ± 0.47 3.08 ± 0.53 1.76 ± 0.21
Vmax (nmol s–1) 0.10 0.19 0.16
kcat/KM (M–1 s–1) (6.24 ± 1.41) × 10–4 1.23 × 10–3 ± 2.96 × 10–4 1.86 × 10–3 ± 3.18 × 10–4
As our previous results suggested, EcBcsGΔN displays an increasing affinity for cellooligosaccharides with an increasing DP, although due to limited solubility, cellohexaose represents the highest DP that can be assayed using our experimental design (Figure 4B). We measured apparent KM values of 3.20 ± 0.47, 3.08 ± 0.53, and 1.76 ± 0.21 mM for celloligosaccharides with DP values of 4, 5, and 6, respectively (Figure 4B and Table 2). A similar trend in calculated kcat values was observed with values of 2.00 × 10–6 ± 1.41 × 10–7, 3.80 × 10–6 ± 3.00 × 10–7, and 3.27 × 10–6 ± 1.97 × 10–7 s–1, respectively, and the resulting catalytic efficiencies (kcat/KM) were then calculated to be (6.24 ± 1.41) × 10–4, 1.23 × 10–3 ± 2.96 × 10–4, and 1.86 × 10–3 ± 3.18 × 10–4 M–1 s–1 for DP values of 4, 5, and 6, respectively (Table 2). In the case of cellohexaose, we measured an order of magnitude increase in catalytic efficiency in the presence of the acceptor (i.e., 1.86 × 10–3 ± 3.18 × 10–4 compared to 1.81 × 10–4 ± 2.81 × 10–5 M–1 s–1 without an acceptor), suggesting that the enzyme is far more active as a transferase than as an esterase, at least with the artificial substrate p-NPPE.
Measurement of the kinetic parameters of other pEtN transferases has not been reported at present, likely due to the absence of commercially available substrate analogues with which to do so. Accordingly, a comparison between the kinetic parameters of EcBcsGΔN measured here and those of other pEtN transferases is not possible. However, the measured kcat and KM values we report are poor in comparison to those of other biochemically characterized enzymes of virtually any identity; a majority of enzymes have reported kcat (s–1) values of >1 and catalytic efficiencies of at least 1.0 × 103 M–1 s–1. (45) However, when examined against characterized enzymes that modify other bacterial exopolysaccharides, the BcsG values are more typical. For example, the PNAG de-N-acetylases PgaB from E. coli and IcaB from Staphylococcus epidermidis have low reported kcat (PgaB, 0.0013 s–1; IcaB, 0.0007 s–1) and kcat/KM values (PgaB, 0.26 M–1 s–1; IcaB, 0.03 M–1 s–1), using a similar assay design with a synthetic fluorogenic substrate 4-methylumbelliferyl acetate. (43,46) In addition to PNAG, the alginate epimerase AlgG from Pseudomonas aeruginosa (kcat/KM ∼ 0.02 M–1 s–1) is also consistent with EcBcsGΔN. (47) It is worth noting that the analyses mentioned above were all performed with artificial co-substrates that are not present in a cellular context, which may also explain in part the catalytic inefficiencies of these enzymes to act upon them. Regardless, the generally low turnover numbers found in exopolysaccharide-modifying enzymes have been rationalized previously by the partial degree of modification achieved in the mature polymer, (39,43) which we mentioned above. Accordingly, only 50% of the saccharide units are C6-pEtN substituted in the mature pEtN cellulose polymer, (14) which is in better agreement with the low EcBcsGΔN turnover number and catalytic efficiency.
The rate of cellulose synthesis by BcsAB has been measured, and surprisingly, Omadjela and co-workers reported a rate for BcsAB that is orders of magnitude greater than that of BcsG. (18) However, they similarly acknowledge for their in vitro assay, as we did above for BcsG, that these assays can fail to capture nuances of the biological complexity of the Bcs macrocomplex and that the rate of cellulose synthesis that they report for BcsAB is probably much lower in vivo. Direct rate comparisons between these disparately designed assays should also be interpreted with caution. However, it should be noted that the native E. coli Bcs macrocomplex has been resolved with two copies of BcsG associated with a single BcsA catalytic subunit. (23) Taken together, these observations may partly reconcile the differences in rates between BcsG and BcsA in such a way as to generate the level of pEtN modification required for biofilm formation in vivo.

Concluding Remarks

ARTICLE SECTIONS
Jump To

We have demonstrated that full EcBcsGΔN activity in vitro is dependent upon Zn2+ binding, the putative catalytic nucleophile and acid–base arrangement (Ser278/Cys243/His396), the previously unidentified active site residues Arg458 and Ser244, and proper formation of a disulfide bond between Cys290 and Cys306. A kinetic analysis of EcBcsGΔN demonstrated that it is specific for the transfer of phosphoethanolamine over other donor substrates tested. Our data also showed that EcBcsG shares equivalent catalytic residues with the characterized colistin resistance enzymes, although our model does not support a two-Zn2+ model for catalysis that was suggested for one other group member, MCR-1. Instead, our data suggest a shared mono-Zn2+ mechanism for the pEtN transferase family and that BcsG possesses an active site similar to, but distinct from, that of the colistin resistance enzymes, thereby pointing to new roles for the family.
On the basis of our current data, we propose a catalytic mechanism for transfer of phosphoethanolamine to cellulose by BcsG that involves activation of Ser278 by abstraction of the hydroxyl proton by His443 and Glu442 (Figure 6). Following a nucleophilic attack of the phosphate, the loss of the diacylglycerol product is probably achieved by protonation of the oxyanion intermediate, stabilized by the Zn2+ ion, possibly through Cys243 and His396. Although a catalytic Cys base is unprecedented in the pEtN family, the Glu442/His443 acid–base pair would be too distant to serve this role, separated by at least 5 Å. Our data presented herein support a catalytic role for Cys243, and its presence in place of the typical glutamic acid may be rationalized by the lower overall catalytic efficiency required of BcsG in general. The essential substrate recognition and catalytic features observed thus provide a molecular level understanding of BcsG and will facilitate the design of biofilm inhibitors targeting this enzyme.

Figure 6

Figure 6. Proposed mechanism of BcsG as a pEtN transferase. In the first mechanistic step, the hydroxyl proton is abstracted from Ser278 by His443 and Glu442 (2). Protonation and collapse of the resulting oxyanion intermediate are plausibly achieved by His396 and Cys243, resulting in the covalent enzyme intermediate and release of the diacylglycerol co-product (3). In the second mechanistic step, the C6 hydroxyl of cellulose is deprotonated, possibly by Cys243, and attacks the phosphoryl-Ser278 intermediate. The resulting Ser278 oxyanion is protonated, likely by the adjacent His443, followed by release of the pEtN cellulose co-product (4). Our enzymatic data also suggest that Ser244 and Arg458 serve as important polar contacts during both mechanistic steps, especially Arg458 in the second mechanistic step.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.1c00605.

  • A complete list of plasmids, primers, and strains used in this study, the synthesis of p-NPPP, and representative 1H, 13C, and 31P NMR shifts for p-NPPP (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
  • Authors
    • Alexander C. Anderson - Department of Biology and , Wilfrid Laurier University, Waterloo, ON N2L3C5, CanadaPresent Address: A.C.A.: Department of Molecular and Cellular Biology, University of Guelph, Guelph, ON N1G2W1, CanadaOrcidhttps://orcid.org/0000-0002-1870-2903
    • Alysha J. N. Burnett - Department of Biology and , Wilfrid Laurier University, Waterloo, ON N2L3C5, Canada
    • Shirley Constable - Department of Biology and , Wilfrid Laurier University, Waterloo, ON N2L3C5, Canada
    • Lana Hiscock - Department of Biology  and  Department of Chemistry & Biochemistry and , Wilfrid Laurier University, Waterloo, ON N2L3C5, Canada
    • Kenneth E. Maly - Department of Chemistry & Biochemistry and , Wilfrid Laurier University, Waterloo, ON N2L3C5, CanadaOrcidhttps://orcid.org/0000-0002-3695-4995
  • Author Contributions

    J.T.W. and A.C.A. conceived the research and acquired funding. A.C.A. and A.J.N.B. designed and carried out the enzymology experiments and prepared the manuscript. L.H. and K.E.M. carried out the synthesis and validation of p-NPPE and p-NPPP. S.C. carried out some of the mutant-based enzymology. J.T.W., A.C.A., A.J.N.B., S.C., L.H., and K.E.M. were involved in manuscript review and editing. All authors have given approval to the final version of the manuscript.

  • Funding

    The authors acknowledge the support of the Natural Sciences and Engineering Research Council of Canada (NSERC) in the form of a grant to J.T.W. (229971) and a graduate scholarship (PGS-D) to A.C.A. A.C.A. and A.J.N.B. were also supported by Ontario Graduate Scholarships via Wilfrid Laurier University.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

The authors thank D. Brewer and A. Charchoglyan at the University of Guelph for expert technical assistance with mass spectrometry experiments.

Abbreviations

ARTICLE SECTIONS
Jump To

pEtN

phosphoethanolamine

c-di-GMP

cyclic dimeric guanosine monophosphate

TPR

tetratricopeptide repeat

p-NPPE

p-nitrophenyl phosphoethanolamine

p-NPPP

p-nitrophenyl phosphopropanolamine

p-NPP

p-nitrophenyl phosphate

Ni-NTA

nickel-nitrilotriacetic acid

IPTG

isopropyl β-d-thiogalactopyranoside

EDTA

ethylenediaminetetraacetic acid

DTT

dithiothreitol

DP

degree of polymerization.

References

ARTICLE SECTIONS
Jump To

This article references 47 other publications.

  1. 1
    Mah, T.-F. C.; O’Toole, G. A. Mechanisms of Biofilm Resistance to Antimicrobial Agents. Trends Microbiol. 2001, 9 (1), 3439,  DOI: 10.1016/S0966-842X(00)01913-2
  2. 2
    Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S. A.; Kjelleberg, S. Biofilms: An Emergent Form of Bacterial Life. Nat. Rev. Microbiol. 2016, 14 (9), 563575,  DOI: 10.1038/nrmicro.2016.94
  3. 3
    Marsh, P. Dental Plaque as a Microbial Biofilm Dental Plaque – Existing Perspective. Caries Res. 2004, 38, 204211,  DOI: 10.1159/000077756
  4. 4
    Flickinger, S. T.; Copeland, M. F.; Downes, E. M.; Braasch, A. T.; Tuson, H. H.; Eun, Y.-J.; Weibel, D. B. Quorum Sensing between Pseudomonas Aeruginosa Biofilms Accelerates Cell Growth. J. Am. Chem. Soc. 2011, 133 (15), 59665975,  DOI: 10.1021/ja111131f
  5. 5
    Fux, C. A.; Costerton, J. W.; Stewart, P. S.; Stoodley, P. Survival Strategies of Infectious Biofilms. Trends Microbiol. 2005, 13 (1), 3440,  DOI: 10.1016/j.tim.2004.11.010
  6. 6
    Römling, U.; Bokranz, W.; Rabsch, W.; Zogaj, X.; Nimtz, M.; Tschäpe, H. Occurrence and Regulation of the Multicellular Morphotype in Salmonella Serovars Important in Human Disease. Int. J. Med. Microbiol. 2003, 293 (4), 273285,  DOI: 10.1078/1438-4221-00268
  7. 7
    Saldaña, Z.; Xicohtencatl-Cortes, J.; Avelino, F.; Phillips, A. D.; Kaper, J. B.; Puente, J. L.; Girón, J. A. Synergistic Role of Curli and Cellulose in Cell Adherence and Biofilm Formation of Attaching and Effacing Escherichia Coli and Identification of Fis as a Negative Regulator of Curli. Environ. Microbiol. 2009, 11 (4), 9921006,  DOI: 10.1111/j.1462-2920.2008.01824.x
  8. 8
    Spiers, A. J.; Bohannon, J.; Gehrig, S. M.; Rainey, P. B. Biofilm Formation at the Air-Liquid Interface by the Pseudomonas Fluorescens SBW25 Wrinkly Spreader Requires an Acetylated Form of Cellulose. Mol. Microbiol. 2003, 50 (1), 1527,  DOI: 10.1046/j.1365-2958.2003.03670.x
  9. 9
    Hu, L.; Grim, C. J.; Franco, A. A.; Jarvis, K. G.; Sathyamoorthy, V.; Kothary, M. H.; McCardell, B. A.; Tall, B. D. Analysis of the Cellulose Synthase Operon Genes, BcsA, BcsB, and BcsC in Cronobacter Species: Prevalence among Species and Their Roles in Biofilm Formation and Cell-Cell Aggregation. Food Microbiol. 2015, 52, 97105,  DOI: 10.1016/j.fm.2015.07.007
  10. 10
    Ross, P.; Mayer, R.; Benziman, M. Cellulose Biosynthesis and Function in Bacteria. Microbiol. Rev. 1991, 55 (1), 3558,  DOI: 10.1128/mr.55.1.35-58.1991
  11. 11
    Canale-Parola, E.; Borasky, R.; Wolfe, R. S. Studies on Sarcina Ventriculi III. Localization of Cellulose. J. Bacteriol. 1961, 81 (2), 311318,  DOI: 10.1128/jb.81.2.311-318.1961
  12. 12
    Scott, W.; Lowrance, B.; Anderson, A. C.; Weadge, J. T. Identification of the Clostridial Cellulose Synthase and Characterization of the Cognate Glycosyl Hydrolase, CcsZ. PLoS One 2020, 15 (12), e0242686,  DOI: 10.1371/journal.pone.0242686
  13. 13
    Whitfield, G. B.; Marmont, L. S.; Howell, P. L. Enzymatic Modifications of Exopolysaccharides Enhance Bacterial Persistence. Front. Microbiol. 2015, 6, 471,  DOI: 10.3389/fmicb.2015.00471
  14. 14
    Thongsomboon, W.; Serra, D. O.; Possling, A.; Hadjineophytou, C.; Hengge, R.; Cegelski, L. Phosphoethanolamine Cellulose: A Naturally Produced Chemically Modified Cellulose. Science (Washington, DC, U. S.) 2018, 359 (6373), 334338,  DOI: 10.1126/science.aao4096
  15. 15
    Whitfield, G. B.; Marmont, L. S.; Bundalovic-Torma, C.; Razvi, E.; Roach, E. J.; Khursigara, C. M.; Parkinson, J.; Howell, P. L. Discovery and Characterization of a Gram- Positive Pel Polysaccharide Biosynthetic Gene Cluster. PLoS Pathog. 2020, 16 (4), e1008281,  DOI: 10.1371/journal.ppat.1008281
  16. 16
    Römling, U. Molecular Biology of Cellulose Production in Bacteria. Res. Microbiol. 2002, 153 (4), 205212,  DOI: 10.1016/S0923-2508(02)01316-5
  17. 17
    Römling, U.; Galperin, M. Y. Bacterial Cellulose Biosynthesis: Diversity of Operons, Subunits, Products, and Functions. Trends Microbiol. 2015, 23 (9), 545557,  DOI: 10.1016/j.tim.2015.05.005
  18. 18
    Omadjela, O.; Narahari, A.; Strumillo, J.; Mélida, H.; Mazur, O.; Bulone, V.; Zimmer, J. BcsA and BcsB Form the Catalytically Active Core of Bacterial Cellulose Synthase Sufficient for in Vitro Cellulose Synthesis. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (44), 1785617861,  DOI: 10.1073/pnas.1314063110
  19. 19
    Morgan, J. L. W.; McNamara, J. T.; Zimmer, J. Mechanism of Activation of Bacterial Cellulose Synthase by Cyclic Di-GMP. Nat. Struct. Mol. Biol. 2014, 21 (5), 489496,  DOI: 10.1038/nsmb.2803
  20. 20
    Morgan, J. L. W.; Strumillo, J.; Zimmer, J. Crystallographic Snapshot of Cellulose Synthesis and Membrane Translocation. Nature 2013, 493 (7431), 181186,  DOI: 10.1038/nature11744
  21. 21
    Whitfield, C.; Mainprize, I. L. TPR Motifs: Hallmarks of a New Polysaccharide Export Scaffold. Structure 2010, 18 (2), 151153,  DOI: 10.1016/j.str.2010.01.006
  22. 22
    Low, K. E.; Howell, P. L. Gram-Negative Synthase-Dependent Exopolysaccharide Biosynthetic Machines. Curr. Opin. Struct. Biol. 2018, 53, 3244,  DOI: 10.1016/j.sbi.2018.05.001
  23. 23
    Acheson, J. F.; Derewenda, Z. S.; Zimmer, J. Architecture of the Cellulose Synthase Outer Membrane Channel and Its Association with the Periplasmic TPR Domain. Structure 2019, 27 (12), 18551861.e3,  DOI: 10.1016/j.str.2019.09.008
  24. 24
    Nojima, S.; Fujishima, A.; Kato, K.; Ohuchi, K.; Shimizu, N.; Yonezawa, K.; Tajima, K.; Yao, M. Crystal Structure of the Flexible Tandem Repeat Domain of Bacterial Cellulose Synthesis Subunit C. Sci. Rep. 2017, 7 (1), 13018,  DOI: 10.1038/s41598-017-12530-0
  25. 25
    Anderson, A. C.; Burnett, A. J. N.; Hiscock, L.; Maly, K. E.; Weadge, J. T. The Escherichia Coli Cellulose Synthase Subunit G (BcsG) Is a Zn2+-Dependent Phosphoethanolamine Transferase. J. Biol. Chem. 2020, 295 (18), 62256235,  DOI: 10.1074/jbc.RA119.011668
  26. 26
    Mazur, O.; Zimmer, J. Apo- and Cellopentaose-Bound Structures of the Bacterial Cellulose Synthase Subunit BcsZ. J. Biol. Chem. 2011, 286 (20), 1760117606,  DOI: 10.1074/jbc.M111.227660
  27. 27
    Sun, L.; Vella, P.; Schnell, R.; Polyakova, A.; Bourenkov, G.; Li, F.; Cimdins, A.; Schneider, T. R.; Lindqvist, Y.; Galperin, M. Y.; Schneider, G.; Römling, U. Structural and Functional Characterization of the BcsG Subunit of the Cellulose Synthase in Salmonella Typhimurium. J. Mol. Biol. 2018, 430 (18), 31703189,  DOI: 10.1016/j.jmb.2018.07.008
  28. 28
    Fang, X.; Ahmad, I.; Blanka, A.; Schottkowski, M.; Cimdins, A.; Galperin, M. Y.; Römling, U.; Gomelsky, M. GIL, a New C-di-GMP-binding Protein Domain Involved in Regulation of Cellulose Synthesis in Enterobacteria. Mol. Microbiol. 2014, 93 (3), 439452,  DOI: 10.1111/mmi.12672
  29. 29
    Zouhir, S.; Abidi, W.; Caleechurn, M.; Krasteva, P. V. Structure and Multitasking of the C-Di-GMP-Sensing Cellulose Secretion Regulator BcsE. mBio 2020, 11, 4,  DOI: 10.1128/mBio.01303-20
  30. 30
    Acheson, J. F.; Ho, R.; Goularte, N. F.; Cegelski, L.; Zimmer, J. Molecular Organization of the E. Coli Cellulose Synthase Macrocomplex. Nat. Struct. Mol. Biol. 2021, 28, 310,  DOI: 10.1038/s41594-021-00569-7
  31. 31
    Abidi, W.; Zouhir, S.; Caleechurn, M.; Roche, S.; Krasteva, P. V. Architecture and Regulation of an Enterobacterial Cellulose Secretion System. Sci. Adv. 2021, 7 (5), 116,  DOI: 10.1126/sciadv.abd8049
  32. 32
    Hinchliffe, P.; Yang, Q. E.; Portal, E.; Young, T.; Li, H.; Tooke, C. L.; Carvalho, M. J.; Paterson, N. G.; Brem, J.; Niumsup, P. R.; Tansawai, U.; Lei, L.; Li, M.; Shen, Z.; Wang, Y.; Schofield, C. J.; Mulholland, A. J.; Shen, J.; Fey, N.; Walsh, T. R.; Spencer, J. Insights into the Mechanistic Basis of Plasmid-Mediated Colistin Resistance from Crystal Structures of the Catalytic Domain of MCR-1. Sci. Rep. 2017, 7, 39392,  DOI: 10.1038/srep39392
  33. 33
    McCall, K. A.; Huang, C.; Fierke, C. A. Function and Mechanism of Zinc Metalloenzymes. J. Nutr. 2000, 130 (5), 1437S1446S,  DOI: 10.1093/jn/130.5.1437S
  34. 34
    Fage, C. D.; Brown, D. B.; Boll, J. M.; Keatinge-Clay, A. T.; Trent, M. S. Crystallographic Study of the Phosphoethanolamine Transferase EptC Required for Polymyxin Resistance and Motility in Campylobacter Jejuni. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2014, 70 (10), 27302739,  DOI: 10.1107/S1399004714017623
  35. 35
    Wanty, C.; Anandan, A.; Piek, S.; Walshe, J.; Ganguly, J.; Carlson, R. W.; Stubbs, K. A.; Kahler, C. M.; Vrielink, A. The Structure of the Neisserial Lipooligosaccharide Phosphoethanolamine Transferase A (LptA) Required for Resistance to Polymyxin. J. Mol. Biol. 2013, 425 (18), 33893402,  DOI: 10.1016/j.jmb.2013.06.029
  36. 36
    Anandan, A.; Evans, G. L.; Condic-Jurkic, K.; O’Mara, M. L.; John, C. M.; Phillips, N. J.; Jarvis, G. A.; Wills, S. S.; Stubbs, K. A.; Moraes, I.; Kahler, C. M.; Vrielink, A. Structure of a Lipid A Phosphoethanolamine Transferase Suggests How Conformational Changes Govern Substrate Binding. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (9), 22182223,  DOI: 10.1073/pnas.1612927114
  37. 37
    Moynihan, M. M.; Murkin, A. S. Cysteine Is the General Base That Serves in Catalysis by Isocitrate Lyase and in Mechanism-Based Inhibition by 3-Nitropropionate. Biochemistry 2014, 53 (1), 178187,  DOI: 10.1021/bi401432t
  38. 38
    Zhao, Y.; Meng, Q.; Lai, Y.; Wang, L.; Zhou, D.; Dou, C.; Gu, Y.; Nie, C.; Wei, Y.; Cheng, W. Structural and Mechanistic Insights into Polymyxin Resistance Mediated by EptC Originating from Escherichia Coli. FEBS J. 2019, 286 (4), 750764,  DOI: 10.1111/febs.14719
  39. 39
    Arciola, C. R.; Campoccia, D.; Ravaioli, S.; Montanaro, L. Polysaccharide Intercellular Adhesin in Biofilm: Structural and Regulatory Aspects. Front. Cell. Infect. Microbiol. 2015, 5, 7,  DOI: 10.3389/fcimb.2015.00007
  40. 40
    Little, D. J.; Milek, S.; Bamford, N. C.; Ganguly, T.; Difrancesco, B. R.; Nitz, M.; Deora, R.; Howell, P. L. The Protein BpsB Is a Poly-β-1,6-N-Acetyl-D-Glucosamine Deacetylase Required for Biofilm Formation in Bordetella Bronchiseptica. J. Biol. Chem. 2015, 290 (37), 2282722840,  DOI: 10.1074/jbc.M115.672469
  41. 41
    Little, D. J.; Poloczek, J.; Whitney, J. C.; Robinson, H.; Nitz, M.; Howell, P. L. The Structure- and Metal-Dependent Activity of Escherichia Coli PgaB Provides Insight into the Partial de-N-Acetylation of Poly-β-1,6-N-Acetyl- D-Glucosamine. J. Biol. Chem. 2012, 287 (37), 3112631137,  DOI: 10.1074/jbc.M112.390005
  42. 42
    Little, D. J.; Li, G.; Ing, C.; DiFrancesco, B. R.; Bamford, N. C.; Robinson, H.; Nitz, M.; Pomes, R.; Howell, P. L. Modification and Periplasmic Translocation of the Biofilm Exopolysaccharide Poly- −1,6-N-Acetyl-D-Glucosamine. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (30), 1101311018,  DOI: 10.1073/pnas.1406388111
  43. 43
    Pokrovskaya, V.; Poloczek, J.; Little, D. J.; Griffiths, H.; Howell, P. L.; Nitz, M. Functional Characterization of Staphylococcus Epidermidis IcaB, a De-N-Acetylase Important for Biofilm Formation. Biochemistry 2013, 52 (32), 54635471,  DOI: 10.1021/bi400836g
  44. 44
    Suardíaz, R.; Lythell, E.; Hinchliffe, P.; Van Der Kamp, M.; Spencer, J.; Fey, N.; Mulholland, A. J. Catalytic Mechanism of the Colistin Resistance Protein MCR-1. Org. Biomol. Chem. 2021, 19, 38133819,  DOI: 10.1039/D0OB02566F
  45. 45
    Bar-Even, A.; Noor, E.; Savir, Y.; Liebermeister, W.; Davidi, D.; Tawfik, D. S.; Milo, R. The Moderately Efficient Enzyme: Evolutionary and Physicochemical Trends Shaping Enzyme Parameters. Biochemistry 2011, 50 (21), 44024410,  DOI: 10.1021/bi2002289
  46. 46
    Chibba, A.; Poloczek, J.; Little, D. J.; Howell, P. L.; Nitz, M. Synthesis and Evaluation of Inhibitors of E. Coli PgaB, a Polysaccharide de-N-Acetylase Involved in Biofilm Formation. Org. Biomol. Chem. 2012, 10 (35), 71037107,  DOI: 10.1039/c2ob26105g
  47. 47
    Jerga, A.; Raychaudhuri, A.; Tipton, P. A. Pseudomonas Aeruginosa C5-Mannuronan Epimerase: Steady-State Kinetics and Characterization of the Product. Biochemistry 2006, 45 (2), 552560,  DOI: 10.1021/bi051862l

Cited By

ARTICLE SECTIONS
Jump To

This article has not yet been cited by other publications.

  • Abstract

    Figure 1

    Figure 1. Schematic of the reaction catalyzed by BcsG. Phosphoethanolamine is added to approximately 50% of glucose units of bacterial cellulose, exclusively at the C6 position with an unknown pattern across the polymer. (14,25)

    Figure 2

    Figure 2. Amino acid replacements of essential catalytic residues abrogate EcBcsGΔN activity. (A) The structure of the EcBcsGΔN active site (PDB entry 6PD0) is depicted here and was previously mapped by functional complementation in vivo. (25) (B) The measured rates of esterase (i.e., carbohydrate acceptor-free conditions) and transferase (i.e., in the presence of a cellopentaose acceptor) activities of EcBcsGΔN amino acid variants within the catalytic site demonstrate these residues are essential in vitro with the exception of Y277. Asterisks denote statistical significance [Tukey’s multiple comparisons, q(33) > 8.40 and p < 0.0001 for all comparisons].

    Figure 3

    Figure 3. Zn2+ binding and disulfide bond formation are essential catalytic features of EcBcsGΔN. Treatment of EcBcsGΔN with the metal chelating agent EDTA or the reducing agent DTT results in impaired enzymatic activities, consistent with impaired biofilm-forming phenotypes observed in vivo. Asterisks denote significant differences from respective controls [Tukey’s multiple comparisons, q(14) > 6.5 and p < 0.004 for all comparisons].

    Figure 4

    Figure 4. Steady-state kinetics of EcBcsGΔN. (A) Steady-state esterase (WT) and transferase (mutant) kinetics using p-NPPE as the phosphoethanolamine donor and 3 mM cellopentaose as the acceptor co-substrate. (B) Steady-state transferase kinetics using 7 mM p-NPPE as the donor and cellooligosaccharides with DP values of 4–6 as the acceptor co-substrates.

    Figure 5

    Figure 5. Conserved His396 functions in catalysis and not secondary metal binding. The active sites of (A) BcsG (PDB entry 6PD0), (B) IcrMc (PDB entry 6BND), and (C) MCR-1 (PDB entry 5LRM) display a conserved His residue in the proximity of the active site. Although involved in binding a second Zn2+ ion in some structures of MCR-1, this His residue is within H-bonding distance of the adjacent main chain carbonyl in each of the three structures and likely functions in proton relay, supported by kinetic studies of the BcsG H396A variant.

    Figure 6

    Figure 6. Proposed mechanism of BcsG as a pEtN transferase. In the first mechanistic step, the hydroxyl proton is abstracted from Ser278 by His443 and Glu442 (2). Protonation and collapse of the resulting oxyanion intermediate are plausibly achieved by His396 and Cys243, resulting in the covalent enzyme intermediate and release of the diacylglycerol co-product (3). In the second mechanistic step, the C6 hydroxyl of cellulose is deprotonated, possibly by Cys243, and attacks the phosphoryl-Ser278 intermediate. The resulting Ser278 oxyanion is protonated, likely by the adjacent His443, followed by release of the pEtN cellulose co-product (4). Our enzymatic data also suggest that Ser244 and Arg458 serve as important polar contacts during both mechanistic steps, especially Arg458 in the second mechanistic step.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 47 other publications.

    1. 1
      Mah, T.-F. C.; O’Toole, G. A. Mechanisms of Biofilm Resistance to Antimicrobial Agents. Trends Microbiol. 2001, 9 (1), 3439,  DOI: 10.1016/S0966-842X(00)01913-2
    2. 2
      Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S. A.; Kjelleberg, S. Biofilms: An Emergent Form of Bacterial Life. Nat. Rev. Microbiol. 2016, 14 (9), 563575,  DOI: 10.1038/nrmicro.2016.94
    3. 3
      Marsh, P. Dental Plaque as a Microbial Biofilm Dental Plaque – Existing Perspective. Caries Res. 2004, 38, 204211,  DOI: 10.1159/000077756
    4. 4
      Flickinger, S. T.; Copeland, M. F.; Downes, E. M.; Braasch, A. T.; Tuson, H. H.; Eun, Y.-J.; Weibel, D. B. Quorum Sensing between Pseudomonas Aeruginosa Biofilms Accelerates Cell Growth. J. Am. Chem. Soc. 2011, 133 (15), 59665975,  DOI: 10.1021/ja111131f
    5. 5
      Fux, C. A.; Costerton, J. W.; Stewart, P. S.; Stoodley, P. Survival Strategies of Infectious Biofilms. Trends Microbiol. 2005, 13 (1), 3440,  DOI: 10.1016/j.tim.2004.11.010
    6. 6
      Römling, U.; Bokranz, W.; Rabsch, W.; Zogaj, X.; Nimtz, M.; Tschäpe, H. Occurrence and Regulation of the Multicellular Morphotype in Salmonella Serovars Important in Human Disease. Int. J. Med. Microbiol. 2003, 293 (4), 273285,  DOI: 10.1078/1438-4221-00268
    7. 7
      Saldaña, Z.; Xicohtencatl-Cortes, J.; Avelino, F.; Phillips, A. D.; Kaper, J. B.; Puente, J. L.; Girón, J. A. Synergistic Role of Curli and Cellulose in Cell Adherence and Biofilm Formation of Attaching and Effacing Escherichia Coli and Identification of Fis as a Negative Regulator of Curli. Environ. Microbiol. 2009, 11 (4), 9921006,  DOI: 10.1111/j.1462-2920.2008.01824.x
    8. 8
      Spiers, A. J.; Bohannon, J.; Gehrig, S. M.; Rainey, P. B. Biofilm Formation at the Air-Liquid Interface by the Pseudomonas Fluorescens SBW25 Wrinkly Spreader Requires an Acetylated Form of Cellulose. Mol. Microbiol. 2003, 50 (1), 1527,  DOI: 10.1046/j.1365-2958.2003.03670.x
    9. 9
      Hu, L.; Grim, C. J.; Franco, A. A.; Jarvis, K. G.; Sathyamoorthy, V.; Kothary, M. H.; McCardell, B. A.; Tall, B. D. Analysis of the Cellulose Synthase Operon Genes, BcsA, BcsB, and BcsC in Cronobacter Species: Prevalence among Species and Their Roles in Biofilm Formation and Cell-Cell Aggregation. Food Microbiol. 2015, 52, 97105,  DOI: 10.1016/j.fm.2015.07.007
    10. 10
      Ross, P.; Mayer, R.; Benziman, M. Cellulose Biosynthesis and Function in Bacteria. Microbiol. Rev. 1991, 55 (1), 3558,  DOI: 10.1128/mr.55.1.35-58.1991
    11. 11
      Canale-Parola, E.; Borasky, R.; Wolfe, R. S. Studies on Sarcina Ventriculi III. Localization of Cellulose. J. Bacteriol. 1961, 81 (2), 311318,  DOI: 10.1128/jb.81.2.311-318.1961
    12. 12
      Scott, W.; Lowrance, B.; Anderson, A. C.; Weadge, J. T. Identification of the Clostridial Cellulose Synthase and Characterization of the Cognate Glycosyl Hydrolase, CcsZ. PLoS One 2020, 15 (12), e0242686,  DOI: 10.1371/journal.pone.0242686
    13. 13
      Whitfield, G. B.; Marmont, L. S.; Howell, P. L. Enzymatic Modifications of Exopolysaccharides Enhance Bacterial Persistence. Front. Microbiol. 2015, 6, 471,  DOI: 10.3389/fmicb.2015.00471
    14. 14
      Thongsomboon, W.; Serra, D. O.; Possling, A.; Hadjineophytou, C.; Hengge, R.; Cegelski, L. Phosphoethanolamine Cellulose: A Naturally Produced Chemically Modified Cellulose. Science (Washington, DC, U. S.) 2018, 359 (6373), 334338,  DOI: 10.1126/science.aao4096
    15. 15
      Whitfield, G. B.; Marmont, L. S.; Bundalovic-Torma, C.; Razvi, E.; Roach, E. J.; Khursigara, C. M.; Parkinson, J.; Howell, P. L. Discovery and Characterization of a Gram- Positive Pel Polysaccharide Biosynthetic Gene Cluster. PLoS Pathog. 2020, 16 (4), e1008281,  DOI: 10.1371/journal.ppat.1008281
    16. 16
      Römling, U. Molecular Biology of Cellulose Production in Bacteria. Res. Microbiol. 2002, 153 (4), 205212,  DOI: 10.1016/S0923-2508(02)01316-5
    17. 17
      Römling, U.; Galperin, M. Y. Bacterial Cellulose Biosynthesis: Diversity of Operons, Subunits, Products, and Functions. Trends Microbiol. 2015, 23 (9), 545557,  DOI: 10.1016/j.tim.2015.05.005
    18. 18
      Omadjela, O.; Narahari, A.; Strumillo, J.; Mélida, H.; Mazur, O.; Bulone, V.; Zimmer, J. BcsA and BcsB Form the Catalytically Active Core of Bacterial Cellulose Synthase Sufficient for in Vitro Cellulose Synthesis. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (44), 1785617861,  DOI: 10.1073/pnas.1314063110
    19. 19
      Morgan, J. L. W.; McNamara, J. T.; Zimmer, J. Mechanism of Activation of Bacterial Cellulose Synthase by Cyclic Di-GMP. Nat. Struct. Mol. Biol. 2014, 21 (5), 489496,  DOI: 10.1038/nsmb.2803
    20. 20
      Morgan, J. L. W.; Strumillo, J.; Zimmer, J. Crystallographic Snapshot of Cellulose Synthesis and Membrane Translocation. Nature 2013, 493 (7431), 181186,  DOI: 10.1038/nature11744
    21. 21
      Whitfield, C.; Mainprize, I. L. TPR Motifs: Hallmarks of a New Polysaccharide Export Scaffold. Structure 2010, 18 (2), 151153,  DOI: 10.1016/j.str.2010.01.006
    22. 22
      Low, K. E.; Howell, P. L. Gram-Negative Synthase-Dependent Exopolysaccharide Biosynthetic Machines. Curr. Opin. Struct. Biol. 2018, 53, 3244,  DOI: 10.1016/j.sbi.2018.05.001
    23. 23
      Acheson, J. F.; Derewenda, Z. S.; Zimmer, J. Architecture of the Cellulose Synthase Outer Membrane Channel and Its Association with the Periplasmic TPR Domain. Structure 2019, 27 (12), 18551861.e3,  DOI: 10.1016/j.str.2019.09.008
    24. 24
      Nojima, S.; Fujishima, A.; Kato, K.; Ohuchi, K.; Shimizu, N.; Yonezawa, K.; Tajima, K.; Yao, M. Crystal Structure of the Flexible Tandem Repeat Domain of Bacterial Cellulose Synthesis Subunit C. Sci. Rep. 2017, 7 (1), 13018,  DOI: 10.1038/s41598-017-12530-0
    25. 25
      Anderson, A. C.; Burnett, A. J. N.; Hiscock, L.; Maly, K. E.; Weadge, J. T. The Escherichia Coli Cellulose Synthase Subunit G (BcsG) Is a Zn2+-Dependent Phosphoethanolamine Transferase. J. Biol. Chem. 2020, 295 (18), 62256235,  DOI: 10.1074/jbc.RA119.011668
    26. 26
      Mazur, O.; Zimmer, J. Apo- and Cellopentaose-Bound Structures of the Bacterial Cellulose Synthase Subunit BcsZ. J. Biol. Chem. 2011, 286 (20), 1760117606,  DOI: 10.1074/jbc.M111.227660
    27. 27
      Sun, L.; Vella, P.; Schnell, R.; Polyakova, A.; Bourenkov, G.; Li, F.; Cimdins, A.; Schneider, T. R.; Lindqvist, Y.; Galperin, M. Y.; Schneider, G.; Römling, U. Structural and Functional Characterization of the BcsG Subunit of the Cellulose Synthase in Salmonella Typhimurium. J. Mol. Biol. 2018, 430 (18), 31703189,  DOI: 10.1016/j.jmb.2018.07.008
    28. 28
      Fang, X.; Ahmad, I.; Blanka, A.; Schottkowski, M.; Cimdins, A.; Galperin, M. Y.; Römling, U.; Gomelsky, M. GIL, a New C-di-GMP-binding Protein Domain Involved in Regulation of Cellulose Synthesis in Enterobacteria. Mol. Microbiol. 2014, 93 (3), 439452,  DOI: 10.1111/mmi.12672
    29. 29
      Zouhir, S.; Abidi, W.; Caleechurn, M.; Krasteva, P. V. Structure and Multitasking of the C-Di-GMP-Sensing Cellulose Secretion Regulator BcsE. mBio 2020, 11, 4,  DOI: 10.1128/mBio.01303-20
    30. 30
      Acheson, J. F.; Ho, R.; Goularte, N. F.; Cegelski, L.; Zimmer, J. Molecular Organization of the E. Coli Cellulose Synthase Macrocomplex. Nat. Struct. Mol. Biol. 2021, 28, 310,  DOI: 10.1038/s41594-021-00569-7
    31. 31
      Abidi, W.; Zouhir, S.; Caleechurn, M.; Roche, S.; Krasteva, P. V. Architecture and Regulation of an Enterobacterial Cellulose Secretion System. Sci. Adv. 2021, 7 (5), 116,  DOI: 10.1126/sciadv.abd8049
    32. 32
      Hinchliffe, P.; Yang, Q. E.; Portal, E.; Young, T.; Li, H.; Tooke, C. L.; Carvalho, M. J.; Paterson, N. G.; Brem, J.; Niumsup, P. R.; Tansawai, U.; Lei, L.; Li, M.; Shen, Z.; Wang, Y.; Schofield, C. J.; Mulholland, A. J.; Shen, J.; Fey, N.; Walsh, T. R.; Spencer, J. Insights into the Mechanistic Basis of Plasmid-Mediated Colistin Resistance from Crystal Structures of the Catalytic Domain of MCR-1. Sci. Rep. 2017, 7, 39392,  DOI: 10.1038/srep39392
    33. 33
      McCall, K. A.; Huang, C.; Fierke, C. A. Function and Mechanism of Zinc Metalloenzymes. J. Nutr. 2000, 130 (5), 1437S1446S,  DOI: 10.1093/jn/130.5.1437S
    34. 34
      Fage, C. D.; Brown, D. B.; Boll, J. M.; Keatinge-Clay, A. T.; Trent, M. S. Crystallographic Study of the Phosphoethanolamine Transferase EptC Required for Polymyxin Resistance and Motility in Campylobacter Jejuni. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2014, 70 (10), 27302739,  DOI: 10.1107/S1399004714017623
    35. 35
      Wanty, C.; Anandan, A.; Piek, S.; Walshe, J.; Ganguly, J.; Carlson, R. W.; Stubbs, K. A.; Kahler, C. M.; Vrielink, A. The Structure of the Neisserial Lipooligosaccharide Phosphoethanolamine Transferase A (LptA) Required for Resistance to Polymyxin. J. Mol. Biol. 2013, 425 (18), 33893402,  DOI: 10.1016/j.jmb.2013.06.029
    36. 36
      Anandan, A.; Evans, G. L.; Condic-Jurkic, K.; O’Mara, M. L.; John, C. M.; Phillips, N. J.; Jarvis, G. A.; Wills, S. S.; Stubbs, K. A.; Moraes, I.; Kahler, C. M.; Vrielink, A. Structure of a Lipid A Phosphoethanolamine Transferase Suggests How Conformational Changes Govern Substrate Binding. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (9), 22182223,  DOI: 10.1073/pnas.1612927114
    37. 37
      Moynihan, M. M.; Murkin, A. S. Cysteine Is the General Base That Serves in Catalysis by Isocitrate Lyase and in Mechanism-Based Inhibition by 3-Nitropropionate. Biochemistry 2014, 53 (1), 178187,  DOI: 10.1021/bi401432t
    38. 38
      Zhao, Y.; Meng, Q.; Lai, Y.; Wang, L.; Zhou, D.; Dou, C.; Gu, Y.; Nie, C.; Wei, Y.; Cheng, W. Structural and Mechanistic Insights into Polymyxin Resistance Mediated by EptC Originating from Escherichia Coli. FEBS J. 2019, 286 (4), 750764,  DOI: 10.1111/febs.14719
    39. 39
      Arciola, C. R.; Campoccia, D.; Ravaioli, S.; Montanaro, L. Polysaccharide Intercellular Adhesin in Biofilm: Structural and Regulatory Aspects. Front. Cell. Infect. Microbiol. 2015, 5, 7,  DOI: 10.3389/fcimb.2015.00007
    40. 40
      Little, D. J.; Milek, S.; Bamford, N. C.; Ganguly, T.; Difrancesco, B. R.; Nitz, M.; Deora, R.; Howell, P. L. The Protein BpsB Is a Poly-β-1,6-N-Acetyl-D-Glucosamine Deacetylase Required for Biofilm Formation in Bordetella Bronchiseptica. J. Biol. Chem. 2015, 290 (37), 2282722840,  DOI: 10.1074/jbc.M115.672469
    41. 41
      Little, D. J.; Poloczek, J.; Whitney, J. C.; Robinson, H.; Nitz, M.; Howell, P. L. The Structure- and Metal-Dependent Activity of Escherichia Coli PgaB Provides Insight into the Partial de-N-Acetylation of Poly-β-1,6-N-Acetyl- D-Glucosamine. J. Biol. Chem. 2012, 287 (37), 3112631137,  DOI: 10.1074/jbc.M112.390005
    42. 42
      Little, D. J.; Li, G.; Ing, C.; DiFrancesco, B. R.; Bamford, N. C.; Robinson, H.; Nitz, M.; Pomes, R.; Howell, P. L. Modification and Periplasmic Translocation of the Biofilm Exopolysaccharide Poly- −1,6-N-Acetyl-D-Glucosamine. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (30), 1101311018,  DOI: 10.1073/pnas.1406388111
    43. 43
      Pokrovskaya, V.; Poloczek, J.; Little, D. J.; Griffiths, H.; Howell, P. L.; Nitz, M. Functional Characterization of Staphylococcus Epidermidis IcaB, a De-N-Acetylase Important for Biofilm Formation. Biochemistry 2013, 52 (32), 54635471,  DOI: 10.1021/bi400836g
    44. 44
      Suardíaz, R.; Lythell, E.; Hinchliffe, P.; Van Der Kamp, M.; Spencer, J.; Fey, N.; Mulholland, A. J. Catalytic Mechanism of the Colistin Resistance Protein MCR-1. Org. Biomol. Chem. 2021, 19, 38133819,  DOI: 10.1039/D0OB02566F
    45. 45
      Bar-Even, A.; Noor, E.; Savir, Y.; Liebermeister, W.; Davidi, D.; Tawfik, D. S.; Milo, R. The Moderately Efficient Enzyme: Evolutionary and Physicochemical Trends Shaping Enzyme Parameters. Biochemistry 2011, 50 (21), 44024410,  DOI: 10.1021/bi2002289
    46. 46
      Chibba, A.; Poloczek, J.; Little, D. J.; Howell, P. L.; Nitz, M. Synthesis and Evaluation of Inhibitors of E. Coli PgaB, a Polysaccharide de-N-Acetylase Involved in Biofilm Formation. Org. Biomol. Chem. 2012, 10 (35), 71037107,  DOI: 10.1039/c2ob26105g
    47. 47
      Jerga, A.; Raychaudhuri, A.; Tipton, P. A. Pseudomonas Aeruginosa C5-Mannuronan Epimerase: Steady-State Kinetics and Characterization of the Product. Biochemistry 2006, 45 (2), 552560,  DOI: 10.1021/bi051862l
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.1c00605.

    • A complete list of plasmids, primers, and strains used in this study, the synthesis of p-NPPP, and representative 1H, 13C, and 31P NMR shifts for p-NPPP (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect