research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X

Crystal structure of the multiple antibiotic resistance regulator MarR from Clostridium difficile

aDepartment of Chemistry, Fudan University, 220 Handan Road, Shanghai 200433, People's Republic of China
*Correspondence e-mail: xstan@fudan.edu.cn

Edited by I. Tanaka, Hokkaido University, Japan (Received 3 March 2017; accepted 24 May 2017; online 31 May 2017)

Regulators of multiple antibiotic resistance (MarRs) are key players against toxins in prokaryotes. MarR homologues have been identified in many bacterial and archaeal species which pose daunting antibiotic resistance issues that threaten public health. The continuous prevalence of Clostridium difficile infection (CDI) throughout the world is associated with the abuse of antibiotics, and antibiotic treatments of CDI have limited effect. In the genome of C. difficile strain 630, the marR gene (ID 4913953) encodes a MarR protein. Here, MarR from C. difficile (MarRC.difficile) was subcloned and crystallized for the first time. MarRC.difficile was successfully expressed in Escherichia coli in a soluble form and was purified to near-homogeneity (>95%) by a two-step purification protocol. The structure of MarRC.difficile has been solved at 2.3 Å resolution. The crystal belonged to the monoclinic space group P43212, with unit-cell parameters a = b = 66.569, c = 83.654 Å. The structure reported reveals MarRC.difficile to be a dimer, with each subunit consisting of six α-helices and three antiparallel β-hairpins. MarRC.difficile shows high structural similarity to the MarR proteins from E. coli and Staphylococcus aureus, indicating that MarRC.difficile might be a DNA-binding protein.

1. Introduction

Clostridium difficile is an anaerobic human pathogen that causes acute healthcare-associated diarrhoea. The morbidity and mortality rates of C. difficile infection (CDI) have increased dramatically in Europe and in North America (Heinlen & Ballard, 2010[Heinlen, L. & Ballard, J. D. (2010). Am. J. Med. Sci. 340, 247-252.]); the emergence of C. difficile strains that are resistant to multiple antibiotic agents can complicate prevention programs and potential treatments (Hunt & Ballard, 2013[Hunt, J. J. & Ballard, J. D. (2013). Microbiol. Mol. Biol. Rev. 77, 567-581.]). Many studies have demonstrated that various bacterial species employ MarR homologues to sense and exert resistance against many cellular toxins from the environment or host immune system, including multiple antibiotics, oxidative reagents and disinfectants (Cohen et al., 1993[Cohen, S. P., Levy, S. B., Foulds, J. & Rosner, J. L. (1993). J. Bacteriol. 175, 7856-7862.]; Alekshun & Levy, 1999[Alekshun, M. N. & Levy, S. B. (1999). Trends Microbiol. 7, 410-413.]). The regulator of multiple antibiotic resistance (MarR) in Escherichia coli, a member of the MarR family of regulator proteins, modulates bacterial detoxification in response to diverse antibiotics (Hao et al., 2014[Hao, Z., Lou, H., Zhu, R., Zhu, J., Zhang, D., Zhao, B. S., Zeng, S., Chen, X., Chan, J., He, C. & Chen, P. R. (2014). Nature Chem. Biol. 10, 21-28.]). The transcription factors of the MarR family regulate diverse genes involved in multiple antibiotic resistance, the synthesis of virulence determinants and many other important biological processes (Martin et al., 1995[Martin, R. G., Nyantakyi, P. S. & Rosner, J. L. (1995). J. Bacteriol. 177, 4176-4178.]; Alekshun & Levy, 1997[Alekshun, M. N. & Levy, S. B. (1997). Antimicrob. Agents Chemother. 41, 2067-2075.]; Perera & Grove, 2010[Perera, I. C. & Grove, A. (2010). J. Mol. Cell Biol. 2, 243-254.]). The MarR protein, as a member of the MarR family of multiple antibiotic resistance proteins, is a key global regulator in C. difficile. A link between MarR family proteins and antibiotic resistance has been suggested in previous studies (George & Levy, 1983[George, A. M. & Levy, S. B. (1983). J. Bacteriol. 155, 541-548.]). However, the function of MarR in C. difficile is still unknown. Here, we aim to study the biological function of MarR from the perspective of its crystal structure. Thus, the major work in this article is to report the crystal structure of MarR from C. difficile (MarRC.difficile).

In this study, we solved the crystal structure of MarRC.difficile by molecular replacement. Diffraction data were collected to 2.3 Å resolution. The overall structure indicated that MarRC.difficile is a homodimer, with each subunit consisting of six helical regions and three β-strands. Like other MarR proteins, the helical regions in each subunit contribute to the protein–protein interface in the dimer. An analysis of electrostatic surface potential shows a putative DNA-binding site, as observed in other MarR family proteins. This is the first reported crystal structure of this protein from C. difficile.

2. Materials and methods

2.1. Protein preparation

The gene encoding MarR (annotated in GenBank as CAJ67669.1) was PCR-amplified using C. difficile 630 genomic DNA as template, into which NdeI and EcoRI restriction sites were introduced. The purified PCR product was digested with the corresponding restriction enzymes and ligated with T4 DNA ligase into the pET-28a(+) vector (Novagen). The resulting construct contained a hexahistidine tag at the N-terminus of MarR and a thrombin cleavage site. The constructed plasmid was transformed into E. coli BL21 cells for expression. The E. coli cells were grown in LB medium containing 100 µg ml−1 kanamycin at 37°C to an OD600 of 0.6 before IPTG was added to a final concentration of 0.4 mM. Protein expression was induced at 20°C for 12 h before harvesting. The cell pellets were collected and resuspended in NTA buffer (20 mM Tris–HCl pH 7.6, 200 mM NaCl, 10% glycerol) containing 1 mM phenylmethanesulfonyl fluoride (PMSF). After sonication on ice for 30 min, the cell lysate was spun at 11 000 rev min−1 for 30 min. The clear lysate was filtrated and loaded onto a HisTrap column (5 ml column, GE Healthcare), which had been pre-equilibrated with 50 ml NTA buffer (Hao et al., 2014[Hao, Z., Lou, H., Zhu, R., Zhu, J., Zhang, D., Zhao, B. S., Zeng, S., Chen, X., Chan, J., He, C. & Chen, P. R. (2014). Nature Chem. Biol. 10, 21-28.]), for nickel-affinity chromatography. The MarRC.difficile protein was then eluted with a linear imidazole gradient followed by a further purification step using a HiLoad 16/60 Superdex 200 column (GE Healthcare) equilibrated with buffer A (20 mM Tris–HCl pH 7.6, 200 mM NaCl). The eluted protein was purified to >95% homogeneity as determined by 16% SDS–PAGE analysis (Fig. 1[link]). The protein was collected and concentrated for crystallization screening.

[Figure 1]
Figure 1
Elution profile of MarRC.difficile from a Superdex 200 HiLoad 16/60 gel-filtration column on an ÄKTAexplorer FPLC system. The peak at 82 ml (the flow rate was 1 ml min−1) represents the MarRC.difficile dimer. The apparent molecular weight of the eluting species was calculated using standard protein markers (Gel Filtration LMW Calibration Kit, GE Healthcare). The right panel shows SDS–PAGE analysis of the collected fraction.

Information relating to the production of recombinant MarR is summarized in Table 1[link].

Table 1
Information relating to the production of recombinant MarR

Source organism C. difficile
DNA source C. difficile strain 630 genomic DNA
Forward primer 5′-GCGCGCGGGAATTCCATATGTTGATTAAGACTTTAGATAGTAATATATTAAGAG-3′
Reverse primer 5′-GCCCGGAATTCCTATCTCTTTACTTTAAACCATTCATTTTCTACG-3′
Cloning vector pET-28a(+) (Novagen)
Expression vector pET-28a(+) (Novagen)
Expression host E. coli BL21
Complete amino-acid sequence of the construct produced§ MGSSHHHHHHSSGLVPRGSHMLIKTLDSNILREVGTLSRAVNSINDIKYKELKLQKGQFTFLTRICENPGINLVELSNMLKVDKATTTKAIQKLIKAGYVDKKQDKFDKRGYNLTPTDKSLEVYELIIEEENRSIEICFDNFTDEEKQVVTKLLEKMSKNVENEWFKVKR
†The NdeI site for cloning is underlined.
‡The EcoRI site for cloning is underlined.
§The N-terminal hexahistidine tag, linker and thrombin cleavage site are underlined.

2.2. Crystallization

Crystals of MarRC.difficile were grown at 16°C by hanging-drop vapour diffusion. 2 µl purified protein (5 mg ml−1) in 200 mM NaCl, 20 mM Tris pH 7.6 was mixed with 2 µl reservoir buffer [10%(v/v) 2-propanol, 100 mM Tris pH 7.6]. The droplets were equilibrated against 400 µl reservoir buffer. Crystals were looped-out and soaked in cryoprotectant [crystallization buffer containing 20%(v/v) glycerol] before flash-cooling and storage in liquid nitrogen. Crystallization information is summarized in Table 2[link].

Table 2
Crystallization

Method Vapour diffusion, hanging drop
Plate type 24-well
Temperature (K) 289
Protein concentration (mg ml−1) 5
Buffer composition of protein solution 20 mM Tris pH 7.6, 200 mM NaCl
Composition of reservoir solution 10% 2-propanol, 0.1 M Tris–HCl pH 7.6
Volume and ratio of drop 4 µl, 1:1
Volume of reservoir (ml) 0.4

2.3. Data collection and processing

X-ray diffraction data were collected to 2.3 Å resolution on beamline BL17U at the Shanghai Synchrotron Radiation Facility (SSRF). Crystals were flash-cooled in mother liquor at the beamline before data collection. All data were processed and reduced using HKL-2000 (Otwinowski & Minor, 1997[Otwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307-326.]). The space group of the MarR crystals was determined to be P43212, with one molecule in the asymmetric unit and with unit-cell parameters a = b = 66.57, c = 83.65 Å, α = β = γ = 90° for the native protein. Data-collection and processing statistics are summarized in Table 3[link].

Table 3
Data collection and processing

Diffraction source Beamline BL17U, SSRF
Wavelength (Å) 0.988
Temperature (K) 100
Detector ADSC 315r
Crystal-to-detector distance (mm) 250
Rotation range per image (°) 1
Total rotation range (°) 180
Exposure time per image (s) 1
Space group P43212
a, b, c (Å) 66.57, 66.57, 83.65
α, β, γ (°) 90, 90, 90
Mosaicity (°) 0.520
Resolution range (Å) 50–2.297 (2.34–2.30)
Total No. of reflections 207959
No. of unique reflections 8352
Completeness (%) 93.8 (98.8)
Multiplicity 24.9 (28.5)
I/σ(I)〉 77.0 (16.01)
Rmerge 0.080 (0.530)
Rr.i.m. 0.087 (0.604)
Rp.i.m. 0.019 (0.110)
Overall B factor from Wilson plot (Å2) 48.6

2.4. Structure solution and refinement

The structure was solved by molecular replacement using SlyA from Listeria monocytogenes (PDB entry 4mnu; Midwest Center for Structural Genomics, unpublished work) as the starting model. The structure was refined using REFMAC 5.7.0032 (Winn et al., 2011[Winn, M. D. et al. (2011). Acta Cryst. D67, 235-242.]; Potterton et al., 2003[Potterton, E., Briggs, P., Turkenburg, M. & Dodson, E. (2003). Acta Cryst. D59, 1131-1137.]; Murshudov et al., 2011[Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355-367.]). Finally, the structure was deposited in the Protein Data Bank as PDB entry 5eri. The structure-solution and refinement statistics are summarized in Table 4[link] .

Table 4
Structure solution and refinement

Resolution range (Å) 25.73–2.30 (2.357–2.297)
Completeness (%) 93.8 (98.8)
σ Cutoff I > 3σ(I)
No. of reflections, working set 7915 (586)
No. of reflections, test set 394 (26)
Final Rcryst 0.212 (0.280)
Final Rfree 0.250 (0.318)
Cruickshank DPI 0.337
No. of non-H atoms
 Protein 1246
 Ligand 0
 Water 33
 Total 1279
R.m.s. deviations
 Bonds (Å) 0.019
 Angles (°) 1.789
Average B factors (Å2)
 Protein 59.9
 Water 56.2
Ramachandran plot
 Most favoured (%) 99
 Allowed (%) 1

3. Results and discussion

3.1. Overall structure of MarRC.difficile

The crystal structure of MarRC.difficile was determined by molecular replacement using SlyA (PDB entry 4mnu) as a search model. Like other MarR proteins, the MarRC.difficile protein is composed of six α-helices and a three-stranded antiparallel β-hairpin (Fig. 2[link]a). The α2, α3 and α4 helices and two antiparallel β-strands, β2 and β3, are probably responsible for DNA binding, as indicated by the electrostatic surface potential (Figs. 2[link] and 3[link]). The two putative DNA-binding domains in each subunit result in the formation of a channel through the centre of the dimer (Figs. 2[link] and 3[link]). Consistent with previous studies, the putative DNA-binding regions of the MarRC.difficile protein are strongly electropositive, as are other winged-helix DNA-binding proteins (Gajiwala & Burley, 2000[Gajiwala, K. S. & Burley, S. K. (2000). Curr. Opin. Struct. Biol. 10, 110-116.]).

[Figure 2]
Figure 2
The structure of MarRC.difficile. (a) One MarRC.difficile subunit with labelled secondary structure. (b) Ribbon representation of the crystal structure of the MarRC.difficile dimer viewed with the subunit twofold axis close to vertical.
[Figure 3]
Figure 3
Electrostatic surface representation of the MarRC.difficile dimer. The putative DNA-binding sites are indicated by an arrow.

The structure of MarR from E. coli (MarRE.coli) is a homodimer (Alekshun et al., 2001[Alekshun, M. N., Levy, S. B., Mealy, T. R., Seaton, B. A. & Head, J. F. (2001). Nature Struct. Biol. 8, 710-714.]), and it has been verified that MarRE.coli binds the marRAB promoter as a dimer (Martin et al., 1996[Martin, R. G., Jair, K. W., Wolf, R. E. Jr & Rosner, J. L. (1996). J. Bacteriol. 178, 2216-2223.]). A recent study showed that disulfide bonds could be formed between MarRE.coli dimers, resulting in the dissociation of MarRE.coli from its cognate DNA and enhanced bacterial resistance (Zhu et al., 2017[Zhu, R., Hao, Z., Lou, H., Song, Y., Zhao, J., Chen, Y., Zhu, J. & Chen, P. R. (2017). J. Biol. Inorg. Chem., https://doi.org/10.1007/s00775-017-1442-7.]). The MarR family member MprA also functions as a dimer (Brooun et al., 1999[Brooun, A., Tomashek, J. J. & Lewis, K. (1999). J. Bacteriol. 181, 5131-5133.]). PISA analysis suggested that MarRC.difficile is a homodimer. In the crystal structure of MarRC.difficile there is one monomer in the asymmetric unit, with the dimer being composed of two subunits related by a crystallographic twofold rotation. The crystal structure also indicates that α-helices in the N- and C-terminal regions of each subunit interdigitate with those of the other subunit to form a hydrophobic core burying a surface area of 3100 Å2. Two helical regions, α1 and α6 (residues 9–27 in the N-terminus and residues 123–147 in the C-terminus, respectively), are closely juxtaposed and intertwine with the equivalent regions of the second subunit to form a dimer. The dimer is stabilized by several salt bridges, notably that between Arg16 and Glu74′ and that between Lys155 and Glu145′. In addition, a hydrogen bond between the side-chain carbonyl O atom of Glu66 and the guanidinium NH groups of Arg31′ enhances the stability of the dimeric structure.

3.2. Comparison of MarRC.difficile with MarRE.coli and MgrAS.aureus

MgrA from Staphylococcus aureus (MgrAS.aureus) is a regulator of antibiotic resistance and is also an important virulence determinant during infection (Ingavale et al., 2005[Ingavale, S., van Wamel, W., Luong, T. T., Lee, C. Y. & Cheung, A. L. (2005). Infect. Immun. 73, 1423-1431.]). A previous study indicated that the cysteine residue (Cys12) of this protein could be oxidized by various reactive oxygen species. Cysteine oxidation leads to the dissociation of MgrA from DNA, resulting in the initiation of signalling pathways and further enhancing antibiotic resistance in S. aureus (Chen et al., 2006[Chen, P. R., Bae, T., Williams, W. A., Duguid, E. M., Rice, P. A., Schneewind, O. & He, C. (2006). Nature Chem. Biol. 2, 591-595.]). The oxidation-sensing mechanism is widely used by bacteria to counter challenges of environmental pressure (Lee & Helmann, 2006[Lee, J.-W. & Helmann, J. D. (2006). Nature (London), 440, 363-367.]). In conclusion, MgrA from S. aureus is an oxidation sensor.

Previous studies reported that the MarR family of proteins are typically conserved transcription factors that modulate bacterial resistance to multiple antibiotics, oxidative reagents and detergents (Martin & Rosner, 1995[Martin, R. G. & Rosner, J. L. (1995). Proc. Natl Acad. Sci. USA, 92, 5456-5460.]). Various bacterial species such as E. coli can respond to environmental stresses such as toxic chemicals and disinfectants by triggering the dissociation of MarR from the cognate DNA in a copper-dependent manner. The detailed mechanism is that copper(II) oxidizes a unique cysteine residue (Cys80) that resides in the DNA-binding domain of MarRE.coli to generate inter-dimer disulfide bonds, thereby inducing tetramer formation and the dissociation of MarR from the marRAB promoter (Hao et al., 2014[Hao, Z., Lou, H., Zhu, R., Zhu, J., Zhang, D., Zhao, B. S., Zeng, S., Chen, X., Chan, J., He, C. & Chen, P. R. (2014). Nature Chem. Biol. 10, 21-28.]). Therefore, MarR from E. coli is a copper signal oxidation sensor.

Sequence alignment of MarRC.difficile with MarRE.coli and MgrAS.aureus using MUSCLE shows 26 and 19% sequence identity; these proteins share low sequence similarity (Fig. 4[link]). However, superposition of MarRC.difficile with MgrAS.aureus (PDB entry 2bv6; Chen et al., 2006[Chen, P. R., Bae, T., Williams, W. A., Duguid, E. M., Rice, P. A., Schneewind, O. & He, C. (2006). Nature Chem. Biol. 2, 591-595.]) and MarRE.coli (PDB entry 1jgs; Alekshun et al., 2001[Alekshun, M. N., Levy, S. B., Mealy, T. R., Seaton, B. A. & Head, J. F. (2001). Nature Struct. Biol. 8, 710-714.]) shows structural similarity (Fig. 5[link]). The MgrAS.aureus dimer is triangular in shape, with two winged-helix DNA-binding domains; the DNA-binding domain includes two α-helices, two β-sheets and a wing region (Chen et al., 2006[Chen, P. R., Bae, T., Williams, W. A., Duguid, E. M., Rice, P. A., Schneewind, O. & He, C. (2006). Nature Chem. Biol. 2, 591-595.]). The structure of MarRE.coli is a crystallographic dimer, with each subunit containing a winged-helix DNA-binding motif, and this DNA-binding motif contains three α-helices and two β-strands (Alekshun et al., 2001[Alekshun, M. N., Levy, S. B., Mealy, T. R., Seaton, B. A. & Head, J. F. (2001). Nature Struct. Biol. 8, 710-714.]). We found that the crystal structure reveals MarRC.difficile to be a dimer, with each monomer consisting of six α-helices and a three-stranded antiparallel β-hairpin. The putative DNA-binding domain of each subunit includes three α-helices and two antiparallel β-strands. Correspondingly, MarRE.coli and MgrAS.aureus share a similar oxidation-sensing mechanism in which cysteine oxidation leads to the dissociation of MarRE.coli and MgrAS.aureus from DNA. As a result, they exhibit a similar function. However, the function of MarR in C. difficile remains unknown. Structural analysis of MarRC.difficile indicated that two cysteine residues (Cys45 and Cys117) are located in the hydrophobic core; this may suggest that MarR in C. difficile is probably not an oxidation sensor. Although they share structural similarity, these proteins might have diverse molecular mechanisms.

[Figure 4]
Figure 4
Primary-sequence alignment of MarRC.difficile with representative members of the MarR family (MarRE.coli and MgrAS.aureus). The secondary-structural elements of MarRC.difficile are indicated above the sequence alignment: α-helices (α) are illustrated as curly lines and arrows represent β-sheets (β). Numbering is according to the MarRC.difficile primary sequence. Residues that are identical in all homologues are highlighted in red and highly conserved amino acids are shown in blue boxes. Red arrows indicate the cysteine residues of MarRC.difficile; the key cysteine residues of MarRE.coli and MgrAS.aureus are shown in purple boxes.
[Figure 5]
Figure 5
Superimposition of MarRC.difficile with MarRE.coli and MgrAS.aureus. (a) MarRC.difficile aligned with MarRE.coli (MarRC.difficile, green; MarRE.coli, cyan). (b) MarRC.difficile aligned with MgrAS.aureus (MarRC.difficile, green; MgrAS.aureus, purple).

4. Conclusion

Although the MarR protein has been well studied in many species, the MarR protein from C. difficile remains unknown. The relationship between MarR and antibiotic resistance in C. difficile needs to be investigated. The crystal structure reported in this paper reveals MarRC.difficile to be a crystallo­graphic dimer. It shows structural similarity to other MarR family proteins. Furthermore, the structure of MarR in C. difficile suggests a putative DNA-binding site, revealing that the MarR protein in C. difficile might be also a transcription factor that can bind DNA. Based on the structural analysis of MarRC.difficile, we found that the two cysteine residues could not be oxidized easily as they are located in the hydrophobic core. Therefore, MarR from C. difficile may not be an oxidation sensor. The solved crystal structure of MarRC.difficile will be the first step in further functional studies.

Supporting information


Acknowledgements

We thank the staff of the BL17B/BL18U1/BL19U1/BL19U2/BL01B beamlines of the National Center for Protein Sciences Shanghai (NCPSS) at Shanghai Synchrotron Radiation Facility for assistance during data collection. The authors also thank Shanghai Synchrotron Radiation Facility (SSRF) beamline BL17U for X-ray beam time.

References

First citationAlekshun, M. N. & Levy, S. B. (1997). Antimicrob. Agents Chemother. 41, 2067–2075.  CAS PubMed Web of Science Google Scholar
First citationAlekshun, M. N. & Levy, S. B. (1999). Trends Microbiol. 7, 410–413.  Web of Science CrossRef PubMed CAS Google Scholar
First citationAlekshun, M. N., Levy, S. B., Mealy, T. R., Seaton, B. A. & Head, J. F. (2001). Nature Struct. Biol. 8, 710–714.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBrooun, A., Tomashek, J. J. & Lewis, K. (1999). J. Bacteriol. 181, 5131–5133.  PubMed CAS Google Scholar
First citationChen, P. R., Bae, T., Williams, W. A., Duguid, E. M., Rice, P. A., Schneewind, O. & He, C. (2006). Nature Chem. Biol. 2, 591–595.  Web of Science CrossRef CAS Google Scholar
First citationCohen, S. P., Levy, S. B., Foulds, J. & Rosner, J. L. (1993). J. Bacteriol. 175, 7856–7862.  CrossRef CAS PubMed Web of Science Google Scholar
First citationGajiwala, K. S. & Burley, S. K. (2000). Curr. Opin. Struct. Biol. 10, 110–116.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGeorge, A. M. & Levy, S. B. (1983). J. Bacteriol. 155, 541–548.  CAS PubMed Google Scholar
First citationHao, Z., Lou, H., Zhu, R., Zhu, J., Zhang, D., Zhao, B. S., Zeng, S., Chen, X., Chan, J., He, C. & Chen, P. R. (2014). Nature Chem. Biol. 10, 21–28.  CrossRef CAS Google Scholar
First citationHeinlen, L. & Ballard, J. D. (2010). Am. J. Med. Sci. 340, 247–252.  CrossRef PubMed Google Scholar
First citationHunt, J. J. & Ballard, J. D. (2013). Microbiol. Mol. Biol. Rev. 77, 567–581.  Web of Science CrossRef CAS PubMed Google Scholar
First citationIngavale, S., van Wamel, W., Luong, T. T., Lee, C. Y. & Cheung, A. L. (2005). Infect. Immun. 73, 1423–1431.  CrossRef PubMed CAS Google Scholar
First citationLee, J.-W. & Helmann, J. D. (2006). Nature (London), 440, 363–367.  CrossRef PubMed CAS Google Scholar
First citationMartin, R. G., Jair, K. W., Wolf, R. E. Jr & Rosner, J. L. (1996). J. Bacteriol. 178, 2216–2223.  CrossRef CAS PubMed Google Scholar
First citationMartin, R. G., Nyantakyi, P. S. & Rosner, J. L. (1995). J. Bacteriol. 177, 4176–4178.  CrossRef CAS PubMed Google Scholar
First citationMartin, R. G. & Rosner, J. L. (1995). Proc. Natl Acad. Sci. USA, 92, 5456–5460.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMurshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355–367.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationOtwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307–326.  CrossRef CAS PubMed Web of Science Google Scholar
First citationPerera, I. C. & Grove, A. (2010). J. Mol. Cell Biol. 2, 243–254.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPotterton, E., Briggs, P., Turkenburg, M. & Dodson, E. (2003). Acta Cryst. D59, 1131–1137.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWinn, M. D. et al. (2011). Acta Cryst. D67, 235–242.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhu, R., Hao, Z., Lou, H., Song, Y., Zhao, J., Chen, Y., Zhu, J. & Chen, P. R. (2017). J. Biol. Inorg. Chem., https://doi.org/10.1007/s00775-017-1442-7Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X
Follow Acta Cryst. F
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds