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

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X
Volume 70| Part 9| September 2014| Pages 1199-1205

The structure of alanine racemase from Acinetobacter baumannii

CROSSMARK_Color_square_no_text.svg

aDepartment of Biochemistry, University of Otago, Dunedin, New Zealand
*Correspondence e-mail: kurt.krause@otago.ac.nz

(Received 30 May 2014; accepted 1 August 2014; online 29 August 2014)

Acinetobacter baumannii is an opportunistic Gram-negative bacterium which is a common cause of hospital-acquired infections. Numerous antibiotic-resistant strains exist, emphasizing the need for the development of new antimicrobials. Alanine racemase (Alr) is a pyridoxal 5′-phosphate dependent enzyme that is responsible for racemization between enantiomers of alanine. As D-alanine is an essential component of the bacterial cell wall, its inhibition is lethal to prokaryotes, making it an excellent antibiotic drug target. The crystal structure of A. baumannii alanine racemase (AlrAba) from the highly antibiotic-resistant NCTC13302 strain has been solved to 1.9 Å resolution. Comparison of AlrAba with alanine racemases from closely related bacteria demonstrates a conserved overall fold. The substrate entryway and active site of the enzymes were shown to be highly conserved. The structure of AlrAba will provide the template required for future structure-based drug-design studies.

1. Introduction

Acinetobacter baumannii is a small nonmotile Gram-negative bacterium that is capable of replicating under a wide range of environmental conditions (Peleg et al., 2008[Peleg, A. Y., Seifert, H. & Paterson, D. L. (2008). Clin. Microbiol. Rev. 21, 538-582.]). In vulnerable patients, A. baumannii can cause severe infections such as pneumonia, septicaemia, urinary-tract infections and meningitis (Lee et al., 2007[Lee, N.-Y., Lee, H.-C., Ko, N.-Y., Chang, C.-M., Shih, H.-I., Wu, C.-J. & Ko, W.-C. (2007). Infect. Control Hosp. Epidemiol. 28, 713-719.]). There are numerous strains of A. baumannii that are responsible for endemics and epidemics in hospitals (Scott et al., 2007[Scott, P. et al. (2007). Clin. Infect. Dis. 44, 1577-1584.]; Eliopoulos et al., 2008[Eliopoulos, G. M., Maragakis, L. L. & Perl, T. M. (2008). Clin. Infect. Dis. 46, 1254-1263.]) and many of these strains are antibiotic resistant (Zavascki et al., 2010[Zavascki, A. P., Carvalhaes, C. G., Picão, R. C. & Gales, A. C. (2010). Expert Rev. Anti Infect. Ther. 8, 71-93.]). In fact, multidrug-resistant A. baumannii is considered to be among the most difficult Gram-negative bacteria to control and treat (Dijkshoorn et al., 2007[Dijkshoorn, L., Nemec, A. & Seifert, H. (2007). Nature Rev. Microbiol. 5, 939-951.]). Therefore, new drug therapies are vital in combating these infections (Joshi et al., 2003[Joshi, S. G., Litake, G. M., Niphadkar, K. B. & Ghole, V. S. (2003). J. Infect. Chemother. 9, 187-190.]; Gootz & Marra, 2008[Gootz, T. D. & Marra, A. (2008). Expert Rev. Anti Infect. Ther. 6, 309-325.]).

Alanine racemase (EC 5.1.1.1) is a pyridoxal 5′-phosphate (PLP)-dependent enzyme which catalyses racemization between L-alanine and D-alanine (Walsh, 1989[Walsh, C. T. (1989). J. Biol. Chem. 264, 2393-2396.]). D-Alanine is an essential building block of peptidoglycan, a critical component of the bacterial cell wall (Hols et al., 1997[Hols, P., Defrenne, C., Ferain, T., Derzelle, S., Delplace, B. & Delcour, J. (1997). J. Bacteriol. 179, 3804-3807.]). Alanine racemase (Alr) is the sole source of D-alanine in many bacteria and its inhibition is lethal, making the enzyme an attractive antibiotic drug target (Lambert & Neuhaus, 1972[Lambert, M. P. & Neuhaus, F. C. (1972). J. Bacteriol. 109, 1156-1161.]). Kinetic studies of Alr have indicated that racemization occurs via a stepwise two-base mechanism, in which a highly conserved lysine and a tyrosine residue act as acid and base catalysts. These residues either accept or donate the α-hydrogen of the alanine substrate, depending on the direction of the reaction (Watanabe et al., 2002[Watanabe, A., Yoshimura, T., Mikami, B., Hayashi, H., Kagamiyama, H. & Esaki, N. (2002). J. Biol. Chem. 277, 19166-19172.]; Spies & Toney, 2003[Spies, M. A. & Toney, M. D. (2003). Biochemistry, 42, 5099-5107.]).

Alanine racemase has been structurally characterized from numerous bacteria, including Geobacillus (formerly Bacillus) stearo­thermophilus (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]), Pseudomonas aeruginosa (LeMagueres et al., 2003[LeMagueres, P., Im, H., Dvorak, A., Strych, U., Benedik, M. & Krause, K. L. (2003). Biochemistry, 42, 14752-14761.]), Streptomyces lavendulae (Noda et al., 2004[Noda, M., Matoba, Y., Kumagai, T. & Sugiyama, M. (2004). J. Biol. Chem. 279, 46153-46161.]), Mycobacterium tuberculosis (LeMagueres et al., 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]), Escherichia coli (Wu et al., 2008[Wu, D., Hu, T., Zhang, L., Chen, J., Du, J., Ding, J., Jiang, H. & Shen, X. (2008). Protein Sci. 17, 1066-1076.]), Bacillus anthracis (Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]), Enterococcus faecalis (Priyadarshi et al., 2009[Priyadarshi, A., Lee, E. H., Sung, M. W., Nam, K. H., Lee, W. H., Kim, E. E. & Hwang, K. Y. (2009). Biochim. Biophys. Acta, 1794, 1030-1040.]), Bartonella henselae (Abendroth et al., 2011[Abendroth, J., Gardberg, A. S., Robinson, J. I., Christensen, J. S., Staker, B. L., Myler, P. J., Stewart, L. J. & Edwards, T. E. (2011). J. Struct. Funct. Genomics, 12, 83-95.]), Streptococcus pneumoniae (Im et al., 2011[Im, H., Sharpe, M. L., Strych, U., Davlieva, M. & Krause, K. L. (2011). BMC Microbiol. 11, 116.]), Staphylococcus aureus (Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]) and Oenococcus oeni (Palani et al., 2013[Palani, K., Burley, S. K. & Swaminathan, S. (2013). Acta Cryst. F69, 15-19.]). In these structures, the enzyme is a homodimer in which the two monomers interact in a head-to-tail fashion, generating two active sites. Each Alr monomer has two distinct domains: an α/β-barrel at the N-terminus and a predominantly β-stranded domain at the C-terminus. The PLP cofactor forms an internal aldimine linkage with the highly conserved catalytic lysine residue in the α/β-barrel domain (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]). Here, we report the structure and partial kinetic characterization of A. baumannii alanine racemase (AlrAba) from the multidrug-resistant NCTC13302 strain. A crystallization report concerning alanine racemase from this bacterium has appeared in this journal (Nguyen et al., 2013[Nguyen, D.-D., Ngo, H.-P.-T., Hong, M., Pham, T.-V., Lee, J. H., Lee, J. J., Kwon, D. B., Lee, S. H. & Kang, L.-W. (2013). Acta Cryst. F69, 1041-1044.]), but no structure is available and it was not consulted for this work. The structure presented here will provide a template for future structure-based drug-design studies targeting this important enzyme.

2. Methods

2.1. Gene cloning, expression and protein purification

Prior to cloning, genomic DNA was extracted from A. baumannii strain NCTC13302, after which the alanine racemase alr gene was amplified by PCR using the following primers: 5′-ATGATCTACATATGCGTCAAGCAACAGTT-3′ (forward) and 5′-TATCTCGAG­TTAAGTACCCTGACGGAC-3′ (reverse). Full-length A. baumannii alr (residues 1–356) was cloned into pET-26b, which was transformed into E. coli BL21 (DE3) cells. Cells grown overnight at 310 K were used to inoculate the main culture, which was induced with 0.5 mM IPTG at an OD600 of 0.5. After expression for 16 h, the cell pellet was collected by centrifugation and the cells were lysed via sonication. Following ammonium sulfate cuts of 20 and 60%, the pellet fraction was resuspended in 20 mM Tris pH 8.0 and further purified by hydrophobic interaction, anion-exchange and size-exclusion chromatography. Fractions containing AlrAba were greater than 95% pure as indicated by SDS–PAGE.

2.2. Crystallization

Purified AlrAba was concentrated to 6 mg ml−1 using a Vivaspin 20 (10 000 Da molecular-weight cutoff; GE Healthcare) and sitting drops were set up versus 100 mM 2-(N-morpholino)ethanesulfonic acid (MES) pH 6.0, 200 mM CaCl2, 20% polyethylene glycol 6000 (PEG 6000). Single yellow crystals grew within 4 d and were cryoprotected by soaking in mother liquor containing 20% glycerol. Crystals were flash-cooled in liquid nitrogen prior to data collection.

2.3. Data collection and processing

A native AlrAba data set was collected at the Australian Synchrotron (Melbourne, Australia) on the MX1 beamline equipped with an ADSC Quantum 315r image-plate detector. Data were collected at 13.000 keV using an oscillation angle of 0.5° and an exposure time of 2 s per image. Diffraction images were processed with iMosflm (Battye et al., 2011[Battye, T. G. G., Kontogiannis, L., Johnson, O., Powell, H. R. & Leslie, A. G. W. (2011). Acta Cryst. D67, 271-281.]), POINTLESS (Grosse-Kunstleve et al., 2002[Grosse-Kunstleve, R. W., Sauter, N. K., Moriarty, N. W. & Adams, P. D. (2002). J. Appl. Cryst. 35, 126-136.]) and SCALA (Evans, 2006[Evans, P. (2006). Acta Cryst. D62, 72-82.]) within the CCP4 suite (Winn et al., 2011[Winn, M. D. et al. (2011). Acta Cryst. D67, 235-242.]). The crystals of AlrAba belonged to the monoclinic space group P21, with unit-cell parameters a = 47.0, b = 83.0, c = 93.3 Å, α = 90, β = 97.1, γ = 90°. Data were collected and processed to 1.9 Å resolution with statistics as presented in Table 1[link].

Table 1
Data-collection and refinement statistics

Values in parentheses are for the highest resolution shell.

Space group P21
Unit-cell parameters  
a (Å) 47.0
b (Å) 83.0
c (Å) 93.3
β (°) 97.1
No. of observations 406799 (28844)
No. of unique reflections 55904 (3960)
Completeness (%) 99.7 (99.7)
Rmerge (%) 9.2 (26.0)
I/σ(I)〉 16.9 (8.4)
Multiplicity 7.3 (7.3)
Resolution range (Å) 30.15–1.90 (1.95–1.90)
R factor (%) 19.7
Rfree(%) 23.4
Average B factors (Å2)
 All atoms 14.8
 Main chain 13.3
 Side chains 15.3
 Waters 20.5
R.m.s. deviations
 Bond lengths (Å) 0.016
 Bond angles (°) 1.75
No. of atoms  
 Protein 5360
 PLP 30
 Water 512
Rmerge = [\textstyle \sum_{hkl}\sum_{i}|I_{i}(hkl)- \langle I(hkl)\rangle|/][\textstyle \sum_{hkl}\sum_{i}I_{i}(hkl)].
R factor = [\textstyle \sum_{hkl}\big ||F_{\rm obs}|-][|F_{\rm calc}|\big |/][\textstyle \sum_{hkl}|F_{\rm obs}|].

2.4. Structure determination and refinement

The structure of AlrAba was solved via molecular replacement with Phaser (McCoy et al., 2007[McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658-674.]) using the monomer of P. aeruginosa alanine racemase (with ligands and waters removed), to which it has a sequence identity of 41%, as the search model. This was performed assuming the presence of two monomers in the asymmetric unit, as suggested by the Matthews coefficient VM of 2.07 Å3 Da−1 (Matthews, 1968[Matthews, B. W. (1968). J. Mol. Biol. 33, 491-497.]). PHENIX (Adams et al., 2010[Adams, P. D. et al. (2010). Acta Cryst. D66, 213-221.]) was used to build the initial model and to change the amino-acid sequence to that of AlrAba. After several rounds of model building and refinement using Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]) and refinement using REFMAC5 (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.]), the electron density improved and waters and the cofactor PLP were built into the structure. The final structure had an R factor of 19.7% and an Rfree of 23.4%. The r.m.s. deviations of bond lengths and angles were 0.016 Å and 1.75°, respectively. Intermonomer interactions in the structure were analysed using the Protein Interfaces, Surfaces and Assemblies service (PISA) at the European Bioinformatics Institute (https://www.ebi.ac.uk/pdbe/prot_int/pistart.html ; Krissinel & Henrick, 2007[Krissinel, E. & Henrick, K. (2007). J. Mol. Biol. 372, 774-797.]). Comparisons to other structures were made with the respective PDB files as deposited. The final model was validated using PROCHECK (Laskowski et al., 1993[Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). J. Appl. Cryst. 26, 283-291.]), with the resulting Ramachandran plot indicating that 98.2% of the residues were in the most favoured regions, with 1.6% in additionally allowed regions. Additional structure-determination and refinement statistics are presented in Table 1[link].

2.5. Enzyme kinetics

The D-alanine to L-alanine direction of the racemization catalysed by AlrAba was characterized using a coupled-enzyme assay based on Esaki & Walsh (1986[Esaki, N. & Walsh, C. T. (1986). Biochemistry, 25, 3261-3267.]) and as described in previous studies (Strych et al., 2000[Strych, U., Huang, H.-C., Krause, K. L. & Benedik, M. J. (2000). Curr. Microbiol. 41, 290-294.], 2001[Strych, U., Penland, R. L., Jimenez, M., Krause, K. L. & Benedik, M. J. (2001). FEMS Microbiol. Lett. 196, 93-98.]). This assay was carried out at 30°C and begins with the conversion of D-alanine to L-alanine by alanine racemase. The L-alanine is then converted to pyruvate and ammonia by L-alanine dehydrogenase, producing NADH, which is tracked by following the increase in absorbance at 340 nm. Concentrations of D-alanine ranging from 0.1 to 10 mM were assayed in triplicate using 200 ng AlrAba per reaction. The production of NADH was monitored for 10 min with the slope calculated during the last 2 min of the assay. The kinetic constants Km and Vmax were determined using nonlinear regression fitting carried out within GraphPad Prism v.6 (GraphPad Software, La Jolla, California, USA).

3. Results and discussion

3.1. Overall structure of A. baumannii alanine racemase

The tertiary structure of A. baumannii alanine racemase (AlrAba) is a homodimer in which the two monomers interact in a head-to-tail manner. This results in two active sites per enzyme, each comprised of residues from the N-terminal domain of one monomer and the C-terminal domain of the second monomer (Fig. 1[link]a). Each AlrAba monomer contains two distinct domains (Fig. 1[link]b). The N-terminal domain corresponds to residues 1–230 in the structure and predominantly consists of an eight-stranded α/β-barrel. The C-terminal domain corresponds to residues 231–356 in the structure and mainly contains β-strands (two α-helices and eight β-strands). The individual AlrAba monomers are crystallographically distinct and form a dimer in the asymmetric unit. Following refinement, they have a low r.m.s difference of 0.41 Å after Cα-atom superposition. Overall, the AlrAba structure is lacking clear density for residues 247–263 and 355–356 of monomer B. Missing density for these regions has previously been observed in alanine racemases from other species, including M. tuberculosis (LeMagueres et al., 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]), B. henselae (Abendroth et al., 2011[Abendroth, J., Gardberg, A. S., Robinson, J. I., Christensen, J. S., Staker, B. L., Myler, P. J., Stewart, L. J. & Edwards, T. E. (2011). J. Struct. Funct. Genomics, 12, 83-95.]), S. aureus (Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]) and O. oeni (Palani et al., 2013[Palani, K., Burley, S. K. & Swaminathan, S. (2013). Acta Cryst. F69, 15-19.]).

[Figure 1]
Figure 1
(a) Ribbon representation of the A. baumannii alanine racemase dimer. Monomers are coloured blue and red, with the surface representation of one monomer also shown in blue. The PLP cofactors are depicted as black ball-and-stick models. (b) Structure of the A. baumannii alanine racemase monomer. Ribbon representation with α-helices coloured orange and β-sheets shown in green. The PLP cofactor covalently bound to Lys34 is shown as a black ball-and-stick model. This figure was produced in PyMOL (DeLano, 2002[DeLano, W. L. (2002). PyMOL. https://www.pymol.org .]).

The essential PLP cofactor is covalently bound to Lys34 via an internal aldimine linkage and extends towards the centre of the α/β-barrel. The pyridine N1 of the PLP ring is stabilized by hydrogen bonding to Arg209, which is further stabilized by interactions with His159. The phosphate tail of PLP is stabilized by several residues from one monomer. The OP1 of the phosphate group hydrogen bonds to Ile212 and Tyr38, OP2 hydrogen bonds to Try341 and OP3 hydrogen bonds to Ile212 and Ser190 (Fig. 2[link]). Arg132 usually interacts with the phenolic O atom (O3′) of PLP, aiding in maintaining the position of the cofactor (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]). However, Arg132 was not within hydrogen-bonding distance of PLP in the AlrAba structure. There was no additional density in the AlrAba structure consistent with the presence of any additional ligands besides PLP.

[Figure 2]
Figure 2
Active site of A. baumannii Alr, depicting the network of hydrogen bonds securing the PLP cofactor as found in monomer A. Hydrogen bonds are indicated by dashed lines. Bond distances are indicated in Å. Functional groups of amino acids are shown. This figure was produced using ChemDraw Pro (v.12.0.2).

3.2. Structural and biochemical comparison with closely related alanine racemases

3.2.1. Overall topology, individual domains and inter-monomer hinge angle

The multiple sequence alignment presented in Fig. 3[link] indicates a high level of sequence similarity between AlrAba and the closely related alanine racemases from P. aeruginosa (DadXPao), B. henselae (AlrBhe) and E. coli (AlrEco). The PLP-binding motif containing the catalytic Lys34 (sequence SMVKANAYGHG) is largely conserved between the various enzymes. In addition, residues which contribute to the inner and middle layers of the active-site entryway of the enzymes are also strongly conserved (Fig. 3[link]). As noted previously, AlrAba lacks electron density in the region containing Tyr254 (in monomer B only), which previous structural and kinetic studies have indicated to be part of the inner layer of the active-site entryway and the second catalytic residue involved in the reaction mechanism (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]; Spies & Toney, 2003[Spies, M. A. & Toney, M. D. (2003). Biochemistry, 42, 5099-5107.]).

[Figure 3]
Figure 3
Structure-based sequence alignment of alanine racemases from A. baumannii, E. coli, B. henselae and P. aeruginosa. Identical residues are shaded black, while grey shading indicates amino acids with conserved physicochemical properties. An asterisk marks the highly conserved PLP-bound lysine and a black diamond marks the location of the catalytic tyrosine, while a black circle indicates the location of a residue which is often carbamylated in alanine racemases that have a lysine at this position. The purple box encloses the conserved PLP-binding motif containing the catalytic lysine. The red box indicates the region of electron density missing from monomer B of AlrAba. I and M represent residues which form the inner and middle layers of the active-site entryway.

Superposition of the Cα atoms of the AlrAba monomer with alanine racemases from Gram-negative bacteria (DadXPao, AlrBhe and AlrEco) indicates a high level of structural similarity between the enzymes (Table 2[link]). The AlrAba monomer is most similar to AlrEco and DadXPao, with which it shares the lowest r.m.s differences (1.30 Å) and the highest sequence similarity (41%). AlrAba is less similar to AlrBhe, with which it has a higher r.m.s difference (1.86 Å) and a lower sequence identity (29%). Comparison of the individual domains and active site of AlrAba with the other alanine racemases also indicates high structural similarity (Table 2[link]).

Table 2
Average r.m.s differences (Å) between the Cα atoms of AlrAba and other alanine racemases

Numbers in parentheses denote sequence identity with AlrAba. Residues from the other structures equivalent to those in AlrAba monomer A were used for the superpositions.

Alanine racemase PDB entry Whole monomer N-terminal domain C-terminal domain§ Active site
AlrEco 2rjg 1.30 (41%) 1.32 (40%) 1.02 (43%) 0.65 (60%)
AlrBhe 3kw3 1.86 (29%) 1.68 (25%) 1.07 (36%) 0.91 (48%)
DadXPao 1rcq 1.30 (41%) 1.30 (39%) 1.08 (45%) 0.56 (58%)
†Calculated using monomer A for AlrEco and DadXPao and monomer B for AlrBhe.
‡Calculated using residues 1–230.
§Calculated using residues 231–354.
¶Calculated using residues 32–38, 54–58, 75–79, 95–99, 121–134, 154–162, 188–195, 206–213 and 338–345 from monomer B and 252–255, 272–276 and 299–304 from monomer A for AlrEco and DadXPao and vice versa for AlrBhe.

The N-terminal domain of AlrAba superimposed best with those from AlrEco and DadXPao, with which it shares the highest sequence identity (40 and 39%, respectively) and the lowest r.m.s differences (1.32 and 1.30 Å, respectively). Consistent with the superpositions involving whole monomers, the N-terminal domain was less similar to that of AlrBhe (Fig. 4[link]a). In the C-terminal domain, AlrAba superimposed equally well with AlrBhe, AlrEco and DadXPao (1.07, 1.02 and 1.08 Å, respectively; Fig. 4[link]b). Analysis of the active sites indicates that this region superimposed the best, having much lower r.m.s differences than whole monomers or individual domains (AlrBhe, 0.91 Å; AlrEco, 0.65 Å; DadXPao, 0.56 Å). This indicates that the active site is highly conserved between the alanine racemase structures in spite of structural deviations between their domains.

[Figure 4]
Figure 4
Cα-atom superposition of AlrAba and other alanine racemases. Colouring is as follows: A. baumannii, green; P. aeruginosa, red; B. henselae, blue; E. coli, orange. Cα-­atom traces showing superposition between the (a) N-terminal and (b) C-terminal domains. (c) Superposition of the N-terminal α/β-barrel domain of whole alanine racemase monomers visualized as a ribbon representation. The PLP cofactor of AlrAba is depicted as a black ball-and-stick model. The difference in inter-monomer hinge angle between the enzymes is indicated by a black double-headed arrow. This figure was produced in PyMOL (DeLano, 2002[DeLano, W. L. (2002). PyMOL. https://www.pymol.org .]).

Previous studies of alanine racemases have indicated that the angle between the N-terminal and C-terminal domains of the individual monomers (the inter-monomer hinge angle) differs between enzymes (LeMagueres et al., 2003[LeMagueres, P., Im, H., Dvorak, A., Strych, U., Benedik, M. & Krause, K. L. (2003). Biochemistry, 42, 14752-14761.]; Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]). The difference in domain orientations is the reason that individual monomers cannot be optimally superimposed as a whole (LeMagueres et al., 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]; Im et al., 2011[Im, H., Sharpe, M. L., Strych, U., Davlieva, M. & Krause, K. L. (2011). BMC Microbiol. 11, 116.]). AlrAba has a hinge angle of 141.6°, identical to that of DadXPao (141.6°) and very similar to that of AlrEco (140.4°), to which it has the most similar active-site architecture (Table 2[link]). The hinge angle of AlrAba is least similar to that of AlrBhe (132.5°), to which it has the lowest percentage sequence identity. This indicates a positive correlation between sequence identity and inter-monomer hinge angle. To date, the best explanation for hinge-angle differences involves the hydrogen-bond interactions formed between residues from the N- and C-terminal tails of opposite monomers (LeMagueres et al., 2003[LeMagueres, P., Im, H., Dvorak, A., Strych, U., Benedik, M. & Krause, K. L. (2003). Biochemistry, 42, 14752-14761.], 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]; Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]). AlrBhe and alanine racemases from G. stearothermophilus (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]), B. anthracis (Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]) and S. aureus (Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]) have additional residues in these areas capable of forming these hydrogen bonds, resulting in similar hinge angles. Shorter alanine racemases such as AlrEco, DadXPao and AlrAba do not have these additional residues (Fig. 3[link]), resulting in differing hinge angles.

3.2.2. Enzyme kinetics

Kinetic characterization of AlrAba revealed that the enzyme has a Vmax of 11.3 U mg−1 and a Km of 0.56 mM for the racemization of D-alanine to L-alanine (Table 3[link]). The other direction of the racemization was not characterized in this study. This is very similar to the kinetic values reported for AlrEco (Vmax = 8.8 U mg−1, Km = 0.31 mM), with which it shares high structural similarity (Table 2[link]) and a similar hinge angle. However, despite AlrAba sharing a high structural similarity and hinge angle with DadXPao, the kinetic parameters differed significantly between these enzymes (Vmax = 134 U mg−1 and Km = 1.40 mM for DadXPao). This indicates that despite high levels of sequence identity, similar active-site structures and hinge angles, the kinetics parameters of even closely related alanine racemases can differ. This is in agreement with what has been observed in previous studies (Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]; Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]). No comparison could be made with AlrBhe as no kinetic information is available.

Table 3
Kinetic parameters for the racemization between L- and D-alanine by alanine racemases

NR, value not reported.

  L-to-D direction D-to-L direction
Alanine racemase Km (mM) Vmax (U mg−1) Km (mM) Vmax (U mg−1)
AlrAba NR NR 0.56§ 11.3§
AlrEco 1.0§ 356§ 0.31§ 8.8§
AlrBhe NR NR NR NR
DadXPao†† 1.40‡‡ 155‡‡ 1.40‡‡ 134‡‡
†One unit is defined as the amount of enzyme which catalyzes the racemization of 1 µmol of substrate per minute.
‡Kinetic parameters reported in the current work.
§Assay performed at 30°C.
¶Kinetic parameters reported by Wu et al. (2008[Wu, D., Hu, T., Zhang, L., Chen, J., Du, J., Ding, J., Jiang, H. & Shen, X. (2008). Protein Sci. 17, 1066-1076.]).
††Kinetic parameters reported by Strych et al. (2000[Strych, U., Huang, H.-C., Krause, K. L. & Benedik, M. J. (2000). Curr. Microbiol. 41, 290-294.]).
‡‡Assay performed at 23°C.
3.2.3. Dimer interface, substrate entryway and active site

Previous studies have found that alanine racemase is dependent on dimerization for enzyme activity, and that a number of residues involved in the dimer interface are highly conserved (Strych & Benedik, 2002[Strych, U. & Benedik, M. J. (2002). J. Bacteriol. 184, 4321-4325.]; Im et al., 2011[Im, H., Sharpe, M. L., Strych, U., Davlieva, M. & Krause, K. L. (2011). BMC Microbiol. 11, 116.]; Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]). This makes the dimer interface a potential target for structure-aided drug design. The area of the dimer interface of AlrAba is 2360 Å2 and is formed by roughly equal contributions from both monomers (73 and 70 residues from monomer A and monomer B, respectively). This interface contains five salt bridges and 22 hydrogen-bond interactions. The dimer interface of AlrAba is most similar to the interface calculated for AlrBhe (2606 Å2), which has similar numbers of residues involved in dimerization (71 and 75 in monomers A and B, respectively). The interface is less similar to the interfaces calculated for DadXPao (1918 Å2) and AlrEco (2816 Å2), both of which have a larger number of residues involved in dimerization (DadXPao, 82 and 82 in monomers A and B, respectively; AlrEco, 81 and 83 in monomers A and B, respectively).

The substrate entryway of AlrAba is comprised of residues from both monomers and consists of an outer, a middle and an inner layer (with the inner layer being in closest proximity to the PLP cofactor and the outer layer being on the surface of the protein). Comparison of AlrAba with the other enzymes indicates that the inner (Ala163, Tyr254′, Tyr273′ and Tyr341) and middle (Asp164, Arg279, Arg299 and Ile339) layers of the entryway are highly conserved (Fig. 3[link]), consistent with previous studies (LeMagueres et al., 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]; Im et al., 2011[Im, H., Sharpe, M. L., Strych, U., Davlieva, M. & Krause, K. L. (2011). BMC Microbiol. 11, 116.]). The highly conserved nature of the alanine racemase substrate entryway makes it a potential target for the development of enzyme-specific inhibitors.

The active-site residues of AlrAba and other alanine racemases superimpose particularly well, with low r.m.s differences (Table 2[link]). The active-site structure of AlrAba is most similar to that of DadXPao (0.56 Å; 58% sequence identity). AlrAba also shares high similarity to AlrEco (0.65 Å; 60% sequence identity), but diverges most from AlrBhe (0.91 Å; 48% sequence identity). This indicates a positive correlation between sequence identity and structural similarity in this region. As mentioned in §[link]3.1, the active site of AlrAba is composed of residues from both monomers, several of which are involved in a hydrogen-bond network with the PLP cofactor (Figs. 2[link] and 5[link]a) and are conserved between alanine racemases (Fig. 5[link]b).

[Figure 5]
Figure 5
Active site of A. baumannii alanine racemase. (a) 2FoFc electron-density map of the active site contoured at 1.0σ with isomesh map shown (1.6 Å carve). The main-chain and side-chain atoms of the AlrAba active-site residues are depicted as sticks. C atoms are green, O atoms red, N atoms blue, S atoms yellow and phosphates orange. The PLP cofactor is depicted as a ball-and-stick model in which C atoms are coloured black. (b) Superposition of the active-site residues of alanine racemases from A. baumannii (green), B. henselae (blue), P. aeruginosa (red) and E. coli (orange). For A. baumannii, a hybrid view is depicted with residues from monomer B, except for the side chain of His159 which is included from monomer A. The PLP cofactors from each structure are depicted as ball-and-stick models. Primes denote residues contributed by the second monomer. The superposition was performed using the residue ranges stated in Table 2[link]. This figure was produced in PyMOL (DeLano, 2002[DeLano, W. L. (2002). PyMOL. https://www.pymol.org .]).

AlrAba is lacking density for Tyr254′ in one active site and previous studies have shown that this residue is important in substrate binding and catalysis (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]; Morollo et al., 1999[Morollo, A. A., Petsko, G. A. & Ringe, D. (1999). Biochemistry, 38, 3293-3301.]; Watanabe et al., 2002[Watanabe, A., Yoshimura, T., Mikami, B., Hayashi, H., Kagamiyama, H. & Esaki, N. (2002). J. Biol. Chem. 277, 19166-19172.]). As noted above, missing density for this residue has previously been observed in the active sites of several alanine racemases (LeMagueres et al., 2005[LeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471-1481.]; Abendroth et al., 2011[Abendroth, J., Gardberg, A. S., Robinson, J. I., Christensen, J. S., Staker, B. L., Myler, P. J., Stewart, L. J. & Edwards, T. E. (2011). J. Struct. Funct. Genomics, 12, 83-95.]; Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]; Palani et al., 2013[Palani, K., Burley, S. K. & Swaminathan, S. (2013). Acta Cryst. F69, 15-19.]). His159 can be identified in the same active site and its side chain forms a hydrogen bond to Arg209, as well as a parallel displaced π-stacking interaction with the PLP ring. The closest contact distance for this interaction is ∼3 Å, which is consistent with other reports in proteins (McGaughey et al., 1998[McGaughey, G. B., Gagné, M. & Rappé, A. K. (1998). J. Biol. Chem. 273, 15458-15463.]).

In the second active site Tyr254′ is present but the His159 side chain is difficult to localize, with little density beyond the β-carbon. A hybrid view with these residues included from opposite active sites (Fig. 5[link]b) reveals that these two residues are located in their usual positions when compared with the active sites from other bacterial alanine racemase structures.

Differences between the AlrAba active site and the other alanine racemase structures included a lack of density consistent with Lys125 carbamylation (Figs. 3[link] and 5[link]b). This modification is present in both AlrEco and DadXPao; it is proposed to aid in the positioning of Arg132 within the active site (Morollo et al., 1999[Morollo, A. A., Petsko, G. A. & Ringe, D. (1999). Biochemistry, 38, 3293-3301.]; LeMagueres et al., 2003[LeMagueres, P., Im, H., Dvorak, A., Strych, U., Benedik, M. & Krause, K. L. (2003). Biochemistry, 42, 14752-14761.]; Wu et al., 2008[Wu, D., Hu, T., Zhang, L., Chen, J., Du, J., Ding, J., Jiang, H. & Shen, X. (2008). Protein Sci. 17, 1066-1076.]) and is a process that is dependent on a relatively high pH (>8.0; Golemi et al., 2001[Golemi, D., Maveyraud, L., Vakulenko, S., Samama, J.-P. & Mobashery, S. (2001). Proc. Natl Acad. Sci. USA, 98, 14280-14285.]). It is possible that the observed lack of carbamylation in the AlrAba structure is owing to the low pH (6.0) at which the enzyme was crystallized. Furthermore, AlrAba did not contain any extra density consistent with other molecules in its active site. Previous alanine racemase structures have reported the presence of additional molecules in the active site such as sulfate (Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]), chloride (Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]) and acetate (Shaw et al., 1997[Shaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329-1342.]; Couñago et al., 2009[Couñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.]; Scaletti et al., 2012[Scaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82-92.]). The lack of substrate-like molecules in the active site could also be a contributory factor to the unusual position of Arg132 owing to this residue playing an integral role in substrate binding.

4. Conclusions

We have reported the crystal structure of A. baumannii alanine racemase from the highly antibiotic-resistant NCTC13302 strain to 1.9 Å resolution. The overall structure is very similar to closely related alanine racemases from Gram-negative bacteria. Interestingly, the partial kinetic analysis generated in this study indicated an inconsistent relationship between structural similarity and racemization rate. The substrate entryway and active sites of these alanine racemases were shown to be highly conserved and are thus possible targets for the development of enzyme-specific inhibitors. The structural determination and biochemical analysis presented here could provide a template for future structure-based drug-design studies targeting AlrAba.

Supporting information


Acknowledgements

This research was undertaken in part on the MX1 beamline at the Australian Synchrotron, Victoria, Australia, and we would like to thank their staff for assistance with data collection. Expert scientific assistance is acknowledged from Roman Mortuza, Otago University. This work was supported by funding from the University of Otago, the New Zealand Group Ltd, c/o Royal Society of New Zealand, L2 Diagnostics Ltd (New Haven, Connecticut, USA), the National Institutes of Health, the Thrash Foundation and the Maurice Wilkins Centre for Molecular Biodiscovery.

References

First citationAbendroth, J., Gardberg, A. S., Robinson, J. I., Christensen, J. S., Staker, B. L., Myler, P. J., Stewart, L. J. & Edwards, T. E. (2011). J. Struct. Funct. Genomics, 12, 83–95.  CrossRef CAS PubMed Google Scholar
First citationAdams, P. D. et al. (2010). Acta Cryst. D66, 213–221.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBattye, T. G. G., Kontogiannis, L., Johnson, O., Powell, H. R. & Leslie, A. G. W. (2011). Acta Cryst. D67, 271–281.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCouñago, R. M., Davlieva, M., Strych, U., Hill, R. E. & Krause, K. L. (2009). BMC Struct. Biol. 9, 53.  Google Scholar
First citationDeLano, W. L. (2002). PyMOL. https://www.pymol.orgGoogle Scholar
First citationDijkshoorn, L., Nemec, A. & Seifert, H. (2007). Nature Rev. Microbiol. 5, 939–951.  Web of Science CrossRef CAS Google Scholar
First citationEliopoulos, G. M., Maragakis, L. L. & Perl, T. M. (2008). Clin. Infect. Dis. 46, 1254–1263.  Web of Science PubMed Google Scholar
First citationEmsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126–2132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEsaki, N. & Walsh, C. T. (1986). Biochemistry, 25, 3261–3267.  CrossRef CAS PubMed Web of Science Google Scholar
First citationEvans, P. (2006). Acta Cryst. D62, 72–82.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGolemi, D., Maveyraud, L., Vakulenko, S., Samama, J.-P. & Mobashery, S. (2001). Proc. Natl Acad. Sci. USA, 98, 14280–14285.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGootz, T. D. & Marra, A. (2008). Expert Rev. Anti Infect. Ther. 6, 309–325.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGrosse-Kunstleve, R. W., Sauter, N. K., Moriarty, N. W. & Adams, P. D. (2002). J. Appl. Cryst. 35, 126–136.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHols, P., Defrenne, C., Ferain, T., Derzelle, S., Delplace, B. & Delcour, J. (1997). J. Bacteriol. 179, 3804–3807.  CAS PubMed Web of Science Google Scholar
First citationIm, H., Sharpe, M. L., Strych, U., Davlieva, M. & Krause, K. L. (2011). BMC Microbiol. 11, 116.  Google Scholar
First citationJoshi, S. G., Litake, G. M., Niphadkar, K. B. & Ghole, V. S. (2003). J. Infect. Chemother. 9, 187–190.  CrossRef PubMed CAS Google Scholar
First citationKrissinel, E. & Henrick, K. (2007). J. Mol. Biol. 372, 774–797.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLambert, M. P. & Neuhaus, F. C. (1972). J. Bacteriol. 109, 1156–1161.  CAS PubMed Web of Science Google Scholar
First citationLaskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). J. Appl. Cryst. 26, 283–291.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationLee, N.-Y., Lee, H.-C., Ko, N.-Y., Chang, C.-M., Shih, H.-I., Wu, C.-J. & Ko, W.-C. (2007). Infect. Control Hosp. Epidemiol. 28, 713–719.  Web of Science CrossRef PubMed Google Scholar
First citationLeMagueres, P., Im, H., Dvorak, A., Strych, U., Benedik, M. & Krause, K. L. (2003). Biochemistry, 42, 14752–14761.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLeMagueres, P., Im, H., Ebalunode, J., Strych, U., Benedik, M. J., Briggs, J. M., Kohn, H. & Krause, K. L. (2005). Biochemistry, 44, 1471–1481.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMatthews, B. W. (1968). J. Mol. Biol. 33, 491–497.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMcCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658–674.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMcGaughey, G. B., Gagné, M. & Rappé, A. K. (1998). J. Biol. Chem. 273, 15458–15463.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMorollo, A. A., Petsko, G. A. & Ringe, D. (1999). Biochemistry, 38, 3293–3301.  Web of Science CrossRef PubMed CAS 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 citationNguyen, D.-D., Ngo, H.-P.-T., Hong, M., Pham, T.-V., Lee, J. H., Lee, J. J., Kwon, D. B., Lee, S. H. & Kang, L.-W. (2013). Acta Cryst. F69, 1041–1044.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNoda, M., Matoba, Y., Kumagai, T. & Sugiyama, M. (2004). J. Biol. Chem. 279, 46153–46161.  Web of Science CrossRef PubMed CAS Google Scholar
First citationPalani, K., Burley, S. K. & Swaminathan, S. (2013). Acta Cryst. F69, 15–19.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPeleg, A. Y., Seifert, H. & Paterson, D. L. (2008). Clin. Microbiol. Rev. 21, 538–582.  Web of Science CrossRef PubMed CAS Google Scholar
First citationPriyadarshi, A., Lee, E. H., Sung, M. W., Nam, K. H., Lee, W. H., Kim, E. E. & Hwang, K. Y. (2009). Biochim. Biophys. Acta, 1794, 1030–1040.  Web of Science CrossRef PubMed CAS Google Scholar
First citationScaletti, E. R., Luckner, S. R. & Krause, K. L. (2012). Acta Cryst. D68, 82–92.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationScott, P. et al. (2007). Clin. Infect. Dis. 44, 1577–1584.  Web of Science CrossRef PubMed CAS Google Scholar
First citationShaw, J. P., Petsko, G. A. & Ringe, D. (1997). Biochemistry, 36, 1329–1342.  CrossRef CAS PubMed Web of Science Google Scholar
First citationSpies, M. A. & Toney, M. D. (2003). Biochemistry, 42, 5099–5107.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStrych, U. & Benedik, M. J. (2002). J. Bacteriol. 184, 4321–4325.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStrych, U., Huang, H.-C., Krause, K. L. & Benedik, M. J. (2000). Curr. Microbiol. 41, 290–294.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStrych, U., Penland, R. L., Jimenez, M., Krause, K. L. & Benedik, M. J. (2001). FEMS Microbiol. Lett. 196, 93–98.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWalsh, C. T. (1989). J. Biol. Chem. 264, 2393–2396.  CAS PubMed Web of Science Google Scholar
First citationWatanabe, A., Yoshimura, T., Mikami, B., Hayashi, H., Kagamiyama, H. & Esaki, N. (2002). J. Biol. Chem. 277, 19166–19172.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWinn, M. D. et al. (2011). Acta Cryst. D67, 235–242.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWu, D., Hu, T., Zhang, L., Chen, J., Du, J., Ding, J., Jiang, H. & Shen, X. (2008). Protein Sci. 17, 1066–1076.  Web of Science CrossRef PubMed CAS Google Scholar
First citationZavascki, A. P., Carvalhaes, C. G., Picão, R. C. & Gales, A. C. (2010). Expert Rev. Anti Infect. Ther. 8, 71–93.  Web of Science CrossRef CAS PubMed Google 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
Volume 70| Part 9| September 2014| Pages 1199-1205
Follow Acta Cryst. F
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds