Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Mechanism of d-alanine transfer to teichoic acids shows how bacteria acylate cell envelope polymers

Abstract

Bacterial cell envelope polymers are often modified with acyl esters that modulate physiology, enhance pathogenesis and provide antibiotic resistance. Here, using the d-alanylation of lipoteichoic acid (Dlt) pathway as a paradigm, we have identified a widespread strategy for how acylation of cell envelope polymers occurs. In this strategy, a membrane-bound O-acyltransferase (MBOAT) protein transfers an acyl group from an intracellular thioester onto the tyrosine of an extracytoplasmic C-terminal hexapeptide motif. This motif shuttles the acyl group to a serine on a separate transferase that moves the cargo to its destination. In the Dlt pathway, here studied in Staphylococcus aureus and Streptococcus thermophilus, the C-terminal ‘acyl shuttle’ motif that forms the crucial pathway intermediate is found on a transmembrane microprotein that holds the MBOAT protein and the other transferase together in a complex. In other systems, found in both Gram-negative and Gram-positive bacteria as well as some archaea, the motif is fused to the MBOAT protein, which interacts directly with the other transferase. The conserved chemistry uncovered here is widely used for acylation throughout the prokaryotic world.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: dltX is required for LTA d-alanylation in S. aureus.
Fig. 2: DltX interacts with DltB and DltD as part of a three-member membrane protein complex.
Fig. 3: DltX has a highly conserved C-terminal motif that is critical for LTA d-alanylation but not for complex formation.
Fig. 4: The invariant tyrosine of DltX is required for d-alanylation of DltD, and this chemistry is conserved across many bacterial cell envelope polymer acylation systems.

Similar content being viewed by others

Data availability

All data generated or analysed during this study are included in this published article (and its Supplementary Information files). The following databases were used for this research: UniProt (including UniRef90), InterPro and various NCBI databases such as GenBank. All plasmids and bacterial strains generated during this work are available upon request. Source data are provided with this paper.

References

  1. Silhavy, T. J., Kahne, D. & Walker, S. The bacterial cell envelope. Cold Spring Harb. Perspect. Biol. 2, a000414 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Percy, M. G. & Gründling, A. Lipoteichoic acid synthesis and function in Gram-positive bacteria. Annu. Rev. Microbiol. 68, 81–100 (2014).

    Article  CAS  PubMed  Google Scholar 

  3. Wecke, J., Madela, K. & Fischer, W. The absence of d-alanine from lipoteichoic acid and wall teichoic acid alters surface charge, enhances autolysis and increases susceptibility to methicillin in Bacillus subtilis. Microbiology 143, 2953–2960 (1997).

    Article  CAS  PubMed  Google Scholar 

  4. Blackburn, N. T. & Clarke, A. J. Characterization of soluble and membrane-bound family 3 lytic transglycosylases from Pseudomonas aeruginosa. Biochemistry 41, 1001–1013 (2002).

    Article  CAS  PubMed  Google Scholar 

  5. Lunderberg, J. M. et al. Bacillus anthracis acetyltransferases PatA1 and PatA2 modify the secondary cell wall polysaccharide and affect the assembly of S-layer proteins. J. Bacteriol. 195, 977–989 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Inês Crisóstomo, M. et al. Attenuation of penicillin resistance in a peptidoglycan O-acetyl transferase mutant of Streptococcus pneumoniae. Mol. Microbiol. 61, 1497–1509 (2006).

    Article  PubMed  Google Scholar 

  7. Aubry, C. et al. OatA, a peptidoglycan O-acetyltransferase involved in Listeria monocytogenes immune escape, is critical for virulence. J. Infect. Dis. 204, 731–740 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Pasquina, L. et al. A synthetic lethal approach for compound and target identification in Staphylococcus aureus. Nat. Chem. Biol. 12, 40–45 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Pier, G. B., Coleman, F., Grout, M., Franklin, M. & Ohman, D. E. Role of alginate O acetylation in resistance of mucoid Pseudomonas aeruginosa to opsonic phagocytosis. Infect. Immun. 69, 1895–1901 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Kovács, M. et al. A functional dlt operon, encoding proteins required for incorporation of d-alanine in teichoic acids in Gram-positive bacteria, confers resistance to cationic antimicrobial peptides in Streptococcus pneumoniae. J. Bacteriol. 188, 5797–5805 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Fabretti, F. et al. Alanine esters of enterococcal lipoteichoic acid play a role in biofilm formation and resistance to antimicrobial peptides. Infect. Immun. 74, 4164–4171 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Neuhaus, F. C. & Baddiley, J. A continuum of anionic charge: structures and functions of d-alanyl-teichoic acids in Gram-positive bacteria. Microbiol. Mol. Biol. Rev. 67, 686–723 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Abachin, E. et al. Formation of d-alanyl-lipoteichoic acid is required for adhesion and virulence of Listeria monocytogenes. Mol. Microbiol. 43, 1–14 (2002).

    Article  CAS  PubMed  Google Scholar 

  14. Vincent Collins, L. et al. Staphylococcus aureus strains lacking d-alanine modifications of teichoic acids are highly susceptible to human neutrophil killing and are virulence attenuated in mice. J. Infect. Dis. 186, 214–219 (2002).

    Article  PubMed  Google Scholar 

  15. Poyart, C. et al. Attenuated virulence of Streptococcus agalactiae deficient in d-alanyl-lipoteichoic acid is due to an increased susceptibility to defensins and phagocytic cells. Mol. Microbiol. 49, 1615–1625 (2003).

    Article  CAS  PubMed  Google Scholar 

  16. Fischer, W., Rösel, P. & Koch, H. U. Effect of alanine ester substitution and other structural features of lipoteichoic acids on their inhibitory activity against autolysins of Staphylococcus aureus. J. Bacteriol. 146, 467–475 (1981).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Peschel, A., Vuong, C., Otto, M. & Götz, F. The d-alanine residues of Staphylococcus aureus teichoic acids alter the susceptibility to vancomycin and the activity of autolytic enzymes. Antimicrob. Agents Chemother. 44, 2845–2847 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Peschel, A. et al. Inactivation of the dlt operon in Staphylococcus aureus confers sensitivity to defensins, protegrins, and other antimicrobial peptides. J. Biol. Chem. 274, 8405–8410 (1999).

    Article  CAS  PubMed  Google Scholar 

  19. Guariglia-Oropeza, V. & Helmann, J. D. Bacillus subtilis σV confers lysozyme resistance by activation of two cell wall modification pathways, peptidoglycan O-acetylation and d-alanylation of teichoic acids. J. Bacteriol. 193, 6223–6232 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Koprivnjak, T., Peschel, A., Gelb, M. H., Liang, N. S. & Weiss, J. P. Role of charge properties of bacterial envelope in bactericidal action of human group IIA phospholipase A2 against Staphylococcus aureus. J. Biol. Chem. 277, 47636–47644 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Saar-Dover, R. et al. d-Alanylation of lipoteichoic acids confers resistance to cationic peptides in Group B Streptococcus by increasing the cell wall density. PLoS Pathog. 8, e1002891 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Pasquina, L. W., Santa Maria, J. P. & Walker, S. Teichoic acid biosynthesis as an antibiotic target. Curr. Opin. Microbiol. 16, 531–537 (2013).

    Article  CAS  PubMed  Google Scholar 

  23. Tzipilevich, E., Pollak-Fiyaksel, O., Shraiteh, B. & Ben-Yehuda, S. Bacteria elicit a phage tolerance response subsequent to infection of their neighbors. EMBO J. 41, e109247 (2022).

    Article  CAS  PubMed  Google Scholar 

  24. Matos, R. C. et al. d-Alanylation of teichoic acids contributes to Lactobacillus plantarum-mediated Drosophila growth during chronic undernutrition. Nat. Microbiol. 2, 1635–1647 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Perego, M. et al. Incorporation of d-alanine into lipoteichoic acid and wall teichoic acid in Bacillus subtilis: identification of genes and regulation. J. Biol. Chem. 270, 15598–15606 (1995).

    Article  CAS  PubMed  Google Scholar 

  26. Baddiley, J. & Neuhaus, F. C. The enzymic activation of d-alanine. Biochem. J. 75, 579–587 (1960).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Heaton, M. P. & Neuhaus, F. C. Role of the d-alanyl carrier protein in the biosynthesis of d-alanyl-lipoteichoic acid. J. Bacteriol. 176, 681–690 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Nikolopoulos, N. et al. DltC acts as an interaction hub for AcpS, DltA and DltB in the teichoic acid d-alanylation pathway of Lactiplantibacillus plantarum. Sci. Rep. 12, 13133 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Hofmann, K. A superfamily of membrane-bound O-acyltransferases with implications for Wnt signaling. Trends Biochem. Sci. 25, 111–112 (2000).

    Article  CAS  PubMed  Google Scholar 

  30. Ma, D. et al. Crystal structure of a membrane-bound O-acyltransferase. Nature 562, 286–290 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Reichmann, N. T., Picarra Cassona, C. & Gründling, A. Revised mechanism of d-alanine incorporation into cell wall polymers in Gram-postive bacteria. Microbiology 159, 1868–1877 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Wood, B. M., Santa Maria, J. P. Jr., Matano, L. M., Vickery, C. R. & Walker, S. A partial reconstitution implicates DltD in catalyzing lipoteichoic acid d-alanylation. J. Biol. Chem. 293, 17985–17996 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kamar, R. et al. DltX of Bacillus thuringiensis is essential for d-alanylation of teichoic acids and resistance to antimicrobial response in insects. Front. Microbiol. 8, 1–13 (2017).

    Article  Google Scholar 

  34. Storz, G., Wolf, Y. I. & Ramamurthi, K. S. Small proteins can no longer be ignored. Annu. Rev. Biochem. 83, 753–777 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Abi Khattar, Z. et al. The dlt operon of Bacillus cereus is required for resistance to cationic antimicrobial peptides and for virulence in insects. J. Bacteriol. 191, 7063–7073 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Kumra Taneja, N. et al. d-Alanine modification of a protease-susceptible outer membrane component by the Bordetella pertussis dra locus promotes resistance to antimicrobial peptides and polymorphonuclear leukocyte-mediated killing. J. Bacteriol. 195, 5102–5111 (2013).

    Article  Google Scholar 

  37. Santa Maria, J. P. Jr. et al. Compound–gene interaction mapping reveals distinct roles for Staphylococcus aureus teichoic acids. Proc. Natl Acad. Sci. USA 111, 12510–12515 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hancock, I. C., Wiseman, G. & Baddiley, J. Biosynthesis of the unit that links teichoic acid to the bacterial wall: inhibition by tunicamycin. FEBS Lett. 69, 75–80 (1976).

    Article  CAS  PubMed  Google Scholar 

  39. Campbell, J. et al. Synthetic lethal compound combinations reveal a fundamental connection between wall teichoic acid and peptidoglycan biosynthesis in Staphylococcus aureus. ACS Chem. Biol. 6, 106–116 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Mirdita, M. et al. ColabFold: making protein folding accessible to all. Nat. Methods 19, 679–682 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Lin, S., Lu, X., Chang, C. C. & Chang, T. Y. Human acyl-coenzyme A:cholesterol acyltransferase expressed in Chinese hamster ovary cells: membrane topology and active site location. Mol. Biol. Cell 14, 2447–2460 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Qian, H. et al. Structural basis for catalysis and substrate specificity of human ACAT1. Nature 581, 333–338 (2020).

    Article  CAS  PubMed  Google Scholar 

  44. Long, T., Sun, Y., Hassan, A., Qi, X. & Li, X. Structure of nevanimibe-bound tetrameric human ACAT1. Nature 581, 339–343 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Guan, C. et al. Structural insights into the inhibition mechanism of human sterol O-acyltransferase 1 by a competitive inhibitor. Nat. Commun. 11, 2478 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Sui, X. et al. Structure and catalytic mechanism of a human triacylglycerol-synthesis enzyme. Nature 581, 323–328 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Wang, L. et al. Structure and mechanism of human diacylglycerol O-acyltransferase 1. Nature 581, 329–332 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Jiang, Y., Benz, T. L. & Long, S. B. Substrate and product complexes reveal mechanisms of Hedgehog acylation by HHAT. Science 372, 1215–1219 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Coupland, C. E. et al. Structure, mechanism, and inhibition of Hedgehog acyltransferase. Mol. Cell 81, 5025–5038.e10 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Yang, J., Brown, M. S., Liang, G., Grishin, N. V. & Goldstein, J. L. Identification of the acyltransferase that octanoylates ghrelin, an appetite-stimulating peptide hormone. Cell 132, 387–396 (2008).

    Article  CAS  PubMed  Google Scholar 

  51. Dodds, A. W., Ren, X. D., Willis, A. C. & Law, S. K. The reaction mechanism of the internal thioester in the human complement component C4. Nature 379, 177–179 (1996).

    Article  CAS  PubMed  Google Scholar 

  52. Huguenin-Dezot, N. et al. Trapping biosynthetic acyl-enzyme intermediates with encoded 2,3-diaminopropionic acid. Nature 565, 112–117 (2019).

    Article  CAS  PubMed  Google Scholar 

  53. Jones, C. S., Anderson, A. C. & Clarke, A. J. Mechanism of Staphylococcus aureus peptidoglycan O-acetyltransferase A as an O-acyltransferase. Proc. Natl Acad. Sci. USA 118, e2103602118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Moynihan, P. J. & Clarke, A. J. O-Acetylation of peptidoglycan in Gram-negative bacteria: identification and characterization of peptidoglycan O-acetyltransferase in Neisseria gonorrhoeae. J. Biol. Chem. 285, 13264–13273 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Wang, G., Lo, L. F., Forsberg, L. S. & Maier, R. J. Helicobacter pylori peptidoglycan modifications confer lysozyme resistance and contribute to survival in the host. mBio 3, e00409–e00412 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Franklin, M. J. & Ohman, D. E. Identification of algI and algJ in the Pseudomonas aeruginosa alginate biosynthetic gene cluster which are required for alginate O-acetylation. J. Bacteriol. 178, 2186–2195 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Baker, P. et al. P. aeruginosa SGNH hydrolase-like proteins AlgJ and AlgX have similar topology but separate and distinct roles in alginate acetylation. PLoS Pathog. 10, e1004334 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Sychantha, D. et al. PatB1 is an O-acetyltransferase that decorates secondary cell wall polysaccharides. Nat. Chem. Biol. 14, 79–85 (2018).

    Article  CAS  PubMed  Google Scholar 

  59. Chanasit, W., Gonzaga, Z. J. C. & Rehm, B. H. A. Analysis of the alginate O-acetylation machinery in Pseudomonas aeruginosa. Appl. Microbiol. Biotechnol. 104, 2179–2191 (2020).

    Article  CAS  PubMed  Google Scholar 

  60. Garai, P. & Blanc-Potard, A. Uncovering small membrane proteins in pathogenic bacteria: regulatory functions and therapeutic potential. Mol. Microbiol. 114, 710–720 (2020).

    Article  CAS  PubMed  Google Scholar 

  61. Makarewich, C. A. The hidden world of membrane microproteins. Exp. Cell Res. 388, 111853 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Zallot, R., Oberg, N. & Gerlt, J. A. The EFI web resource for genomic enzymology tools: leveraging protein, genome, and metagenome databases to discover novel enzymes and metabolic pathways. Biochemistry 58, 4169–4182 (2019).

    Article  CAS  PubMed  Google Scholar 

  63. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Do, T. et al. Staphylococcus aureus cell growth and division are regulated by an amidase that trims peptides from uncrosslinked peptidoglycan. Nat. Microbiol. 5, 291–303 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hesser, A. R. The length of lipoteichoic acid polymers controls Staphylococcus aureus cell size and envelope integrity. J. Bacteriol. https://doi.org/10.1128/JB.00149-20 (2020).

  66. Tsirigos, K. D., Peters, C., Shu, N., Käll, L. & Elofsson, A. The TOPCONS web server for consensus prediction of membrane protein topology and signal peptides. Nucleic Acids Res. 43, W401–W407 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Shen, W., Le, S., Li, Y. & Hu, F. SeqKit: a cross-platform and ultrafast toolkit for FASTA/Q file manipulation. PLoS ONE 5, e0163962 (2016).

    Article  Google Scholar 

  68. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 11, 539 (2011).

    Article  Google Scholar 

  69. Eddy, S. R. Profile hidden Markov models. Bioinformatics 14, 755–763 (1998).

    Article  CAS  PubMed  Google Scholar 

  70. Hopf, T. A. et al. The EVcouplings Python framework for coevolutionary sequence analysis. Bioinformatics 35, 1582–1584 (2019).

    Article  CAS  PubMed  Google Scholar 

  71. Haas, R., Koch, H. U. & Fischer, W. Alanyl turnover from lipoteichoic acid to teichoic acid in Staphylococcus aureus. FEMS Microbiol. Lett. 21, 27–31 (1984).

    Article  CAS  Google Scholar 

  72. Koch, H. U., Döker, R. & Fischer, W. Maintenance of d-alanine ester substitution of lipoteichoic acid by reesterification in Staphylococcus aureus. J. Bacteriol. 164, 1211–1217 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Ha, R. et al. Accumulation of peptidoglycan O-acetylation leads to altered cell wall biochemistry and negatively impacts pathogenesis factors of Campylobacter jejuni. J. Biol. Chem. 291, 22686–22702 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Spiers, A. J., Kahn, S. G., Bohannon, J., Travisano, M. & Rainey, P. B. Adaptive divergence in experimental populations of Pseudomonas fluorescens. I. Genetic and phenotypic bases of wrinkly spreader fitness. Genetics 161, 33–46 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Corrigan, R. M., Abbott, J. C., Burhenne, H., Kaever, V. & Gründling, A. c-di-AMP is a new second messenger in Staphylococcus aureus with a role in controlling cell size and envelope stress. PLoS Pathog. 7, e1002217 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Forsberg, L. S. et al. Localization and structural analysis of a conserved pyruvylated epitope in Bacillus anthracis secondary cell wall polysaccharides and characterization of the galactose-deficient wall polysaccharide from avirulent B. anthracis CDC 684. Glycobiology 22, 1103–1117 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank S. Moussa and M. Stone for generating plasmids used in this study, and T. Sisley for streaking strains and starting overnight cultures during COVID-related shift work. We thank F. Hammel and S. Early for critical feedback on the manuscript, as well as all members of the Walker Lab for helpful scientific discussions. We thank T. Pang from the Bernhardt and Rudner laboratories for pTP16. We thank R. Tomaino of the Harvard Medical School Taplin Mass Spectrometry Facility for MS analysis, as well as J. Paulo of the Gygi laboratory for help with preliminary MS experiments. Funding for this work was provided by the National Institutes of Health grants P01 AI083214 and U19 AI158028 to S.W. E.D.S. acknowledges support from National Institutes of Health grant T32 GM139775. B.J.S. acknowledges support from the National Science Foundation Graduate Research Fellowship Program under grant nos. DGE 1745303 and DGE 2140743. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation.

Author information

Authors and Affiliations

Authors

Contributions

B.J.S. and S.W. conceived the project. B.J.S. and E.D.S. designed and performed experiments and analysed data, with input from S.W. B.J.S. and S.W. wrote the original draft of the manuscript and, along with E.D.S., revised the manuscript.

Corresponding author

Correspondence to Suzanne Walker.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Andreas Peschel, Dorte Frees and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 S. aureus strains without dltX cannot survive on tunicamycin but still produce LTAs.

a, Chemical structures of three repeat units of Staphylococcus aureus LTA (left) and one repeat unit of Streptococcus pneumoniae LTA (right)2. d-Alanine esters are shown in blue. b, Full set of controls for the spot titer assay shown in Fig. 1c. S. aureus strains were grown on tryptic soy agar (TSA) plates containing the indicated compounds: dimethyl sulfoxide (DMSO; 1.25 µL per 1.00 mL of TSA), isopropyl-β-d-1-thiogalactopyranoside (IPTG; 1.00 mM), and/or tunicamycin (tunic.; 1.0 µg/mL). Tunic. inhibits WTA biosynthesis, IPTG was used to induce expression of the indicated plasmid-borne cassette, and DMSO was used as a vehicle control. The image of the plate with tunicamycin and IPTG has been reproduced here for clarity. The plasmid harboring dltXABCD is leaky thus resulting in growth of the sixth strain even in the absence of IPTG. c, An α-LTA western blot indicates that all of the strains tested in b and Fig. 1c produce LTAs at roughly equivalent levels to wild-type S. aureus. The strain in Lane 2 (4S5)75 is a mutant that lacks LtaS and contains a suppressor mutation that permits growth in the absence of LTAs. It does not produce any LTAs. The LTAs themselves run as a smear from ~15–37 kDa. d, Spot titer assay results for a partially different set of S. aureus strains. Here, the complementation constructs are under anhydrotetracycline (aTc)-inducible control, and the overall cassettes are genomically integrated. The TSA plates here contained DMSO, tunic., and aTc as indicated, with the concentrations of DMSO and tunic. being the same as for b and that for aTc being 0.4 µM. In these strains, the dltXABCD complementation construct does not appear leaky. e, SDS-PAGE autoradiography of LTAs from S. aureus cells grown in the presence of d-[14C]alanine again shows that dltX is required for LTA d-alanylation. Here, the same set of strains from d was used. f, An α-LTA western blot indicates that all strains tested in d and e produce LTAs at roughly equivalent levels to wild-type S. aureus.

Source data

Extended Data Fig. 2 DltX is a predicted bitopic transmembrane protein with an extracellular C-terminus and is required for DltD to copurify with DltB.

a, (top) The TOPCONS web server66 for consensus-based membrane protein topology prediction was used to predict the topology of S. aureus DltX. The sequence of S. aureus DltX is shown with the predicted localization of each individual residue indicated beneath. (bottom) An AlphaFold240 model of S. aureus DltX generated using ColabFold41. This specific model is from an overall model of DltX with DltD and was selected here because it illustrates the topology of the protein well. We note that in the DltDX model, DltX’s predicted topology is in agreement with the known topology31 of DltD. b, An α-myc western blot indicates that a similar amount of myc-StDltD was present in the solubilized membranes loaded onto the TALON immobilized metal affinity chromatography (IMAC) resin during purifications of StDltB-His10 (and interactors) from E. coli cells expressing StdltB and StdltD, with or without StdltX. A similar amount of myc-StDltD was also present in the flow-through from the resin in the two samples. However, no myc-StDltD could be observed in the IMAC elution fraction from the sample derived from E. coli cells that did not express StdltX.

Source data

Extended Data Fig. 3 The predicted S. aureus DltBDX structure is a high-confidence prediction.

a, (top) The top-ranked AlphaFold2 model of the S. aureus DltBDX complex, with proteins colored as in Fig. 2c, is a high-confidence model. The predicted local Distance Difference Test (plDDT) is a metric of model confidence, with plDDT values > 90 representing very high confidence, 70–90 representing moderate confidence, 50–70 representing low confidence, and < 50 representing very low confidence. The graph in the upper right shows the plDDT values at each residue for all five models. The numbering on the x-axis here corresponds to the residue number for the concatenated sequence of the complex (that is, the sequence of DltX concatenated N-terminally to the sequence of DltB, and then that overall sequence concatenated N-terminally to the sequence of DltD). The rightmost of the two models underneath the ‘Rank 1’ heading shows a graphical depiction of the plDDT values with a color gradient scale (dark blue representing a plDDT value of 100, and red representing a plDDT value of 0). (bottom) Each of the other four models is highly similar to the top-ranked model. Five models were generated by ColabFold, and the models were ranked by predicted template modeling (pTM) scores. b, The relative positions between the three proteins are high-confidence based on the Predicted Aligned Error (PAE). Here, the value at each (x, y) point in the graphs indicates the ‘expected position error at residue x if the predicted and true structures were aligned on residue y’ (https://alphafold.ebi.ac.uk/faq). Low values (blue) indicate low expected position error. Again, the numbering on the x-axis corresponds to the residue number for the concatenated sequence of the complex. Numbering on the y-axis (not shown) would place ‘0’ at the top of the axis. c, Views of the surface representation of the S. aureus DltBDX AlphaFold2 model.

Extended Data Fig. 4 The C-terminal motif of DltX, and particularly its invariant tyrosine residue, is required for S. aureus growth on tunicamycin.

a, Full set of controls for the spot titer assay shown in Fig. 3b. S. aureus strains were grown on TSA plates containing the indicated compounds: DMSO at 1.25 µL per 1.00 mL of TSA, IPTG at 1.00 mM, anhydrotetracycline (aTc) at 0.4 µM, and tunicamycin at 1.0 µg/mL. Tunic. inhibits wall teichoic acid biosynthesis, IPTG was used to induce expression of the indicated plasmid-borne cassette (here, null refers to a strain with an empty IPTG-inducible cassette in the plasmid), aTc was used to induce expression of the indicated chromosomally integrated cassette, and DMSO was used as a vehicle control for the tunic. The image of the plate with tunic., IPTG, and aTc has been reproduced here for clarity. flag-dltX∆motif encodes for an N-terminally FLAG-tagged DltX variant lacking the last six amino acids. b, Full set of controls for the spot titer assay shown in Fig. 3c. Compounds were used at the same concentrations as above, for the same reasons as above. Here, the bracketed sequences in the superscripts of the aTc-inducible cassettes represent the C-terminal motif sequence of the given DltX variant. Red letters denote changes from the wild-type sequence.

Extended Data Fig. 5 Length alterations at the C-terminus of DltX are not well-tolerated.

a, Full set of controls for the spot titer assay shown in Fig. 3d. S. aureus strains were grown on TSA plates containing the indicated compounds: DMSO at 1.25 µL per 1.00 mL of TSA, IPTG at 1.00 mM, anhydrotetracycline (aTc) at 0.4 µM, and tunicamycin at 1.0 µg/mL. Tunic. inhibits wall teichoic acid biosynthesis, IPTG was used to induce expression of the indicated plasmid-borne cassette (here, null refers to a strain with an empty IPTG-inducible cassette in the plasmid), aTc was used to induce expression of the indicated chromosomally integrated cassette, and DMSO was used as a vehicle control for the tunic. The image of the plate with tunic., IPTG, and aTc has been reproduced here for clarity. Here, the bracketed sequences in the superscripts of the aTc-inducible cassettes represent the C-terminal motif sequence of the given DltX variant. Red letters denote changes from the wild-type sequence, and a red hyphen denotes the deletion of a residue. b, A spot titer assay shows that, of all the DltX mutants we have tested that do not allow for growth of a dltX-null strain on tunicamycin when expressed at low levels (from a chromosomally integrated aTc-inducible cassette), only the DltX mutant with an alanine added onto the C-terminus can complement when expressed at high levels (from an IPTG-inducible cassette on a medium copy number plasmid). Here, dltX* indicates that the plasmid encodes for a variant of DltX with the C-terminal sequence indicated in the ‘DltX C-term.’ column. Again, red letters denote changes from the wild-type sequence, and a red hyphen denotes the deletion of a residue. Both dltXwt (wild-type) and all dltX* here encode an N-terminal FLAG tag.

Extended Data Fig. 6 The S. aureus DltBX and DltDX predicted structures are high-confidence predictions.

a, The top-ranked AlphaFold2 model of the S. aureus DltBX complex is a high-confidence model. The predicted local Distance Difference Test (plDDT) is a metric of model confidence, with plDDT values > 90 representing very high confidence, 70–90 representing moderate confidence, 50–70 representing low confidence, and < 50 representing very low confidence. The graph in the upper right shows the plDDT values at each residue for five models. The numbering on the x-axis corresponds to the residue number for the concatenated sequence of the complex (that is, the sequence of DltB concatenated N-terminally to the sequence of DltX). The rightmost of the two full model images depicts the plDDT values with a color gradient scale (dark blue representing a plDDT value of 100, and red representing a value of 0). The zoom-in image shows how the DltX C-terminal motif fits into the DltB tunnel. b, The relative positions between the proteins are high-confidence based on the Predicted Aligned Error (PAE). Here, the value at each (x, y) point in the graphs indicates the ‘expected position error at residue x if the predicted and true structures were aligned on residue y’ (https://alphafold.ebi.ac.uk/faq). Low values (blue) indicate low expected position error. Again, the numbering on the x-axis corresponds to the residue number for the concatenated sequence of the complex. Numbering on the y-axis (not shown) would place ‘0’ at the top of the axis. c, The top-ranked AlphaFold2 model of the S. aureus DltDX complex is a high-confidence model. The rightmost of the two full model images depicts the plDDT values with a color gradient scale matching that found in a. The zoom-in image shows how the DltX C-terminal motif fits into the DltD active site region. The graph in the lower right shows the plDDT values at each residue for five models. The numbering on the x-axis here parallels that in a. d, The relative positions between the proteins are high-confidence based on the PAE. The numbering on the x- and y-axes parallels the numbering in b.

Extended Data Fig. 7 DltB, DltD, and DltX levels are comparable between samples in the DltD d-alanylation in vitro reconstitution assay, and active forms of all three proteins plus DltA and DltC allow for rapid d-alanylation of DltD.

a, A Coomassie-stained SDS-PAGE gel from an in vitro reconstitution experiment run in parallel to the one shown in Fig. 4b but with cold d-alanine rather than d-[14C]alanine. Lane 4 contains all five active proteins, whereas Lanes 3, 5, and 6 are missing the indicated protein(s). Lanes 7–9 contain four active proteins and the indicated mutant. b, A Coomassie-stained SDS-PAGE gel (run using a Tris-glycine gel rather than a Bis-Tris gel as in a) of purified StDltBDX complex, specifically containing wild-type DltB, inactive DltX (Y56F), and wild-type DltD, provided as a reference for comparison to the samples in a. This suggests the impurity below DltB in a may be a degradation product. c, SDS-PAGE autoradiograph from a time course of the in vitro reconstitution with DltA, DltC, and DltBDX. For the ‘10 sec.’ time point, the DltBDX complex was added to the DltC charging reaction, the sample was mixed by pipetting up and down several times, and then immediately 4x XT sample buffer was added and mixed followed by flash freezing of the sample in liquid nitrogen. For all other time points, after the DltBDX complex was added to the DltC charging reaction, samples were incubated at 30 °C for the indicated amount of time before addition of SDS-PAGE loading buffer and immediate flash freezing in liquid nitrogen.

Source data

Extended Data Fig. 8 DltB, DltD, and DltX levels are comparable between samples in the DltX d-alanylation in vitro reconstitution assay, and d-alanylation of DltXY56K is position-specific.

a, A Coomassie-stained SDS-PAGE gel from an in vitro reconstitution experiment run in parallel to the one shown in Fig. 4c but with cold d-alanine rather than d-[14C]alanine. Lanes 2 and 3 contain wild-type (wt) 1× or 3xFLAG-DltX, while Lanes 4 and 5 contain 1x or 3xFLAG-DltXY56K. b, A Coomassie-stained SDS-PAGE gel (run using a Tris-glycine gel rather than a Bis-Tris gel as in a) of purified StDltBDX complex, specifically containing wild-type DltB, DltXN57K, and wild-type DltD, provided as a reference for comparison to the samples in a. This suggests that the background smearing in the Coomassie-stained gels in a and c (right) is a result of the high-pH reaction conditions combined with the specific gel type used (see Supplementary Fig. 2a for further evidence of this). Additionally, the impurity below DltB in a and c (right) may be a degradation product. We note that the mobility of 1xFLAG-DltX in the Bis-Tris gel here is different from that in the Tris-glycine gel in a, but the bands were confirmed by western blot (See Supplementary Figs. 2b-c). c, (left) SDS-PAGE autoradiography of in vitro reconstitution results shows that the DltXY56K mutant is d-alanylated in a position-specific manner. All lanes contain wild-type DltA, DltC, and DltBDX complex, with the exception that the DltX in the complex is the variant with the C-terminal sequence indicated by the key. These reconstitutions were run alongside those shown in Fig. 4c, and the samples were run on the same gel and thus imaged together. (right) A Coomassie-stained SDS-PAGE gel from an in vitro reconstitution run in parallel to the one shown in c (left) but with cold d-alanine rather than d-[14C]alanine. Again, all lanes contain wild-type DltA, DltC, and DltBDX complex, with the exception that the DltX in the complex is the variant with the C-terminal sequence indicated by the key. In the key, blue coloring indicates the lysine, and red coloring indicates the tyrosine-to-phenylalanine swap.

Source data

Extended Data Fig. 9 A C-terminal motif reminiscent of DltX’s is found at the C-termini of MBOAT proteins that partner with DltD-like proteins in various prokaryotic cell envelope polymer acylation pathways.

a, Graphical depictions illustrate the genomic colocalization of genes encoding MBOAT proteins (green) and DltD-like proteins (blue) in bacterial cell envelope polymer acylation systems, as well as a system of unknown function found in some archaea. The chemical structures show a portion of the relevant polymer for each system, with the modification (acetylation in each of these cases), in blue54,57,76. The lower right depicts a potential polymer acylation system of unknown function found in the archaeal species Nitrosopumilus adriaticus. We have designated the genes mbt1 for MBOAT1 and mbtp1 for MBOAT1 Partner. b, AlphaFold2 models of the two-protein complexes encoded by the genes from a. The models are all shown cytosolic side-down. c, Sequence logo generated by EVcouplings70 with N. gonorrhoeae PatA as the query sequence showing the conservation of the final 51 amino acid positions of an alignment of bacterial cell envelope acylation-associated MBOAT proteins (as well as similar MBOAT proteins of unknown function), with the overall consensus sequence and corresponding NgPatA sequence shown underneath. The height of each individual letter represents the degree of conservation of that specific amino acid at the given position. The final transmembrane (TM) helix indicated is a prediction based on AlphaFold. d, Active sites of the MBOAT proteins of three of the systems shown in this figure (the corresponding active site for NgPatA is found in Fig. 4f) from AlphaFold2 models of the MBOAT proteins alone. These images are all shown with an ‘aerial’ view of the active site (looking toward the cytosol from the extracytoplasmic region). The catalytic histidine known to be required for all studied MBOAT proteins is shown in cyan, and the invariant tyrosine present in this class of MBOAT proteins is shown in purple.

Extended Data Fig. 10 The MBOAT proteins in non-Dlt pathway cell envelope acylation systems have additional C-terminal helices that allow the C-terminal motif of these proteins to nestle into the active site.

Alignments of predicted structures of MBOAT proteins from bacterial cell envelope polymer acetylation systems (and a system of unknown function in archaea) with the predicted structure of the DltBX complex show that the non-DltB MBOAT proteins likely share a similar architecture to DltB but with a C-terminal extension of several additional helices. All of the predicted structures here were generated using the ColabFold implementation of AlphaFold2 and aligned using the PyMOL ‘super’ command.

Supplementary information

Supplementary Information

Supplementary Methods, Figs. 1–6, Tables 19–22 and references.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–18.

Supplementary Data 1

All ColabFold output files for the AlphaFold predictions generated in this work.

Source data

Source Data Fig. 1

Unprocessed autoradiograph.

Source Data Fig. 2

Unprocessed gels and western blot.

Source Data Fig. 3

Unprocessed gel.

Source Data Fig. 4

Unprocessed autoradiographs.

Source Data Extended Data Fig. 1

Unprocessed western blots and autoradiograph.

Source Data Extended Data Fig. 2

Unprocessed western blot.

Source Data Extended Data Fig. 7

Unprocessed gels and autoradiograph.

Source Data Extended Data Fig. 8

Unprocessed gels and autoradiograph.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schultz, B.J., Snow, E.D. & Walker, S. Mechanism of d-alanine transfer to teichoic acids shows how bacteria acylate cell envelope polymers. Nat Microbiol 8, 1318–1329 (2023). https://doi.org/10.1038/s41564-023-01411-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01411-0

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology