QT 625123 DG
QT 625123 DG
QT 625123 DG
Title
The Molecular Basis of Substrate Recognition, Catalysis, and Regulation in Sortase Enzymes
Permalink
https://escholarship.org/uc/item/625123dg
Author
Jacobitz, Alex William
Publication Date
2015
Peer reviewed|Thesis/dissertation
Los Angeles
Sortase Enzymes
by
2015
ABSTRACT OF THE DISSERTATION
Sortase Enzymes
By
covalently links protein substrates to the cell wall or polymerizes proteinaceous pili. Sortase
enzymes are membrane anchored cysteine transpeptidases that recognize a cell wall sorting
signal motif at the C-terminus of their primary protein substrate and covalently attach it to an
amino-nucleophile located in their secondary substrate. For cell wall anchoring sortases, this
secondary substrate is the cell wall precursor lipid II, while for pilin polymerizing sortases it is a
lysine sidechain within their pilin protein substrate. A number of major questions have remained
unanswered concerning the binding interactions that govern the substrate specificity of these
enzymes, their catalytic mechanism, and the mechanism through which appendages that contact
the active site regulate their activity. This dissertation advances the understanding of the sortase
mechanism through the elucidation of new sortase structures, and the characterization of their
dynamic behavior.
ii
Chapter three of this thesis describes the solution structure of the class D sortase from
Bacillus anthracis by NMR. Class D enzymes anchor proteins involved in bacterial sporulation
to the cell wall, and this is the first structure of a class D sortase to ever be determined. NMR
analysis of the enzyme uncovered a rigid substrate binding pocket and a novel dimerization
interface. Chapter four describes the crystal structure of the class B sortase from Staphylococcus
aureus bound to a substrate analog. This work provided new insight into the biophysical basis of
substrate recognition. The structure combined with computational modeling and molecular
dynamics simulations led to the discovery of a substrate-stabilized oxyanion hole that is used to
stabilize tetrahedral reaction intermediates. Molecular dynamics simulations and the high degree
of sequence conservation inherent in these enzymes suggest that all members of the sortase
superfamily will stabilize high energy reaction intermediates in a similar manner. Chapter five
investigates the additional structural features some sortases use to regulate access to their active
site, focusing on sortase C-1 from Streptococcus pneumoniae as a model. Using NMR,
mutagenesis, and biochemical experiments this work demonstrated that the enzyme utilizes a
rigid N-terminal appendage, termed the “lid,” to maintain the enzyme in an inactive state. The
results of in vitro biochemical assays indicate that the lid prevents cleavage of the primary
substrate by preventing access to the catalytic cysteine residue. Both in vitro activity and access
to the active site cysteine could be increased by the incorporation of mutations which were
shown by NMR to destabilize the lid and increase its flexibility. Based on these results we
propose that on the cell surface, lid containing class C enzymes exist in a dormant state and are
only activated during pilus biogenesis by interactions with either their substrate or other factors
on the cell surface. The results of these experiments have provided new insight into substrate
iii
The dissertation of Alex Jacobitz is approved.
James U. Bowie
Beth A. Lazazzera
2015
iv
DEDICATION
To my father, who taught me the value of hard work and perseverance, and always
faced adversity with humor and grace. This would not have been possible without
- Thought is the wind, knowledge the sail, and mankind the vessel -
v
TABLE OF CONTENTS
2.11 Figures................................................................................................................................49
vi
Chapter 3. Solution structure of the sortase required for efficient production of infectious
3.3.3 Disordered Loops Positioned near the Active Site Might Mediate in vitro ...............
Dimerization ..........................................................................................................89
3.5 Figures................................................................................................................................98
vii
4.1 Overview ..........................................................................................................................121
Complex ...............................................................................................................139
4.6 Figures..............................................................................................................................143
Chapter 5. The “lid” of Streptococcus pneumoniae SrtC1 is not Dynamic in Solution and
viii
5.3.1 Initial NMR analysis indicates SpnSrtC1 forms dimers in solution ....................171
5.3.2 NMR dynamics data indicate that the SpnSrtC1 lid is not dynamic in vitro .......172
5.3.3 Mutations to the “lid” of SpnSrtC1 make the region highly dynamic in
solution.................................................................................................................175
5.3.5 A closed lid is important for protecting the active site cysteine from
oxidation ..............................................................................................................178
5.6 Figures..............................................................................................................................188
ix
LIST OF FIGURES
Figure 3.1 NMR spectra and amino acid sequence of BaSrtC ................................................98
Figure 3.5 Comparison of representative structures from class A–D sortases ......................106
Figure S3.1 Residue by residue T2, T1 and 15N{1H} Heteronuclear NOE data for BaSrtCΔ55 117
Figure 4.2 Structure and interactions of the NPQT* modifier in complex with SrtB...........145
Figure 4.3 Expanded view of the active site in the energy-minimized model of the SrtB-
Figure 4.6 Structures and free energy profiles of the SrtA and SrtB thioacyl complexes ....151
x
Figure 4.7 The SrtB transpeptidation mechanism showing how the sorting signal may
Figure 5.4 NMR shows SpnSrtC1 mutants have more dynamic lids ....................................192
Figure 5.5 SpnSrtC1 lid mutants are more active in vitro .....................................................194
Figure 5.6 The lid of SpnSrtC1 protects the active site cysteine from modification by a
xi
LIST OF TABLES
Table 3.1 Structural Statistics for the Solution Structure of BaSrtCΔ55 ...............................102
xii
ACKNOWLEDGEMENTS
First and foremost I would like to thank my advisor, Dr. Robert Clubb for his guidance,
support, and for the unlimited supply of coffee he provided: without which my PhD would not
have been possible. My collaborators, Dr. Jeff Wereszczynski, Dr. J. Andrew McCammon, Dr.
Sung Wook Yi, and Dr. Michael Jung for helping in the design and execution of numerous
experiments, and I would like to additionally thank Jeff specifically for teaching me to run
molecular dynamics simulations. Dr. Michael Sawaya for patiently teaching me X-ray
crystallography. Dr. Robert Peterson for helping me to better understand NMR and for writing
new pulse programs when necessary. My rotation mentors, Dr. Todd Yeates and Dr. William
Gelbart for their advice on topics in and out of the lab. My committee members: Dr. James
Bowie, Dr. Pascal Egea, Dr. Beth Lazazzera, and Dr. Joseph Loo for their thoughts and
guidance. The undergraduates, John Cho, Allen Tsiyer, and Scott McConnell, and rotation
students, Brendan Amer, Grace Huang, Angelyn Nguyen, Reid O’Brien Johnson, Kevin Eden,
Mimi Ho, and Jessica Davis, whose efforts in the lab proved to be invaluable. Members of the
Clubb lab, Dr. Scott Robson, Dr. Valerie Villareal, Dr. Thomas Spirig, Dr. Ethan Weiner, Dr.
Reza Malmirchegini, Dr. Albert Chan, Megan Sjodt, Brendan Amer, Grace Huang, Michele
Kattke, and Ramsay Macdonald, whose friendship, scientific discourse, and moral support
helped to get me through graduate school. My friends and family for putting up with me for the
last 6 years. The Cellular and Molecular Biology Training Grant, Audree Fowler Fellowship in
Protein Science, John M. Jordan Memorial Award for Research in Biochemistry, and the UCLA
xiii
career, and for the additional opportunities for enriching my scientific development that they
have provided.
Chapter three of this dissertation is a version of the manuscript: Robson S. A., Jacobitz A.
W., Clubb R.T., Phillips M.L., Clubb R.T. Solution Structure of the Sortase Required for
permission. Copyright 2011, American Chemical Society. We express great appreciation to Dr.
Joseph Loo and members of the Clubb laboratory for their assistance.
Wereszczynski, J., Yi, S. W., Amer, B. R., Huang, G. L., Nguyen, A. V., Sawaya, M. R., Jung,
of Biological Chemistry. This work was supported, in whole or in part, by National Institutes of
Health Grants AI52217 (to R. T. C. and M. E. J.), GM31749 (to J. A. M.), and T32
GM007185 (to A. W. J. and G. L. H.). This work was also supported by the National Science
Foundation, the Howard Hughes Medical Institute, the Center for Theoretical Biological Physics,
the National Biomedical Computation Resource, and the National Science Foundation
Supercomputer Centers.
xiv
Chapter Five of this dissertation is a version of a manuscript in preparation titled: The
“lid” of S. pneumoniae SrtC1 is not dynamic in solution and regulates substrate access to the
active site. Jacobitz A. W., Yi, S. W., McConnell, S., Peterson, R., Jung, M. E., and Clubb, R. T.
2015.
xv
Vita
xvi
PUBLICATIONS
Jacobitz, A. W., Wereszczynski, J., Yi, S. W., Amer, B. R., Huang, G. L., Nguyen, A. V., Sawaya, M.
R., Jung, M. E., McCammon, J. A. & Clubb, R. T. Structural and computational studies of the
Staphylococcus aureus Sortase B-substrate complex reveal a substrate-stabilized oxyanion hole. J. Biol.
Robson, S. A., Jacobitz, A. W., Phillips, M. L. & Clubb, R. T. Solution Structure of the Sortase Required
for Efficient Production of Infectious Bacillus anthracis Spores. Biochemistry. 51, 7953–7963 (2012).
doi:10.1021/bi300867t
King, N. P., Jacobitz, A. W., Sawaya, M. R., Goldschmidt, L. & Yeates, T. O. Structure and folding of a
xvii
Chapter 1
1
1.1 Overview
transpeptidation reaction that covalently attaches proteins to the cell wall, or polymerizes protein
substrates into long filamentous pili. These surface anchored proteins and pili serve to augment
the pathogenic potential of their bacterial proprietor by allowing the organism to more
effectively interact with host cells and other bacteria. Sortase anchored surface proteins perform
a variety of functions including adhesion, immune evasion, nutrient acquisition, and spore
formation. Numerous studies have shown that the removal of sortase enzymes is generally not
lethal to the bacterium, but often renders the bacterium avirulent. This observation has led to a
wealth of research into the biology, structures, mechanism, regulation, and inhibition of these
enzymes.
2
1.2 Introduction to Sortase Biology
Gram-positive organisms are recognized by the presence of a thick cell wall. While its
function in protecting the cell against mechanical and osmotic stresses is well known, the cell
wall also provides an additional function as a scaffold for the display of a large number of
surface proteins by sortase enzymes1–3. All sortase enzymes catalyze the formation of a peptide
bond that covalently joins two substrates together. For most sortases, this transpeptidation
reaction links a substrate protein bearing an appropriate cell wall sorting signal (CWSS)
containing a C-terminal 5 residue recognition motif and transmembrane helix to the cell wall, but
a select subgroup of the enzymes instead catalyze the formation of an isopeptide bond between a
CWSS containing protein and a lysine sidechain of a pilin motif containing protein. Based on
sequence similarity and structural homology, sortases can be categorized into 6 distinct families.
These families each typically operate to anchor a specific type of substrate, with the class A
enzymes being constitutively expressed and anchoring a number of diverse proteins to the cell
surface, and each of the other types being relegated to a separate operon and only expressed
along with their substrates under specific conditions. Most Gram-positives have at least one
sortase gene, with certain bacteria harboring genes for multiple sortases which typically function
non-redundantly to attach a specific subset of surface proteins to the cell wall. For bacteria with a
single sortase, that sortase is typically class A and will attach multiple proteins to the cell wall,
with the most prolific of these enzymes being sortase A from Listeria monocytogenes, which is
predicted to display 43 distinct proteins4,5. Other enzymes, like Staphylococcus aureus sortase B,
The majority of sortase enzymes studied to date catalyze the covalent attachment of their
substrates to the cell wall, and many of these enzymes have been shown to be particularly
3
important in virulence. Genetic knock-outs of sortase A and B have both been shown to reduce
virulence in a mouse model of infection for the human pathogens S. aureus7,8, and Bacillus
anthracis9, without affecting the ability of these organisms to grow in rich media. Sortase A KOs
have also been shown to reduce virulence of several other human pathogens including Listeria
importance of this class of enzymes and led to the development of a number of small-molecule
sortase inhibitors which may eventually lead to the production of a novel anti-virulence
therapeutic15.
In addition to their importance as virulence factors, additional work has gone into
understanding the need for multiple classes of sortase and these studies have shown that the
different classes function non-redundantly to attach specific substrates to specific regions of the
cell wall. While all sortase enzymes are known to recognize a CWSS, the individual classes
generally recognize different 5 residue sorting signal motifs, and the structural basis of this
sorting signals, these different classes of enzymes have also been shown to anchor their
substrates to distinct locations on the cell wall. While most sortases mount their proteins to the
outermost surface of the cell wall, certain class B enzymes which mount proteins involved in
delivering heme-iron to the cell have been shown to anchor proteins to a more buried location
within the cell wall, which is likely necessary to develop a heme transport conduit to deliver
heme from the surface of the cell wall to the cell membrane18,19. Sortase enzymes and their
substrates have also been shown to colocalize to distinct foci on the bacterial surface. Class A
sortases in cocci have been shown to specifically anchor substrates at division septa, where
subsequent growth and expansion will lead to a ring-like distribution of surface proteins across
4
the cell wall20–22. For rod shaped bacteria, cell surface anchoring has been localized to helical
patterns along the cell surface1,23. Certain class C sortases which are typically involved in pilus
polymerization have also been implicated in surface protein localization, and class A and C
sortases have been shown to anchor pili to different locations on the cell surface24. While it is
now well established that certain sortases are capable of localizing their substrates to discrete
locations on the cell wall, further study is needed to understand the mechanisms that cause this
localization to occur.
Instead of attaching proteins to the cell wall, a subset of sortase enzymes catalyze the
polymerization of pili. These sortases are typically found clustered in genomic islands along with
their substrates, and some bacteria can harbor genomic islands coding for the production of
multiple distinct of pili25. These pilus islands typically harbor at least 2 sortases and at least 2
substrates, and the molecular mechanism by which they coordinate to assemble pili is only
beginning to be understood4. Typically these operons code for one primary pilus protein that is
polymerized to make up the shaft of the pilus, and then one or two accessory proteins that
function as either the base of the pilus for attachment to the cell wall, or as an additional adhesin
to be attached as the tip, or elaborate the length of the shaft26. Many of the class C sortases have
high sequence identity, and some studies have suggested that some of these proteins are
multifaceted, polymerizing the pilus, attaching accessory pilus proteins, and attaching the pilus
to the wall, and that other sortases in the operon may function semi-redundantly24,27–29.
Interestingly, several more recent studies have indicated that individual sortases may each be
dedicated to specific steps in the pilin polymerization process, recognizing different pilin motifs
at from separate domains within a single pilin protein substrate to build more complex branched,
or “knobbed” structures30,31.
5
While much is known about the basic biology of sortase enzymes, structural studies in
this field have answered, and will continue to answer questions about their substrate recognition,
molecular mechanism, regulation, and the interactions they have with other proteins on the cell
surface that are all necessary to build a complete understanding of the sorting reaction that these
6
1.3 References
1. Liew PX, Wang CLC, Wong S-L. Functional characterization and localization of a
Bacillus subtilis sortase and its substrate and use of this sortase system to covalently
anchor a heterologous protein to the B. subtilis cell wall for surface display. J Bacteriol.
2012;194(1):161-175. doi:10.1128/JB.05711-11.
2. Paterson GK, Mitchell TJ. The biology of Gram-positive sortase enzymes. Trends
4. Spirig T, Weiner EM, Clubb RT. Sortase enzymes in Gram-positive bacteria. Mol
2005;187(14):4928-4934. doi:10.1128/JB.187.14.4928-4934.2005.
class of surface protein during Staphylococcus aureus pathogenesis. Proc Natl Acad Sci U
S A. 2002;99(4):2293-2298. doi:10.2307/3057958.
sortase mutants defective in the display of surface proteins and in the pathogenesis of
7
animal infections. Proc Natl Acad Sci U S A. 2000;97(10):5510-5515.
doi:10.1073/pnas.080520697.
8. Weiss WJ, Lenoy E, Murphy T, et al. Effect of srtA and srtB gene expression on the
9. Zink SD, Burns DL. Importance of srtA and srtB for growth of Bacillus anthracis in
10. Bierne H, Mazmanian SK, Trost M, et al. Inactivation of the srtA gene in Listeria
11. Garandeau C, Réglier-Poupet H, Dubail I, Beretti J-L, Berche P, Charbit A. The sortase
http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=127754&tool=pmcentrez&ren
12. Chen S, Paterson GK, Tong HH, Mitchell TJ, DeMaria TF. Sortase A contributes to
2005;253(1):151-154. doi:10.1016/j.femsle.2005.09.052.
8
13. Paterson GK, Mitchell TJ. The role of Streptococcus pneumoniae sortase A in
doi:10.1016/j.micinf.2005.06.009.
14. Kharat AS, Tomasz A. Inactivation of the srtA Gene Affects Localization of Surface
2765.2003.
15. Suree N, Jung ME, Clubb RT. Recent advances towards new anti-infective agents that
inhibit cell surface protein anchoring in Staphylococcus aureus and other gram-positive
doi:10.2174/138955707782110097.
16. Jacobitz AW, Wereszczynski J, Yi SW, et al. Structural and computational studies of the
17. Suree N, Liew CK, Villareal VA, et al. The Structure of the Staphylococcus aureus
doi:10.1074/jbc.M109.022624.
2005;280(16):16263-16271. doi:10.1074/jbc.M500071200.
9
19. Mazmanian SK, Skaar EP, Gaspar AH, et al. Passage of Heme-Iron Across the Envelope
doi:10.1126/science.1081147.
20. Raz A, Fischetti VA. Sortase A localizes to distinct foci on the Streptococcus pyogenes
doi:10.1073/pnas.0808301105.
21. DeDent A, Bae T, Missiakas DM, Schneewind O. Signal peptides direct surface proteins
2668. doi:10.1038/emboj.2008.185.
23. Bruck S, Personnic N, Prevost M-C, Cossart P, Bierne H. Regulated Shift from Helical to
2011;193(17):4425-4437. doi:10.1128/JB.01154-10.
24. Fälker S, Nelson AL, Morfeldt E, et al. Sortase-mediated assembly and surface topology
2958.2008.06396.x.
2008;16(1):33-40. doi:10.1016/j.tim.2007.10.010.
10
26. Hendrickx APA, Budzik JM, Oh S-Y, Schneewind O. Architects at the bacterial surface -
sortases and the assembly of pili with isopeptide bonds. Nat Rev Microbiol.
2011;9(3):166-176. doi:10.1038/nrmicro2520.
27. Mora M, Bensi G, Capo S, et al. Group A Streptococcus produce pilus-like structures
containing protective antigens and Lancefield T antigens. Proc Natl Acad Sci U S A.
2005;102(43):15641-15646. doi:10.1073/pnas.0507808102.
28. Ton-That H, Marraffini LA, Schneewind O. Sortases and pilin elements involved in pilus
doi:10.1111/j.1365-2958.2004.04117.x.
29. Neiers F, Madhurantakam C, Fälker S, et al. Two crystal structures of pneumococcal pilus
sortase C provide novel insights into catalysis and substrate specificity. J Mol Biol.
2009;393(3):704-716. doi:10.1016/j.jmb.2009.08.058.
30. El Mortaji L, Terrasse R, Dessen A, Vernet T, Di Guilmi AM. Stability and Assembly of
doi:10.1074/jbc.M109.082776.
31. El Mortaji L, Fenel D, Vernet T, Di Guilmi AM. Association of RrgA and RrgC into the
352. doi:10.1021/bi201591n.
11
Chapter 2
Transpeptidases
12
The work described in this chapter is a version of a manuscript to be submitted for publication.
Jacobitz, A. W., Kattke, M. D., Wereszczynski, J., and Clubb, R. T., 2015.
13
2.1 Overview
reaction that can be used to anchor surface proteins to the peptidoglycan or polymerize them into
long filamentous pili. Gram-positive organisms require these surface proteins to effectively
interact with their environment and to mount infections. Because many of these surface proteins
are necessary for pathogenesis, sortase enzymes are considered virulence factors, and this
designation has led to extensive characterization of their structures and mechanism. To date, over
40 structures of sortase enzymes have been deposited in the PDB and this wealth of structural
information combined with numerous in vitro and in vivo analyses of their catalytic activity has
led to a thorough understanding of the molecular mechanism of substrate binding and catalysis.
These structural and mechanistic observations have prompted a number of computational studies
to understand the dynamic behavior of these enzymes and to develop inhibitors. Here we review
14
2.2 Introduction
to protect themselves against environmental stressors and as a scaffold for the display of a large
number of surface proteins by sortase enzymes. Sortase enzymes are cysteine transpeptidases
that link their two substrates together via the formation of a new covalent bond1–3. Many of these
enzymes catalyze the covalent attachment of surface proteins to peptidoglycan, while a subset of
the family instead catalyzes the polymerization of filamentous pili on the cell surface. Since their
discovery over 15 years ago4, genome sequencing efforts have revealed that Gram-positive
bacteria harboring a conventional cell wall typically encode multiple sortase enzymes5–7. The
surface proteins which these enzymes attach to the cell wall are important for adhesion, nutrient
acquisition, spore formation, and immune evasion2,6,8, and it has been shown that sortase knock-
outs are defective in establishing persistent infections9,10. Since many surface proteins are
necessary for efficiently interacting with a host organism, and their deletion results in a defect in
pathogenesis, sortase enzymes have been classified as virulence factors9,10. This designation has
led to the development of several small-molecule inhibitors of sortase, and the search for
molecules with increased potency and specificity is a subject of ongoing research8,11–15. Beyond
their importance as a potential drug target, the ability of these enzymes to covalently join two
distinct substrates through the formation of a new peptide bond has led to their development as a
biochemical reagent to covalently attach proteins to each other, attach fluorophores or drugs to
The sortase superfamily is large. At present over 800 sortase enzymes in 300 species of
bacteria have been identified6. Based on their primary sequences these enzymes have been
classified into six distinct families, whose members share related atomic structures and
15
frequently have similar functions. On the cell surface all sortases characterized to date function
as cysteine transpeptidase enzymes that attach surface proteins bearing a C-terminal cell wall
charged cytoplasmic anchor and a 5 residue recognition motif to an amino nucleophile located on
the cell surface. While the basic chemistry of this reaction appears to be highly conserved,
sortases can function to either attach proteins to the cell wall, or assemble pili (Fig. 1). All
sortases characterized to date are believed to be associated with the extracellular membrane via
Additionally, sequence conservation and a number of structural and biochemical studies have
determined that the active site of all sortase enzymes studied to date consists of a His-Cys-Arg
catalytic triad22,23. In atomic structures of sortases, these residues are invariantly clustered in the
active site such that the cysteine residue is bracketed by the histidine and arginine residues.
Sortase enzymes are also known to harbor a highly conserved TLXTC motif leading to, and
encompassing the catalytic cysteine24 which could be important for stabilizing the position of the
active site cysteine residue and maintaining it in a catalytically competent orientation. While the
basic architecture is conserved, unique class specific variations enable distinct CWWS and
nucleophiles to be recognized, and in some instances the variations appear to play critical
regulatory roles.
A number of excellent reviews have been written regarding the overall function of
sortases in biology, their specific role in pilin assembly, their development as biochemical
reagents and efforts to discover inhibitors of this important family2,7,8,25–29. Here we review what
is known about the structural basis of catalysis and sortase function on the cell surface,
highlighting recent structural and computational studies that have revealed unique family
16
specific structural variations, as well as the mechanism of sorting signal recognition and enzyme
regulation.
The majority of early work to understand the structure and mechanism of sortase
enzymes was conducted on the Staphylococcus aureus SrtA (SaSrtA) enzyme, and this enzyme
is typically considered to be the archetypal member of the family. Here we briefly summarize
what has been learned about the molecular basis of its function from biochemical and structural
studies. The first structure of the enzyme was determined by the Clubb group in 200130 (Fig. 2).
Initial analysis of the protein sequence revealed the presence of an N-terminal signal sequence,
and transmembrane helix preceding a 181 residue soluble domain containing the absolutely
conserved catalytic cysteine residue4. The structure of this active soluble domain, SaSrtAΔ59 was
determined by NMR and showcases the now canonical closed 8 stranded beta barrel architecture.
This represented the first example of a new protein fold, now defined by the Structural
Classification of Proteins (SCOP) project31 as a “sortase fold.” This core domain architecture is
highly conserved across all families of sortase with small alterations present in the various other
classes (expanded upon below). The structure is well resolved with the exception of a single loop
between strands β6 and β7. Subsequently, a crystal structure of the same construct was solved at
2.0 Å resolution32. The two structures are very similar with an RMSD of 1.97 over all Cα, but
there are several slight differences: first, the crystal structure showcases elongated β7 and β8
strands, and a short helix in the β6/β7 loop in two of the 3 proteins in the asymmetric unit. This
loop was unstructured in the solution NMR structure, and the authors cite the lack of a single
structure for these residues in the asymmetric unit as additional evidence for its flexibility. The
17
active site architecture is also similar in both structures with the catalytic histidine, cysteine, and
arginine residues positioned together within a deep groove. The conserved TLXTC motif which
ends with the catalytic cysteine residue in all sortases largely forms stabilizing interactions with
other residues in the protein’s core, with the exception of the “X” residue (an Ile for SaSrtA)
which extends below the active site cysteine to form a portion of the hydrophobic pocket that
surrounds the active site. SaSrtA has additionally been shown to require Ca2+ for efficient
catalysis, and removal of the ion results in 5 fold lower activity30. NMR chemical shift mapping
revealed that the Ca2+ ion binds in a pocket between the β3/β4 loop and the β6/β7 loop, where it
is coordinated by the sidechains of E105, E108, D112, E171 and a backbone carbonyl from
N11433. NMR dynamics experiments also indicated that binding of calcium altered the mobility
of the β6/β7 loop which contacts the active site, replacing fast, ps-ns timescale motions with
slower, µs-ms timescale motions which are typically associated with ligand binding and
catalysis.
Substrates for the enzyme are recognized by the presence of a (CWSS) harboring the
canonical LPXTG sorting signal. Catalysis begins with a nucleophilic attack from the active site
cysteine residue on the carbonyl carbon in the peptide bond between the Thr and Gly residues in
the sorting signal. This forms a tetrahedral intermediate that quickly collapses to form a semi-
stable thioacyl intermediate covalently attaching the substrate to the sortase’s active site
cysteine7,24,34. The secondary substrate, a pentaglycine cross-bridge peptide on the cell wall
precursor lipid II, then performs a nucleophilic attack on the carbonyl carbon in the newly
formed thioacyl bond between the sortase and substrate, forming a second tetrahedral
intermediate that collapses to form a new peptide bond between the primary substrate and lipid
II34–36. This covalent complex is then incorporated into the cell wall via the standard cell wall
18
synthesis machinery. If the secondary substrate is not available, the reaction can proceed via a
secondary, terminal pathway wherein water performs the second nucleophilic attack and cleaves
the thioacyl bond between sortase and substrate, hydrolyzing the peptide and releasing it to
solvent (Fig. 2D). In this latter reaction, the enzyme acts as a protease. On the cell surface this
would have the presumably negative effect of cleaving the protein from its membrane anchor
Members of the sortase-superfamily have been classified into distinct sub-families based
elaborate their surfaces with a diverse array of protein substrates. These enzymes frequently
operate non-redundantly with each sortase displaying proteins that contain specific sorting signal
motifs. This 5 residue sorting signal is generally similar to the canonical LPXTG motif seen for
SaSrtA with variations to the Leu and Gly positions of the recognition motif occurring most
frequently among signals for the different classes, while the Pro and Thr residues are largely
invariant6,37. Class A and B enzymes are prevalent in bacteria in the Firmicutes phylum. Many
class A enzymes are thought to perform a “housekeeping” role in that they are constitutively
expressed and generally anchor multiple protein substrates to the cell wall with a variety of
functions dependent on the organism. Typically in organisms with only a single sortase, the
sortase will be of the class A variety. Most class B enzymes also anchor proteins to the cell wall.
In many instances these enzymes anchor proteins involved in iron acquisition and typically exist
on a separate, iron regulated operon with their substrate(s). Not all class B enzymes are involved
in heme acquisition, and some members of this group are instead responsible for assembling
19
pili38,39. Class C enzymes are the most unique members of the sortase superfamily. Most
members of this class function to construct pili, requiring them to function as pilus polymerases
by joining the CWSS of one pilin subunit to a lysine residue in a second pilin subunit. The
polymerization of these subunits results in the construction of long (up to 1 µm) filamentous
fibers which are used for adhesion to host cells and have also been implicated in biofilm
formation7,40,41. The remaining types of sortases are less well studied but appear to attach
proteins to the cell wall like their class A and B counterparts. Many species of bacilli encode
class D enzymes that attach proteins involved in spore formation to the cell wall. Members of the
class E enzymes are predominantly found in Actinobacteria species. They are thought perform a
housekeeping role similar to class A enzymes in Firmicutes and have also been shown to anchor
proteins involved in the formation of aerial hyphae6,42. Class F enzymes are also present in
Actinobacteria, but their function is currently unknown. Here we focus on the structural
variations inherent in each class of sortase, as the biological implications of these enzymes and
At present, over 40 structures of sortase enzyme have been reported (Table 1). All
sortases studied to date are constructed from the same conserved 8 β-stranded sortase fold shown
in Fig. 2. However, members from each class exhibit unique variations that may be reflective of
their distinct functionalities in the cell. Generally the most significant discrepancies can be
localized to 4 distinct structural foci: (i) the N-terminal segment of the enzyme between the
transmembrane helix and the first β-strand of the barrel, (ii) the loop between strands β6 and β7
(the β6/β7 loop), (iii) the loop between strands β7 and β8 (the β7/β8 loop), and (iv) the C-
20
terminus of the enzyme which for some class C sortases harbors an additional transmembrane
helix, although this has not been visualized in any reported structure thus far20,21 (Fig. 3). In
many instances these unique structural differences have been shown to influence the enzyme’s
ability to recognize its specific sorting signal motif, and in others they are postulated to perform
a regulatory role by controlling substrate access to the active site. In particular, N-terminal
appendages in some enzymes have been shown to modulate sorting signal access to the active
site, while large structural variation in the β6/β7 and β7/β8 loops has been implicated in substrate
binding. Below we compare the known structures of sortases, highlighting class specific
Class A enzymes. Beyond the archetypal SaSrtA enzyme from S. aureus described in
detail above, class A enzyme structures have been reported for Streptococcus pyogenes43,
enzymes are believed to perform a housekeeping role and recognize the LPXTG sorting signal.
The β6/β7 loop of these enzymes contains a short helix that in structures of SaSrtA and BaSrtA
bound to their cognate sorting signal plays a critical role in signal recognition47,48. Interestingly,
members of the class A family differ in the mobility of this key active site loop. In SaSrtA the
loop is disordered in the absence of substrate, whereas all other class A enzymes display an
ordered β6/β7 loop and thus maintain a preformed binding site for the sorting signal. Many of the
structures also differ substantially from one another in the surface near the active site histidine
residue. This difference was first highlighted in the crystal structure of S. pyogenes SrtA
(SpySrtA)43. The catalytic core of this enzyme assumes the same canonical 8 stranded β-barrel
sortase fold, but unlike SaSrtA, the active site Cys-S is oriented towards the active site His-δN, at
a distance of 5.4 Å (in SaSrtA these groups are pointed away from another and separated by 6.5
21
Å). Most notably, a unique groove runs between these two residues whose walls are defined by
residues in helix H1 and the β7/β8 loop, and whose base is defined by residues in the β4/β5 loop.
This groove provides a potential exit tunnel for the substrate that could accommodate binding of
a full length protein bearing amino acids C-terminal of the LPXTG motif: a feature which was
not obvious in SaSrtA structures wherein the active site histidine residue is actually occluded
from solvent and this channel closed by interactions between residues in the β7/β8 loop and helix
H1. The structure of the S. agalactiae SrtA (SagSrtA) has also been determined and is most
similar to the SpySrtA structure, maintaining the same exit channel between His118 and Cys184
active site residues with its most distinguishing feature being an altered conformation for its
β6/β7 loop which is likely a byproduct of its crystallization conditions as this loop is involved in
coordinating a Zn2+ ion in conjunction with a Glu158 from an adjacent protein in the asymmetric
unit45. The active site His118 residue is also involved in coordinating a Zn2+ ion along with
His136 from a neighboring SagSrtA in the asymmetric unit, but this interaction does not seem to
move the residue into a non-physiological position, as it aligns well with the histidine residues
from active sites of other SrtA enzymes. The B. anthracis SrtA (BaSrtA) structure is noteworthy
in that it possesses an N-terminal appendage which wraps behind helix H1 and contacts the
active site, potentially acting as a cap to prevent access by the secondary substrate, lipid II (Fig.
3A)44. The most recently determined sortase structure, that of S. mutans SrtA (SmSrtA), was
crystallized with a significant portion of its N-terminus intact46. Interestingly, this enzyme
crystallized in an apparent dimer in which an extended N-terminal helix interacts with residues in
the active site of a symmetry related molecule. The authors suggest that this interaction is likely
an artifact of crystallization, but it is interesting to consider that in many recent sortase structures
solved with extended N-termini (especially when considering other classes of sortase, discussed
22
further below), these regions frequently maintain some form of interaction with the active site. It
should also be noted that an additional structure of a class A enzyme, SrtA from Streptococcus
pneumoniae (SpnSrtA), exists in the PDB as a β-strand-swapped dimer, in which the β6/β7 loop
extends from one monomer into an adjacent monomer in the asymmetric unit placing the β7 and
β8 strands in the β-barrel core of an adjacent molecule in the asymmetric unit. No publications
associated with this PDB entry could be found in PubMed, and it is unclear whether this
Class B enzymes. The class B sortases which are typically involved in iron acquisition are
largely similar in structure to the class A sortases discussed previously with the exception of a
longer, structured β6/β7 loop. The first structure of a class B enzyme was of the S. aureus SrtB
sulfhydryl modifiers. When compared to the canonical SaSrtA structure, the most obvious
differences are in the presence of two additional helices N-terminal of the first β-strand and a
much longer loop inserted between strands β6 and β7 which contains an additional α-helix (Fig.
3B). These structural differences could potentially be important for dictating the class B
enzyme’s preference for anchoring substrates to buried, uncrosslinked portions of the cell
wall50,51. The unmodified structures of SaSrtB and B. anthracis SrtB (BaSrtB) were solved later
the same year and exhibit remarkably similar structures (Cα RMSD = 3.23Å). The main
differences between these two enzymes are localized to a slight conformational change in the
β6/β7 loop, a lack of density for a section of BaSrtB’s β7/β8 loop, and an altered conformation of
the short loop between helices H1 and H2. An active site cysteine mutant of the SrtB enzyme
from Clostridium difficile (CdSrtB) was also solved earlier this year by X-ray crystallography52
and displays an almost identical structure to SaSrtB (RMSD = 1.93 Å for all Cα). Interestingly,
23
this class B enzyme may bear a more generalized function as a genetic analysis has predicted 7
possible substrates, none of which are iron associated as is typical of other class B enzymes52.
Beyond these fairly similar class B sortases, the structure of two additional pilin polymerizing
class B sortases from S. pyogenes (SpySrtB) and S. pneumoniae (SpnSrtB) (also referred to as
SrtG-1) have also been solved38,53. These proteins are extremely unique in that their genetic
localization and biochemical analysis would have grouped them with class C pilin polymerizing
sortases, but their structures indicate that they are in fact most similar to class B types (Fig. 3).
The SpySrtB and SpnSrtB structures have a similar arrangement of secondary structure elements,
including the helix containing β6/β7 loop, but also bear an additional short β-strand after this
conserved helix. This alteration adds an uncommon 9th β-strand to these proteins, but this strand
is not inserted into the conserved β-barrel core and instead pairs with a portion of strand β6 that
extends above the core of the beta barrel. Understanding how these class B sortases fulfill the job
of the structurally distinct class C sortases to polymerize pili instead of attaching proteins to the
cell wall will likely be an extremely important step towards understanding the structural
Class C enzymes. Class C sortase structures are most notable for the presence of a
conserved, elongated N-terminal region that occludes the active site, termed “the lid” (Fig.
3C)7,28,54,55. This class has also been the recipient of a significant number of structural studies,
with a total of 15 structures currently deposited in the PDB. Interestingly, every class C, pilin-
polymerizing sortase studied to date contains this lid region, which invariably contains a
DP(F/W/Y) motif. In all but 1 of the class C sortases studied to date, the conserved DP(F/W/Y)
motif contacts the active site through a hydrogen bond between the Asp residue in the lid and the
arginine residue in the active site as well as through a sulfur-aromatic interaction between the
24
active site Cys-S and the aromatic sidechain in the conserved lid motif. Because of these key
interactions, these two residues have been referred to as “anchors” of the lid54,55. The first class C
enzymes to be structurally characterized were SrtC-1 and SrtC-3 from S. pneumoniae (SpnSrtC1
and SpnSrtC3) by Manzano and coworkers in 200854. Upon analyzing the structures the authors
noted that the B-factors for the lid region covering the active site were the highest in the structure
and suggested that the lid was likely mobile in solution and that this mobility would unmask the
active site allowing for recognition of the substrate. Subsequent structures of class C sortases
from S. agalactiae21,56,57, A. oris58, and S. pneumoniae59 furthered this idea with additional
structures showing similar patterns of elevated B-factors or stretches of missing electron density
in regions flanking the anchor residues in the lid. Although many organisms utilize multiple class
C sortases to polymerize pili, systems for which multiple SrtC enzymes have been solved for an
individual organism have shown minimal structural variation. These systems typically show only
small shifts in the position or amount of density found for the N-terminal extension harboring the
lid, with perhaps the most significant deviation from the norm being the addition of a short C-
terminal α-helix opposite the active site in the structure of SpnSrtC3 which has not been
functionally characterized54. Although the structures of most class C sortases show the lid
blocking access to the active site, a structure of S. agalactiae SrtC-1 (SagSrtC1) bound to
MTSET, a nonspecific sulfhydryl modifier, also maintains the lid in a closed state indicating that
the active site cysteine for this enzyme is at least partially accessible even while the lid is
closed56. Mutations to the lid region have shown no effect on transpeptidation in vivo20,21, while
similar mutations have instead been shown to decrease enzyme stability55 and increase the rate of
hydrolysis in vitro21 leading to the theory that the lid performs a regulatory function.
Interestingly SagSrtC1 has been solved in multiple crystal forms, and one of these structures,
25
solved in space group C2, showcases the lid in an “open” conformation providing a model for
what is presumed to be a necessary lid-opening event expected to precede catalysis (Fig. 3C,
compare left to right)56. The enzyme maintains the typical sortase fold and, excluding the N-
terminal extension preceding the β-barrel core, is extremely similar to SagSrtC1 structures
previously solved in space groups P212121 and P312 with an average backbone RMSD of 0.72 Å.
However, in space group C2, residues A38-E71 which typically form the flexible lid structure,
instead form an extended helical structure with the aromatic lid anchor residue of the conserved
DP(F/W/Y) motif, Y51, displaced from the active site by over 30 Å to a position where it stacks
against the backbone of helix H2. An additional crystal structure of S. suis SrtC-1 (SsSrtC1) also
maintains a similar open form of the lid, with the same extended helix replacing what was
spore formation and have received less attention than classes A-C, with only two structures
currently available in the PDB. The first class D structure was determined by NMR for the B.
anthracis SrtD enzyme (BaSrtD, and also previously referred to as SrtC) in 201261. The BaSrtD
solution structure is most structurally similar to class A enzymes, as it lacks the lid seen in class
C enzymes and the elongated β6/β7 loop that is characteristic of class B enzymes; however,
unlike SaSrtA, BaSrtD reveals an ordered 310 helix within the β6/β7 loop, indicating a rigid
binding pocket for recognition of substrates containing an LPNTA motif (Fig. 3D). This solution
structure reveals a catalytic core formed by residues 56-198, which adopts the conserved eight,
β-stranded sortase fold with two short helices positioned on opposite faces of the protein. The N-
26
approximately 30 amino acids preceding the catalytic core. Interestingly, NMR studies of
BaSrtD also indicate that the β2/β3 and β4/H1 surface loops adjacent to the active site histidine
are structurally disordered, as residues for these regions are completely unassigned, and the
β7/β8 loop following and preceding the active site cysteine (Cys173) and arginine (Arg185),
respectively, is also poorly ordered. BaSrtD was also shown to form dimers in vitro, and these
surrounding the active site may be responsible for associating BaSrtD with its lipid II substrate.
Recently, a 1.99Å resolution crystal structure was solved for Clostridium perfringens SrtD
(CpSrtD), representing only the second characterized enzyme from this class62. The CpSrtD
structure contains eight β-strands, adopting the characteristic β-barrel sortase fold with the
catalytic triad residues (Cys171, His109, and Arg178) positioned within the enzyme’s active site.
Interestingly, based on sequence conservation, this enzyme was originally assigned to the class E
family of sortases, and CpSrtD contains two alpha helices at its N-terminus, a marked difference
from the BaSrtD structure. CpSrtD also exhibits catalytic activity in vitro, which is enhanced in a
magnesium-dependent manner, making it one of only two sortases (along with SaSrtA) known to
be modulated by the presence of metal ions. In further contrast to the BaSrtD enzyme, both the
β2/β3 and β4/H1 loops are well ordered in the CpSrtD crystal structure, with the β2/β3 loop
supported through crystallographic and dynamic light scattering studies. Together, these
observations suggest that BaSrtD and CpSrtD could represent distinct subclasses of the sortase D
family.
27
production of aerial hyphae. These enzymes have been minimally characterized, with the first
and only structure for a class E enzyme determined recently by the Clubb group63. Two sortase
enzymes from Streptomyces coelicolor, ScSrtE1 and ScSrtE2, have been confirmed as bonafide
class E enzymes in that they selectively process the unique LAXTG substrate motif in vitro64.
These enzymes are each predicted to contain a single transmembrane domain, formed by
residues 139-161 in ScSrtE1 and residues 31-53 in ScSrtE2, followed by an extracellular domain
containing the characteristic sortase fold with the conserved catalytic triad (Cys320, His251,
Arg329) positioned within the active site. Although most seemingly related to class C enzymes,
the ScSrtE1 structure reveals a largely disordered N-terminus. ScSrtE1 also maintains an
elongated β6/β7 loop with a short, preformed helix, which suggests a rigid substrate-binding
pocket.
Class F enzymes. Lastly, class F enzymes are very poorly understood, with unknown
function, unknown substrate motif specificity, and no structural characterization to date. Here,
we attempt to gain insight to class F enzymes by analyzing five remaining sortase genes from S.
coelicolor that belong to the class F sortase family7. Two of these five enzymes are predicted to
domains of the remaining three enzymes were identified using sequence alignment with
ClustalW2, and homology models of the extracellular domains of these five class F enzymes
generated with the Phyre 2 server have low sequence coverage, ranging from 69-83%, indicating
their divergence from other sortase classes. These SrtF enzymes are apparently most similar to
class D sortases, as the BaSrtD structure was utilized as the top template in all five models.
Notable features of these SrtF models include a preformed helix within the β6/β7 loop for four
out of five enzymes, as well as a disordered N-terminus that is predicted with high confidence.
28
In addition to structural variations within enzyme classes, some sortases have also been
shown to localize to distinct foci on the membrane66–68, leading to the hypothesis that sortases
may associate with one another on the cell surface69. Several sortases have already been shown
to form dimeric assemblies in vitro61,70, and recently, the functional role of SaSrtA dimerization
was investigated in vivo71. Zhu et al. generated a triple alanine mutant of three residues of
SaSrtA, which was shown to disrupt dimerization in vivo. S. aureus strains expressing this
monomeric SaSrtA showed improved surface display of adhesive proteins, as well as more
efficient invasion of host mammalian cells, suggesting that the monomeric form has increased
activity in vivo. This is in stark contrast to an earlier study which claimed that SaSrtA was more
active when chemically cross-linked to maintain its dimeric form in vitro70. Interestingly, MD
simulations of WT SrtAΔ59 and the monomeric mutant form of the protein showed that the triple
alanine mutation induces a “fixed” β6/β7 loop, and the authors proposed that monomer-dimer
equilibrium regulates the enzymatic activity of SaSrtA most likely by influencing the
conformation of the β6/β7 loop71. Determining the prevalence of in vivo complex formation for
other sortase homodimers and exploring the potential of heterodimeric and higher order
assemblies between different classes of sortases and/or other cell surface translocation or cell
wall synthesis machinery present on the cell surface is an active area of research.
Sorting signal recognition has become one of the most intriguing areas of recent study for
sortase enzymes. Many sortase enzymes have been shown to recognize their substrates with high
specificity and have the ability to differentiate between the appropriate sorting motif or their
substrates and that of other sortase enzymes in vitro and in vivo47,59,72–76. This strict recognition
29
allows the sortases to function non-redundantly to catalyze reactions with specific substrates
even when coexpressed in the same cell. In addition to this strict specificity for primary
substrates, different species of bacteria utilize distinct cross-bridge peptides within their lipid II
cell wall precursor molecule which are typically only recognized as a secondary substrate by the
cell-wall associated sortases for that species. The pilin associated sortases on the other hand are
required to recognize a 4 residue lysine-containing pilin motif located within their protein
substrate. This motif is typically different for different sortases, and is thought to regulate
specificity in systems with multiple class C pilin associated sortases75,76. Understanding the
principles that guide substrate recognition for sortase enzymes continues to be an extremely
difficult area of study. Since sortase enzymes will perform an off target hydrolysis reaction in the
absence of an appropriate nucleophile, slowly cleaving their substrate, structural studies of these
complexes stable enough for structural studies. To date, 4 structures have been solved of
facilitate discussion of the binding interactions that govern recognition of the sorting signal
substrate, we henceforth refer to residues in each sorting signal in relation to their distance from
the scissile bond, where residues in an L-P-X-T-G sorting signal are referred to as P4-P3-P2-P1-
P1’, and the binding sites for those residues on the enzyme as S4-S3-S2-S1-S1’, respectively.
C193A mutant with an LPETG peptide soaked into the crystal32. While the complex appeared to
show the LPETG peptide lying across the surface of the enzyme near the active site, the validity
of this complex was quickly called into question as the highly conserved Leu residue at P4 of the
substrate and the Thr residue at P1 were both seen pointing into solvent. Given that these
30
residues are highly conserved across all substrates for this enzyme6, and that an in vitro analysis
with peptide libraries had shown that both of these residues are extremely important for substrate
recognition73, it was expected that these residues would be recognized by numerous contacts to
the enzyme. Additionally, there was only density for the bound peptide in 1 of the 3 proteins in
the asymmetric unit, indicating extremely weak binding. A breakthrough came when the Jung
containing a sulfhydryl moiety in place of its carboxyl oxygens. This molecule mimics the native
LPXTG substrate, but can form a disulfide bond with the active site cysteine leading to the
NMR analysis of the SaSrtA-LPAT* complex led to the localization of the substrate binding
site78, and eventually the structure of a covalent SaSrtA-LPAT* complex47 (Fig. 4A). This
structure revealed a reorganization of both the β6/β7 loop to include the appearance of a short 310
helix, and an opening of the β7/β8 loop upon binding the substrate, potentially indicating a
regulatory mechanism which would open the loop to expose the active site histidine residue to
solvent and allow for binding of the secondary substrate only after the primary substrate had
15
bound. Additional N relaxation analysis revealed that the β6/β7 loop, which was highly
dynamic in the unbound state, becomes more rigid upon binding of the substrate47. In addition to
significant changes in the active site of the enzyme, the structure also revealed the major residues
that make up the binding pocket for the substrate and showed that the previous peptide-soaked
crystal structure had most likely displayed the peptide non-specifically bound. This non-specific
binding could have either been due to the low affinity of the interaction (Km = 7.33 mM)79, or
due to crystal packing interactions preventing the reorganization of the β6/β7 loop which appears
to be necessary to appropriately construct the binding pocket for recognition of the peptide. One
31
of the most notable features of SaSrtA-LPAT* complex is the large hydrophobic S4 pocket made
up of residues V161, V166, V168, and L169 in the reordered β6/β7 loop, and I199 on strand β8
which binds the P4 Leu residue. The complex also explains the requirement for a Pro residue at
position P3, as this is seen forming a kink in the peptide backbone which directs the rest of the
peptide towards the active site cysteine. Interestingly, the residue in the P2 position which, based
on sequences of known substrates6 and biochemical data73 can be any amino acid, is seen
pointing into the active site where it makes contact with A118 and I182. This space does not
appear to be large enough to accommodate the presence of larger amino acid side chains which
was curious given that these are known to be recognized by the enzyme. Additionally, the P1
threonine residue which is conserved in over 95% of all sortase substrates is seen here pointing
out of the active site towards solvent, only making a single hydrogen bond to the active site
histidine side chain. The positions of the active site residues in this structure are compatible with
biochemical data and the placement of the active site arginine residue hydrogen bonding
exclusively with the backbone carbonyl of the P2 Ala residue seemed to indicate that this
absolutely conserved residue is likely involved in orienting the substrate for catalysis as opposed
to activating the active site Cys or acting as a general base in catalysis as had been proposed
previously.
of its native NPQTN substrate served to confirm several observations from the SaSrtA-LPAT*
structure and clarify some of the more ambiguous ones (Fig. 4B)37. The P4 residue, an
asparagine in this sorting signal, is primarily recognized by contacts to the β6/β7 loop, as in the
SaSrtA-LPAT* structure, although in a more solvent exposed position, concordant with the more
hydrophilic nature of the asparagine sidechain. This is also compatible with biochemical data
32
from the McCafferty group which has shown that replacement of the β6/β7 loop of SaSrtA with
that of SaSrtB is sufficient for altering the specificity of the resultant chimeric protein to the
SaSrtB sorting signal’s NPQTN motif80. The P3 proline residue also appears to provide the same
function as seen in the SaSrtA-LPAT* structure, kinking the peptide to point it towards the
catalytic cysteine. There are also some notable differences between this and the SaSrtA-LPAT*
structure, one of which being that the β6/β7 loop for this enzyme appears to be maintained in a
binding-competent “lock and key” conformation since the structure of the complex was almost
identical to that of the unbound crystal structure (RMSD = 0.44 Å for all Cα). The P2 and P1
residues also appear to be bound in a unique conformation in the SaSrtB-NPQT* complex. The
P2 Gln residue rests along the wall of the pocket, pointing out towards solvent, as opposed to
pointing into the base of the active site as the P2 Ala residue does in the SaSrtA-LPAT*
structure. Interestingly, the P1 Thr residue in the SaSrtB-NPQT* structure is buried in the active
site (Thr-in position) where it forms two hydrogen bonds with the active site arginine residue,
instead of pointing towards solvent and hydrogen bonding with the active site histidine residue as
The third and most recent structure utilizing the same type of disulfide bond forming
substrate mimic, that of BaSrtA-LPAT*, bears similarities to both of the previously described
structures (Fig. 4C)48. The substrate mimic peptide in this structure maintains the same general
orientation as seen in both previous structures providing further evidence that the β6/β7 loop is
instrumental in building the S4 site and dictating specificity for the P4 residue, while the P3 Pro
residue is necessary for providing a kink in the peptide backbone to orient the P1 and P2 residues
toward the active site cysteine. Interestingly, this enzyme substrate complex maintains a Thr-in
conformation, reminiscent of that seen in the SaSrtB-NPQT* substrate complex, where the P1
33
Thr is buried within the binding pocket leading to the active site, but here it does not interact
with the active site arginine. The active site arginine instead forms a hydrogen bond with a
backbone carbonyl in the peptide P3 Pro residue, similar to what was seen in the SaSrtA-LPAT*
structure, albeit in this structure the H-bond was to the adjacent P2 Ala carbonyl. Unlike the
flexible β6/β7 loop found in the unmodified SaSrtA, the β6/β7 loop of BaSrtA was already
shown to be structured in the absence of substrate44. The structure of the modified complex
confirmed the presence of a preformed binding pocket as the backbone atoms for these residues
exhibit an RMSD of 0.76 Å. Conversely, the β7/β8 loop, which was disordered in the
unmodified BaSrtA structure, undergoes a disordered to ordered transition when bound to the
substrate mimic which is correlated with a ~7 Å displacement of the active site cysteine residue
upon binding the peptide. Taken together, these three substrate-mimic bound structures show
conclusively that the S4 site is built by residues in the β6/β7 loop which is essential in
determining the specificity of these enzymes for their P4 residue, and that the highly conserved
Pro residue and P3 is necessary to properly orient the residues at positions P2 and P1 for
interaction with the active site (Fig. 4D). While both the SaSrtB-NPQT* and SaSrtA-LPAT*
structures share an overall similar conformation for the P1 and P2 residues, the interactions seen
between the P1 Thr and the catalytic triad are not consistent across the three structures and likely
indicate internal flexibility inherent to this portion of the active site for some members of this
enzyme family. These inconsistencies could also be caused, or compounded, by potential slight
inaccuracies in complexes with these substrate mimics caused by the use of an unnatural
An energy minimized model of the thioacyl intermediate that exists in the SaSrtB-NPQT
complex provided additional information about the nature of the catalytic mechanism. This
34
model (Fig. 5) shows that the highly conserved P1-Thr residue acts through a hydrogen bonding
network to stabilize the position of the active site arginine residue so that it is maintained in close
proximity to the thioacyl bond. In this position it can act as part of an oxyanion hole, in concert
with the backbone amide of E224, immediately C-terminal to the active site cysteine, to stabilize
the two high-energy tetrahedral intermediates which occur during catalysis37. This novel
specificity for the substrate and thereby prevent unwanted catalysis, as biochemical analysis
shows that the enzyme is intolerant to even conservative mutations (Ser or Val) to the Thr at the
P1 position of its substrate37. A similar energy minimized model was produced for the BaSrtA-
LPAT complex which displayed an analogous hydrogen bonding interaction between the P1 Thr
residue and the active site arginine, along with a similar orientation for the P1 Thr carbonyl in
the thioacyl bond that places it in proximity to the arginine guanidino group for stabilization of
the negatively charged tetrahedral intermediate48. It should be noted however, that the geometry
of these two models is not identical, and that the active site arginine in the BaSrtA structure is in
a more extended structure that allows it to make additional hydrogen bonding interactions with
the backbone carbonyl of the P3 Pro residue, in place of a direct hydrogen bonding interaction
Thr-in conformations, Jacobitz et. al. performed molecular dynamics (MD) simulations for three
use of umbrella sampling calculations, the free energy landscape of transitions between the Thr-
out and Thr-in states was mapped for each complex, and it was found that SaSrtA-LPAT could
adopt both the Thr-in and Thr-out states with equal probability, while for both NPQT and NPAT
35
substrates, SaSrtB only samples the Thr-in state. Based on these results, it appears as if the Thr-
in conformation is likely the more evolutionarily conserved state, and that the inherent flexibility
of SaSrtA allows for the Thr-out conformation with an LPAT substrate that was captured by
Suree et al47.
Additional MD studies with SaSrtA have provided additional insight into the recognition
the sorting signal free and bound states, Kappel et al. proposed that sorting signal binding is a
mixture of conformational selection and induced fit mechanisms81. For example, the β6/β7 loop
conformations in the apo state, some of which were relatively close to the bound configurations.
In contrast, the β7/β8 “open” state from the NMR structure was only stable in the presence of a
bound sorting signal, suggesting an induced fit mechanism. In addition, analysis of the sorting
signal-bound conformations showed that an allosteric network runs throughout the protein,
linking the calcium, sorting signal, and proposed lipid-II binding regions to one another. In a
complementary study, Moritsugu et al. used the multiscale enhanced sampling method to probe
the allosteric effects of the calcium ion and sorting signal82. Simulations of each combination of
bound states showed that binding of both molecules is required to stabilize the β6/β7 and β7/β8
loops in conformations observed in the NMR-LPAT* structure. Overall, these simulations point
towards a mechanism in which calcium, sorting signal, and potentially lipid II binding are
modulated by a dynamic network that includes the β6/β7 loop region in SaSrtA.
Other aspects of the SaSrtA recognition and catalytic processes have also been explored
approach to probe the roles of the conserved sorting signal Leu and Pro residues in substrate
36
binding83. Comparative simulations with LPAT, APAT, and LAAT substrates demonstrated that
contacts between the leucine sidechain and SaSrtA help to stabilize the β6/β7 loop, whereas the
kink that is induced by the proline appears to be essential for recognition. In another study, Tian
and Eriksson performed simulations in which His120 and Cys184 were in their zwitterionic and
neutral forms84. Their results showed that Arg197 adopts distinct conformations based upon the
charged state of the protein, which helps to stabilize the catalytically active form. It should be
noted that each of these studies was performed with the sorting signal in the Thr-out state.
Although the global effects of the Thr-in and Thr-out states on the induced fit/conformational
selection process, allosteric networks, and recognition processes are likely similar, subtle
differences may exist that influence some of the fine details that resulted from these simulations.
When compared to the recent accumulation of information regarding the binding of the
primary sorting signal substrate, there is relatively little known about the location of the binding
Progress on this front has been slowed most significantly by the poor solubility and difficulty
inherent in obtaining large quantities of the intact lipid II substrate. To date, the most convincing
line of evidence stems from NMR chemical shift mapping of the SaSrtA-LPAT* complex with a
triglycine peptide titrated into the sample to mimic the pentaglycine cross-bridge present in the S.
aureus lipid II molecule by Suree et al47. Analysis of the chemical shift perturbations resulting
from this addition revealed a continuous surface spanning portions of the β7/β8 loop, β4/H2
loop, and an N-terminal segment of helix H1. While this surface lies near the active site and
bound peptide, it is too large to definitively assign a single binding pocket to the secondary
substrate. It is also interesting to note that these chemical shift changes were not observed when
15
titrating the triglycine peptide against an unbound N labeled SaSrtA but only occur after the
37
SaSrtA-LPAT* complex has formed, presumably due to a rearrangement of the β7/β8 loop
which creates a more open conformation necessary for substrate binding in this enzyme. An X-
ray structure of SaSrtB bound to MTSET, a sulfhydryl modifier, with a triglycine peptide soaked
into the crystal has also been solved and shows the triglycine peptide bound exclusively to the
β7/β8 loop85. While this is compatible with NMR chemical shift data from the SaSrtA-LPAT*
complex, it does not fit the accepted view of the mechanism which would have the incoming
nucleophile deprotonated by the active site histidine residue. In this complex the N-terminal
amine of the triglycine peptide is 6.4 Å from the active site histidine which is occluded from
solvent by a closed β7/β8 loop. This is also incompatible with the clear groove extending past the
active site histidine and cysteine residues visible in SpySrtA, SagSrtA, and SmSrtA structures
that has been hypothesized to be the binding groove for the secondary substrate in these
enzymes43,45,46. Confidently discerning the true location of the secondary substrate binding site
The culmination of the past 16 years of research since the discovery of the sortase
enzymes has been a relatively complete understanding of the molecular mechanism, compiled
from a large number of studies encompassing in vitro biochemical data, mutagenesis, X-ray and
While there will still undoubtedly be variations, improvements, and exceptions added to the
current understanding of the mechanism, the existing model is backed by a large and ever
growing body of evidence. The current model of the molecular mechanism is presented in Figure
6. Kinetic studies suggest that catalysis occurs through a ping-pong mechanism that begins when
38
the sortase recognizes the CWSS of a membrane anchored protein as it binds in a groove made
by residues in the β6/β7 loop, strands β4, β7, β8, and closed by the β2/β3 and β3/β4 loops (Fig.
4D). Here the sorting signal’s L-shaped structure dictated by the highly conserved proline
residue at P3 (>90% conserved) will orient residue P4 for recognition in subsite S4 at the β6/β7
loop, as the remainder of the peptide is directed deeper into the active site. In order for catalysis
to proceed, the enzyme must contain a properly charged active site with cysteine and histidine
residues in their thiolate and imidazolium ion forms, respectively. Based on pKa measurements
of the SaSrtA79,86 and BaSrtA44 enzymes, and the similarly slow in vitro catalytic rates of all
known sortases studied to date, this is expected to comprise <1% of the available population of
enzymes at physiological pH. This is also confirmed by structural analysis of SaSrtA wherein
none of the residues in the active site are close enough to support the presence of any ion pairing,
with a Cys-S-His-δN distance of 6.5 and 7.6 Å for NMR and crystal structures, respectively. If
the cysteine residue is appropriately charged it can perform a nucleophilic attack on the P1 Thr’s
carbonyl carbon, leading to the formation of the first tetrahedral intermediate, which is likely
coordinated by the substrate-stabilized oxyanion hole postulated to exist between the active site
arginine residue and a backbone amide in the β7/β8 loop37. This intermediate quickly collapses
to form a semi-stable thioacyl intermediate between the substrate’s P1-Thr and the active site
cysteine, while the leaving peptide is protonated by the active site histidine residue. The active
site arginine residue may also participate in the reaction at this stage and others by hydrogen
bonding with the backbone carbonyls of the bound substrate to maintain its orientation in the
active site47,48,84. This interaction is potentially dependent on the position (Thr-in or Thr-out) of
the bound P1 Thr residue and also on the particular sortase in question as existing structures of
sortase-substrate complexes have depicted multiple binding positions for this residue and
39
molecular dynamics (MD) studies have indicated that it can sample multiple conformations in
some enzymes which could potentially influence the position of the active site arginine residue37.
In the next step of the reaction, the secondary substrate enters the active site where its amine
group is thought to be deprotonated by the active site histidine residue before performing a
nucleophilic attack on the thioacyl bond. This results in a second tetrahedral intermediate which
tetrahedral intermediate quickly collapses to form a new peptide bond between the incoming
nucleophile and the P1-Thr residue which then exits the active site to be incorporated into the
cell wall via the standard cell wall synthesis machinery, or added to additional protein subunits in
Many of the sortase enzymes have been studied using a variety of different methods in
vitro. The first demonstration of in vitro activity was performed by Ton-That and coworkers
using the prototypical SaSrtA enzyme87. Their assay was the first of many to utilize FRET based
reporters of activity by incorporating donor and quencher fluorophores onto both ends of an
LPXTG sorting signal peptide. Cleavage of this peptide, either through hydrolysis,
hydroxylaminolysis, or the native transpeptidation reaction, liberates the donor from the
quencher and this activity can be detected by an increase in fluorescence. While the initial study
demonstrated only cleavage and hydroxylaminolysis, subsequent studies also demonstrated the
ability of SaSrtA to catalyze the native transpeptidation reaction in vitro using only short peptide
substrates where one contains a sorting signal motif, and the other a series of 1 to 5 glycines88,89.
While this assay was useful for demonstrating a preference for secondary substrates containing at
40
least 2 glycines at the N-terminus89, and for initial characterization of the active site histidine
cysteine and arginine residues22,23, early kinetics analysis was found to be plagued by the inner
filter effect in which very high concentrations of substrate will lead to a donor quenching effect
even after cleavage of the FRET pair, leading to inaccuracies in kinetic measurements90.
Development of an HPLC assay and subsequent revision of the kinetic parameters based on
fitting the parameters to a ping-pong hydrolytic shunt model to more accurately account for loss
of activity due to hydrolysis of the peptide revealed relatively weak binding affinity for the
LPETG sorting signal (Km = 7.33 ± 1.01 mM) and secondary Gly5 substrate (Km = 196 ± 64 µM)
and low kcat of 0.28 ± 0.02 s-1 79. Although this is a fairly low turnover number, the measured
pKas of the active site residues indicate that the majority of the enzyme is actually reverse-
protonated, with only a small percentage (~0.06%) of the enzyme actually in the appropriately
charged, active form in vitro. Accounting for this low population of appropriately charged
enzyme indicates that this subpopulation of the enzyme must actually be extremely active, with a
kcat of over 105 M-1 79. It has also been noted that the reaction occurs extremely rapidly (<3 min
from expression to surface attachment) in vivo19 indicating that there are likely other factors
involved in the reaction that are not adequately replicated in the in vitro systems described to
date. This effect could simply be due to the reduction of the search problem from 3 dimensions
to 2 by the incorporation of membrane anchors to both sortase and substrate, or the existence of
as yet uncharacterized interactions between other cell surface factors that either activate the
sortase or modulate the positions of protein and substrate for more efficient catalysis. Kinetic
analysis has also been conducted on several other sortase enzymes, but these have all reported
even slower rates of in vitro catalysis indicating a similar mechanism which is likely plagued by
similar shortcomings in all simplified in vitro systems43,44,80. A number of additional studies have
41
measured the endpoints of reactions at a defined incubation time and lack the comprehensive
analysis necessary to extract kinetic parameters. A summary of the results of in vitro reactions
with sortases performed to date is available in Table 2. The fact that many of these analyses are
highly simplified, along with numerous differences in substrate and buffer requirements for the
various members of the sortase family have made a comparison of the various kinetic rates of
target for which inhibitor development is being actively pursued91. Many classes of sortase
inhibitors have been identified with high-throughput screening (HTS) via monitoring FRET-
based substrate cleavage in vitro or virtual docking of small molecule libraries against sortase
structures in silico91. Several sortase inhibitor classes function through covalent modification of
the active site cysteine, resulting in irreversible inactivation of the enzyme, and several Cys-
structures display sortase enzymes covalently linked with general thiol modifiers; however,
several structures exhibit complexes with novel, sortase-specific compounds that provide insight
into active site interactions and potential derivatization strategies. Furthermore, structure-activity
relationship (SAR) and molecular docking studies for several lead compounds have facilitated
bringing a novel class of therapeutics for Gram-positive infections closer within reach.
Maresso et al. identified AAEK inhibitors using an in vitro FRET-based HTS, and
subsequently, solved the crystal structures of BaSrtB linked to AAEK1 and AAEK2 inhibitors13.
42
Sortases specifically activate AAEK inhibitors via a β-elimination reaction that generates an
olefin intermediate which covalently modifies the active site cysteine. AAEK compounds
currently inhibit different class sortases from staphylococci and bacilli species with IC50s in the
low micromolar range (~5-50 µM). A comparison of the structures of native and inhibitor-bound
BaSrtB reveals several key differences, including rotation of the active site cysteine by ~180° to
accommodate the ligand, displacement of the conserved arginine away from active site, and
higher mobility and different conformations of residues from the H5/β6 loop. Interestingly, the
aryl group of the thienylpropanone adduct engages in stacking interactions with a critical
tyrosine (Tyr138) within the BaSrtB active site. Furthermore, the inhibitor-bound form reveals
two available binding sites, a cationic site above and anionic site below the aryl group that could
To identify more potent SrtA inhibitors, Suree et al. performed an in vitro HTS of a
30,000 compound library that resulted in three promising leads, the rhodanine, pyridazinone, and
pyrazolethione classes11. All three lead classes inhibited SaSrtA and BaSrtA reversibly with
IC50s in the low to sub-micromolar range. SAR studies of the most promising class, the
pyridazinone lead, identified critical regions of the chemical scaffold that are required for sortase
inhibition. The binding modes of these inhibitors were modeled using induced-fit docking
protocol and the SaSrtA-LPAT* structure. In this model, the pyridazinone phenyl ring is buried
in a hydrophobic pocket consisting of side chains from Ile199 on the β8 strand and Val166 and
Val168 from the β6/β7 loop of SaSrtA, while the carbonyl oxygen on the pyridazinone ring is
positioned towards the active site Arg197 side chain. The inhibitor’s thiol group points towards
the conserved His120, and mutation with a chloro group significantly reduces activity.
43
Interestingly, mutating the R1 substituent to an ethoxy moiety is well-tolerated, presumably due
to additional hydrophobic contacts with side chains of H1 residues, Pro94 and Ala92.
To improve the success rate of virtual screening against the substrate-bound form of
SaSrtA, Chan et al. virtually screened over 55,000 ligands, the top-ranked 500 of which were
docked onto the NMR structure of SaSrtA with a relaxed complex scheme to account for protein
receptor flexibility12. This approach identified 24 unique leads, and in vitro validation of eight
out of 19 compounds yielded a 42.2% success rate with IC50s ranging from 47-368 µM.
compound experimentally, did not rank highly during the first round of virtual screening, but
ranked substantially higher when docked to the ensemble of molecular dynamics (MD) simulated
sortase conformers. This compound appears to exploit regions of the enzyme that have evolved
for substrate binding, as computational docking studies suggest that the DHP group within the 2-
phenyl-2,3-dihydro-1H-perimidine lead scaffold, will bind in the S4 site adjacent to the β6/β7
loop. Interestingly, the naphthalene ring moves within the hydrophobic pocket during MD
simulations, suggesting that adding non-polar substituents here could improve binding. Finally,
the phenyl group of the lead scaffold is positioned towards the conserved catalytic triad residues
and the attached carboxyl moiety is wedged between and the active site histidine and arginine
inhibitor bound to SaSrtA using NMR92. Secondary screening of inhibitor binding by HSQC
provided unique benefits, including identifying interacting sortase residues, estimating potency
based on adduct accumulation, and determining reversibility of binding, and also highlighted the
44
susceptibility of in vitro screening to yield false positives (11 of 41 compounds showed no
complex was then used to design analog molecules, which irreversibly inhibited SaSrtA with
IC50s in the low micromolar range (~3-7 µM). The NMR structure indicates various inhibitor
conformations, although interacting groups are nearly identical, which is possibly due to the
dynamic nature of the β6/β7 and β7/β8 loops. The high affinity interaction is facilitated by
side chain; however, positions 4, 5, and 6 of this moiety did not show contacts with SaSrtA,
suggesting that derivatization to larger substituents could facilitate additional contacts. Within
the structure ensemble, the phenyl group of the inhibitor adduct can interact with L97 and A118
or participate in pi-pi stacking interactions with W194. The linker region preceding the
adamantyl moiety was varied to contain one or two carbonyl groups, which may hydrogen bond
with the active site arginine. In addition, the length of the linker region could be varied to
optimally bury the adamantyl moiety within the enzyme’s hydrophobic pocket, where it was
experimentally determined to interact with V166, V168, L169 in the β6/β7 loop and T180, I182,
Recently, the structure of SmSrtA was solved with the natural product flavonoid
precursor, trans-chalcone46. The structure depicts chalcone bound outside of the SmSrtA active
site; however, this compound was verified biochemically to covalently modify the active site
cysteine residue through Michael addition. Wallock-Richards et al. generated a model of the
and BaSrtB-AAEK. The fortuitous interaction of the N-terminus from the symmetry-related
molecule within the SmSrtA active site guided overlay of the aromatic rings of chalcone at the
45
phenyl rings of Phe67 and Phe69, which presumably represents the binding mode of this
compound. This binding mode seems to be shared among other sortase inhibitor classes,
two phenyl groups is essential for SrtA inhibition93. In fact, the criteria for two aromatic regions
seems to be a common feature of SrtA inhibitors and has been implemented to filter compounds
using a pharmacophore model generated from all SrtA inhibitors available in the literature before
The wealth of structural information available for sortase enzymes is evolving virtual screening
strategies to account for the dynamic nature of SrtA by implementing docking against multiple
strategies, in combination with SAR of lead compounds, have produced many promising
inhibitor classes with IC50s reaching the low- and sub-micromolar range11,15, as well as informed
our understanding of important active site contacts that mediate binding affinity and exploited
2.10 Conclusion
associated with increased pathogenicity. Beyond their medical relevance, these enzymes have
also been shown to be extremely useful tools for site specifically linking proteins and small
molecules in vitro. Because of this, a considerable effort has been put forth to understand the
structural features that dictate substrate and small molecule inhibitor binding, as well as the
46
substrate binding, shed light on the importance of variable regions of sortase structure, elucidated
the molecular mechanism, and produced a number of small molecule sortase inhibitors. In spite
of the many discoveries that have been made in the last decade, many questions remain to be
answered. No conclusive evidence has emerged to pinpoint the secondary substrate binding site
and very little is known about this latter half of the mechanism. While there are a number of
hypotheses, each with varying degrees of experimental backing, there has been no consensus in
the field. It is conceivable that this lack of consensus is the result of individual enzymes or
specific classes recognizing their secondary substrates in very different ways. This could
potentially stem from the needs of certain classes to attach their primary protein substrates to
different locations on the cell surface or recognize a secondary protein substrate instead of lipid
II. Exploration of this interaction in a number of different enzymes will be necessary to draw
useful conclusions about the nature of secondary substrate recognition. A number of enzymes
have also been shown to harbor N-terminal extensions which in some cases contact or
completely occlude the active site. For class C enzymes it appears that lid removal must be a
necessary step in uncovering the active site for catalysis, and while structures of class C enzymes
with their lids displaced have emerged recently, it remains to be seen whether these structures are
physiologically relevant, what would be responsible for removing the lid in vivo, and how this lid
removal would take place. Understanding the various ways sortase enzymes utilize these
additional features in substrate recognition and catalysis is an area of active research. Finally,
although many small molecule sortase inhibitors have been identified, to date none has been
successful enough to move forward towards becoming a therapeutic for treating Gram-positive
infections in humans. Given the rising prevalence of antibiotic resistant bacteria in recent years,
47
pressure to develop viable sortase inhibitors as therapeutics is rising and will undoubtedly lead to
the discovery and characterization of more potent and specific compounds in the future.
48
2.11 Figures
Sortase and substrate are both membrane bound. 1 Sortase recognizes a sorting signal motif (here
the LPXTG sorting signal for SrtA types is shown) within CWSS and performs a nucleophilic
Mechanism for cell wall anchoring proceeds through “a” steps (top) and mechanism for pilus
assembly proceeds through “b” steps (bottom). 3a Cross-bridge peptide from lipid II molecule
performs a nucleophilic attack on the thioacyl intermediate. 4a new peptide bond is formed
between the lipid II molecule and surface protein which is then incorporated into mature cell
wall. 3b a lysine residue within the pilin motif performs nucleophilic attack on thioacyl
intermediate. 4b covalently linked pilin proteins can either be attached to cell wall as in 3a and
49
Figure 2.2 Structure of sortase.
(A) SaSrtA NMR structure shown as cartoon, showcasing 8 Stranded β-barrel, with active site
residues shown as sticks. (B) SaSrtA NMR structure shown as a surface representation in green
with active residues Arg in blue and Cys in yellow. The active site His is occluded by a closed
β7/β8 loop, and there is no obvious groove for a full length peptide to exit the active site. (C)
SpySrtA structure shown as a surface representation, colored as in (B), with active site His
residue also shown in cyan. An open β7/β8 loop creates a clear channel that can be seen running
between active Cys and His residues indicating the likely path of binding for a full length peptide
substrate. (D) Sortase can catalyze both a reversible transpeptidation reaction, and an irreversible
50
Table 2.1 Structurally characterized sortase enzymes.
Organism and Sortase PDB Bound Ligands or Substrates Method
Class A
Class B
51
Class C
Class D
Class E
52
Figure 2.3 Structural variation by class of sortase.
Sortases representative of the major themes seen for each class are displayed as a cartoon
representation, with active site residues shown as sticks. The hallmark sortase beta barrel is
highlighted in blue. Major sources of structural variability are highlighted: N-terminus, red;
53
Figure 2.4 Sorting signal recognition.
(A) The SaSrtA-LPAT* complex. (B) The SaSrtB-NPQT* complex. (C) The BaSrtA-LPAT*
complex, shown with N-terminal appendage removed from view for clarity. Enzymes are shown
as surface representations with SrtA types in light green and SaSrtB in light blue, Substrate
mimics are shown as grey sticks. Active site Cys and Arg residues are shown as gold and blue
surfaces, respectively. (D) Conserved recognition sites for sortase enzymes. Left, SaSrtA shown
green, and S1 in magenta, and active site Arg in blue, Cys in gold, and His in cyan. Right,
Cartoon diagram of SaSrtA with secondary structure elements that contribute to substrate binding
54
Figure 2.5 The substrate stabilized oxyanion hole.
The energy minimized model of the SaSrtB-NPQT thioacyl intermediate displayed with SaSrtB
as light blue cartoon and residues in the active site and oxyanion hole shown as sticks. NPQT
substrate shown as grey sticks. The substrate’s P1 Thr residue’s sidechain hydroxyl, and
backbone carbonyl participate in a hydrogen bonding network with the active site Arg, and the
backbone amide of Glu224 that together build an oxyanion hole to stabilize the high energy
55
Figure 2.6 Molecular mechanism of sortase enzymes.
The active site of sortase consists of a His-Cys-Arg triad, and in its active form, the His and Cys
residues will form a thiolate-imidazolium ion pair (a). The first step in the reaction is the
recognition of an appropriate sorting signal (here the LPXTG sorting signal for SrtA types is
shown), and the active site Cys residue will perform a nucleophilic attack on the carbonyl carbon
at the substrate’s P1 position (b). This leads to a tetrahedral intermediate whose oxyanion is
stabilized by the nearby Arg residue which is likely held in proximity by interactions with the
side chain of the substrate’s P1 residue which is a threonine in over 95% of all substrates (c). The
active His residue concomitantly donates a proton to the leaving group and the tetrahedral
transition state then collapses to form a semi-stable thioacyl intermediate between the substrate’s
P1 residue and the active site Cys (d). Next, the secondary substrate, (Here shown as the lipid II
molecule used by cell wall anchoring sortases) enters the active site where its terminal amine is
56
deprotonated by the active His residue before it performs a nucleophilic attack on the carbonyl
carbon in the thioacyl bond (e) forming a second tetrahedral intermediate (f) before collapsing to
form a new peptide bond between the two substrates, which is then released to leave the
57
Table 2.2 Activity of sortase enzymes in vitro.
Class A
Class B
58
Class C
Class D
“Yes” indicates the reaction was performed in vitro but kinetics parameters were not reported
NR – Not Reported
† Values reported from fluorescence assay and subject to inner filter effect and are likely underestimates
of true parameters
59
2.12 References
1. Liew PX, Wang CLC, Wong S-L. Functional characterization and localization of a
Bacillus subtilis sortase and its substrate and use of this sortase system to covalently
anchor a heterologous protein to the B. subtilis cell wall for surface display. J Bacteriol.
2012;194(1):161-175. doi:10.1128/JB.05711-11.
2. Paterson GK, Mitchell TJ. The biology of Gram-positive sortase enzymes. Trends
enzyme that anchors surface proteins to the cell wall. Science. 1999;285(5428):760-763.
http://www.ncbi.nlm.nih.gov/pubmed/10427003.
doi:10.1128/IAI.72.5.2710-2722.2004.
7. Spirig T, Weiner EM, Clubb RT. Sortase enzymes in Gram-positive bacteria. Mol
60
8. Suree N, Jung ME, Clubb RT. Recent advances towards new anti-infective agents that
inhibit cell surface protein anchoring in Staphylococcus aureus and other gram-positive
doi:10.2174/138955707782110097.
sortase mutants defective in the display of surface proteins and in the pathogenesis of
doi:10.1073/pnas.080520697.
10. Weiss WJ, Lenoy E, Murphy T, et al. Effect of srtA and srtB gene expression on the
11. Suree N, Yi SW, Thieu W, et al. Discovery and structure-activity relationship analysis of
doi:10.1016/j.bmc.2009.08.067.
12. Chan AH, Wereszczynski J, Amer BR, et al. Discovery of Staphylococcus aureus sortase
A inhibitors using virtual screening and the relaxed complex scheme. Chem Biol Drug
13. Maresso AW, Wu R, Kern JW, et al. Activation of inhibitors by sortase triggers
http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3366505&tool=pmcentrez&re
61
14. Maresso AW, Schneewind O. Sortase as a target of anti-infective therapy. Pharmacol Rev.
2008;60(1):128-141. doi:10.1124/pr.107.07110.
15. Zhang J, Liu H, Zhu K, et al. Antiinfective therapy with a small molecule inhibitor of
doi:10.1073/pnas.1408601111.
16. Proft T. Sortase-mediated protein ligation: an emerging biotechnology tool for protein
009-0116-0.
doi:10.1002/cbic.200800724.
18. Mao H, Hart SA, Schink A, Pollok BA. Sortase-mediated protein ligation: a new method
19. Schneewind O, Model P, Fischetti VA. Sorting of protein A to the staphylococcal cell
SrtC2 required for membrane localization and assembly of type 2 fimbriae for
2539. doi:10.1128/JB.00093-12.
62
21. Cozzi R, Malito E, Nuccitelli A, et al. Structure analysis and site-directed mutagenesis of
defined key residues and motives for pilus-related sortase C1 in group B Streptococcus.
the cell wall of Staphylococcus aureus. Cysteine 184 and histidine 120 of sortase form a
doi:10.1074/jbc.M109945200.
surface proteins to the cell wall of Staphylococcus aureus. A conserved arginine residue is
doi:10.1074/jbc.M405282200.
24. Clancy KW, Melvin JA, McCafferty DG. Sortase transpeptidases: insights into
doi:10.1002/bip.21472.
Mechanisms of Their Targeting to the Cell Wall Envelope. Microbiol Mol Biol Rev.
http://www.ncbi.nlm.nih.gov/pubmed/11401711.
63
27. Telford JL, Barocchi MA, Margarit I, Rappuoli R, Grandi G. Pili in Gram-positive
2008;16(1):33-40. doi:10.1016/j.tim.2007.10.010.
29. Danne C, Dramsi S. Pili of gram-positive bacteria: roles in host colonization. Res
30. Ilangovan U, Ton-That H, Iwahara J, Schneewind O, Clubb RT. Structure of sortase, the
transpeptidase that anchors proteins to the cell wall of Staphylococcus aureus. Proc Natl
31. Andreeva A, Howorth D, Chothia C, Kulesha E, Murzin AG. SCOP2 prototype: a new
D314. doi:10.1093/nar/gkt1242.
32. Zong Y, Bice TW, Ton-That H, Schneewind O, Narayana SVL. Crystal structures of
2004;279(30):31383-31389. doi:10.1074/jbc.M401374200.
33. Naik MT, Suree N, Ilangovan U, et al. Staphylococcus aureus Sortase A transpeptidase.
Calcium promotes sorting signal binding by altering the mobility and structure of an
64
34. Schneewind O, Fowler A, Faull KF. Structure of the cell wall anchor of surface proteins in
http://www.ncbi.nlm.nih.gov/pubmed/7701329.
35. Perry AM, Ton-That H, Mazmanian SK, Schneewind O. Anchoring of surface proteins to
the cell wall of Staphylococcus aureus. III. Lipid II is an in vivo peptidoglycan substrate
doi:10.1074/jbc.M109194200.
proteins. A branched peptide that links the carboxyl terminus of proteins to the cell wall. J
37. Jacobitz AW, Wereszczynski J, Yi SW, et al. Structural and computational studies of the
38. Kang HJ, Coulibaly F, Proft T, Baker EN. Crystal Structure of Spy0129, a Streptococcus
doi:10.1371/journal.pone.0015969.
65
41. Ton-That H, Schneewind O. Assembly of pili on the surface of Corynebacterium
42. Elliot MA, Karoonuthaisiri N, Huang J, et al. The chaplins: a family of hydrophobic cell-
43. Race PR, Bentley ML, Melvin JA, et al. Crystal Structure of Streptococcus pyogenes
44. Weiner EM, Robson S, Marohn M, Clubb RT. The Sortase A enzyme that attaches
proteins to the cell wall of Bacillus anthracis contains an unusual active site architecture. J
45. Khare B, Krishnan V, Rajashankar KR, et al. Structural differences between the
Streptococcus agalactiae housekeeping and pilus-specific sortases: SrtA and SrtC1. PLoS
47. Suree N, Liew CK, Villareal VA, et al. The Structure of the Staphylococcus aureus
66
Signal Is Recognized. J Biol Chem. 2009;284(36):24465-24477.
doi:10.1074/jbc.M109.022624.
48. Chan AH, Yi SW, Terwilliger AL, Maresso AW, Jung ME, Clubb RT. Structure of the
Bacillus anthracis Sortase A Enzyme Bound to Its Sorting Signal: a Flexible Amino-
49. Zong Y, Mazmanian SK, Schneewind O, Narayana SV. The Structure of Sortase B, a
Cysteine Transpeptidase that Tethers Surface Protein to the Staphylococcus aureus Cell
50. Mazmanian SK, Skaar EP, Gaspar AH, et al. Passage of Heme-Iron Across the Envelope
doi:10.1126/science.1081147.
2005;280(16):16263-16271. doi:10.1074/jbc.M500071200.
52. Chambers CJ, Roberts AK, Shone CC, Acharya KR. Structure and function of a
53. Shaik MM, Lombardi C, Maragno Trindade D, et al. A structural snapshot of type II pilus
doi:10.1074/jbc.M115.647834.
67
54. Manzano C, Contreras-Martel C, El Mortaji L, et al. Sortase-mediated pilus fiber
1848. doi:10.1016/j.str.2008.10.007.
55. Manzano C, Izoré T, Job V, Di Guilmi AM, Dessen A. Sortase activity is controlled by a
2009;48(44):10549-10557. doi:10.1021/bi901261y.
56. Khare B, Fu Z-Q, Huang I-H, Ton-That H, Narayana SVL. The crystal structure analysis
of group B Streptococcus sortase C1: a model for the “lid” movement upon substrate
57. Cozzi R, Prigozhin D, Rosini R, et al. Structural Basis for Group B Streptococcus Pilus 1
doi:10.1371/journal.pone.0049048.
58. Persson K. Structure of the sortase AcSrtC-1 from Actinomyces oris. Acta Crystallogr D
59. Neiers F, Madhurantakam C, Fälker S, et al. Two crystal structures of pneumococcal pilus
sortase C provide novel insights into catalysis and substrate specificity. J Mol Biol.
2009;393(3):704-716. doi:10.1016/j.jmb.2009.08.058.
60. Lu G, Qi J, Gao F, Yan J, Tang J, Gao GF. A novel “open-form” structure of sortaseC
68
61. Robson SA, Jacobitz AW, Clubb RT, Phillips ML, Clubb RT. Solution Structure of the
62. Suryadinata R, Seabrook SA, Adams TE, Nuttall SD, Peat TS. Structural and biochemical
63. Kattke M, Jacobitz AW, Clubb RT. The Structure of the SrtE1 Sortase from Streptomyces
doi:10.1111/j.1365-2958.2012.07983.x.
65. Krogh A, Larsson B, von Heijne G, Sonnhammer EL. Predicting transmembrane protein
topology with a hidden Markov model: application to complete genomes. J Mol Biol.
2001;305(3):567-580. doi:10.1006/jmbi.2000.4315.
66. Raz A, Fischetti VA. Sortase A localizes to distinct foci on the Streptococcus pyogenes
doi:10.1073/pnas.0808301105.
67. Hu P, Bian Z, Fan M, Huang M, Zhang P. Sec translocase and sortase A are colocalised in
2008;53(2):150-154. doi:10.1016/j.archoralbio.2007.08.008.
69
68. Kline KA, Kau AL, Chen SL, et al. Mechanism for sortase localization and the role of
2009;191(10):3237-3247. doi:10.1128/JB.01837-08.
69. Mandlik A, Das A, Ton-That H. The molecular switch that activates the cell wall
anchoring step of pilus assembly in gram-positive bacteria. Proc Natl Acad Sci U S A.
2008;105(37):14147-14152. doi:10.1073/pnas.0806350105.
70. Lu C, Zhu J, Wang Y, et al. Staphylococcus aureus sortase A exists as a dimeric protein in
Staphylococcus aureus cell surface mediates its cell wall sorting activity. Exp Biol Med
72. Donahue EH, Dawson LF, Valiente E, et al. Clostridium difficile has a single sortase,
doi:10.1186/s12866-014-0219-1.
73. Kruger RG, Otvos B, Frankel BA, Bentley M, Dostal P, McCafferty DG. Analysis of the
74. Fälker S, Nelson AL, Morfeldt E, et al. Sortase-mediated assembly and surface topology
2958.2008.06396.x.
70
75. El Mortaji L, Terrasse R, Dessen A, Vernet T, Di Guilmi AM. Stability and Assembly of
doi:10.1074/jbc.M109.082776.
76. El Mortaji L, Fenel D, Vernet T, Di Guilmi AM. Association of RrgA and RrgC into the
352. doi:10.1021/bi201591n.
77. Jung ME, Clemens JJ, Suree N, et al. Synthesis of (2R,3S) 3-amino-4-mercapto-2-butanol,
a threonine analogue for covalent inhibition of sortases. Bioorg Med Chem Lett.
2005;15(22):5076-5079. doi:10.1016/j.bmcl.2005.07.073.
78. Liew CK, Smith BT, Pilpa R, et al. Localization and mutagenesis of the sorting signal
226. doi:10.1016/j.febslet.2004.06.070.
79. Frankel BA, Kruger RG, Robinson DE, Kelleher NL, McCafferty DG. Staphylococcus
aureus sortase transpeptidase SrtA: insight into the kinetic mechanism and evidence for a
doi:10.1021/bi050141j.
80. Bentley ML, Gaweska H, Kielec JM, McCafferty DG. Engineering the substrate
specificity of Staphylococcus aureus Sortase A. The beta6/beta7 loop from SrtB confers
doi:10.1074/jbc.M610519200.
71
81. Kappel K, Wereszczynski J, Clubb RT, McCammon JA. The binding mechanism,
doi:10.1002/pro.2168.
83. Biswas T, Pawale VS, Choudhury D, Roy RP. Sorting of LPXTG peptides by archetypal
sortase A: role of invariant substrate residues in modulating the enzyme dynamics and
84. Tian B-X, Eriksson LA. Catalytic Mechanism and Roles of Arg197 and Thr183 in the
doi:10.1021/jp2058113.
85. Zong Y, Mazmanian SK, Schneewind O, Narayana SVL. The Structure of Sortase B, a
86. Connolly KM, Smith BT, Pilpa R, Ilangovan U, Jung ME, Clubb RT. Sortase from
Staphylococcus aureus Does Not Contain a Thiolate-Imidazolium Ion Pair in Its Active
72
87. Ton-That H, Liu G, Mazmanian SK, Faull KF, Schneewind O. Purification and
1999;96(22):12424-12429.
http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=22937&tool=pmcentrez&rend
88. Ton-That, H., Mazmanian, S. K., Alksne, L., and Schneewind O. Anchoring of Surface
chromatography assay and revision of kinetic parameters for the Staphylococcus aureus
doi:10.1016/j.ab.2003.10.023.
91. Cascioferro S, Totsika M, Schillaci D. Sortase A: an ideal target for anti-virulence drug
73
Staphylococcus aureus sortase A transpeptidase. Bioorg Med Chem. 2014;22(21):5988-
6003. doi:10.1016/j.bmc.2014.09.011.
93. Oh K-B, Kim S-H, Lee J, Cho W-J, Lee T, Kim S. Discovery of diarylacrylonitriles as a
doi:10.1021/jm0498708.
94. Raj KK, Ganesh Kumar V, Leela Madhuri C, et al. Designing of potential inhibitors
doi:10.1016/j.jmgm.2015.05.009.
95. Uddin R, Saeed K. An exhaustive yet simple virtual screening campaign against Sortase A
from multiple drug resistant Staphylococcus aureus. Mol Biol Rep. 2014;41(8):5167-5175.
doi:10.1007/s11033-014-3384-2.
96. Bentley ML, Lamb EC, McCafferty DG. Mutagenesis Studies of Substrate Recognition
and Catalysis in the Sortase A Transpeptidase from Staphylococcus aureus. J Biol Chem.
2008;283(21):14762-14771. doi:10.1074/jbc.M800974200.
97. Chen I, Dorr BM, Liu DR. A general strategy for the evolution of bond-forming enzymes
doi:10.1073/pnas.1101046108.
74
98. Maresso AW, Chapa TJ, Schneewind O. Surface protein IsdC and sortase B are required
doi:10.1128/JB.01011-06.
99. Van Leeuwen HC, Klychnikov OI, Menks MAC, Kuijper EJ, Drijfhout JW, Hensbergen
PJ. Clostridium difficile sortase recognizes a (S/P)PXTG sequence motif and can
2014;588(23):4325-4333. doi:10.1016/j.febslet.2014.09.041.
100. Marraffini L a, Schneewind O. Targeting proteins to the cell wall of sporulating Bacillus
75
Chapter 3
76
The work described in this chapter has been reproduced with permission from:
Robson, S. A., Jacobitz, A. W., Clubb, R. T., Phillips, M. L., Clubb, R. T. Solution Structure of
the Sortase Required for Efficient Production of Infectious Bacillus anthracis Spores.
doi:10.1021/bi300867t.http://pubs.acs.org/doi/abs/10.1021/bi300867t
Society
structural and genetic analysis between the sortase structure presented here and existing sortase
enzymes.
77
3.1 Overview
Bacillus anthracis forms metabolically dormant endospores that upon germination can
cause lethal anthrax disease in humans. Efficient sporulation requires the activity of the SrtC
sortase (BaSrtC), a cysteine transpeptidase that covalently attaches the BasH and BasI proteins to
the peptidoglycan of the forespore and predivisional cell, respectively. To gain insight into the
molecular basis of protein display, we used nuclear magnetic resonance to determine the
structure and backbone dynamics of the catalytic domain of BaSrtC (residues Ser56–Lys198).
The backbone and heavy atom coordinates of structurally ordered amino acids have coordinate
precision of 0.42 ± 0.07 and 0.82 ± 0.05 Å, respectively. BaSrtCΔ55 adopts an eight-stranded β-
barrel fold that contains two short helices positioned on opposite sides of the protein.
Surprisingly, the protein dimerizes and contains an extensive, structurally disordered surface that
is positioned adjacent to the active site. The surface is formed by two loops (β2−β3 and β4–H1
loops) that surround the active site histidine, suggesting that they may play a key role in
associating BaSrtC with its lipid II substrate. BaSrtC anchors proteins bearing a noncanonical
LPNTA sorting signal. Modeling studies suggest that the enzyme recognizes this substrate using
a rigid binding pocket and reveals the presence of a conserved subsite for the signal. This first
structure of a class D member of the sortase superfamily unveils class-specific features that may
facilitate ongoing efforts to discover sortase inhibitors for the treatment of bacterial infections.
78
3.2 Introduction
Bacterial surface proteins play key roles in microbial physiology and pathogenesis.
Gram-positive bacteria display surface proteins using sortase enzymes, a large superfamily of
cysteine transpeptidases, which covalently join proteins bearing an appropriate C-terminal cell
wall sorting signal (CWSS) to strategically positioned amino groups located on the cell
surface.(1-7) Typically, bacteria encode multiple sortases that either attach proteins to the cross-
bridge peptide of the cell wall or assemble pili, long proteinaceous structures that extend from
the cell surface. Both processes occur through a related transpeptidation reaction that is best
characterized for the sortase A enzyme from Staphylococcus aureus (SaSrtA). SaSrtA anchors
proteins to the cell wall containing a CWSS that consists of an LPXTG motif (where X denotes
any amino acid), followed by a segment of hydrophobic amino acids, and a tail that is comprised
primarily of positively charged residues. The C-terminal charged tail presumably retards export,
positioning the protein for processing by the extracellular membrane-associated SaSrtA enzyme.
Catalysis occurs through a ping-pong mechanism that is initiated when the active site cysteine
residue nucleophilically attacks the backbone carbonyl carbon of the threonine residue within the
LPXTG motif, breaking the threonine–glycine peptide bond to create a sortase–protein complex
in which the components are linked via a thioacyl bond.(8, 9) SaSrtA then transfers the protein to
the cell wall precursor lipid II, when the amino group in this molecule nucleophilically attacks
and transpeptidation reactions that synthesize the cell wall then incorporate this product into the
bridge peptide. Instead of anchoring proteins to the cell wall, some sortases assemble pili.(2, 3)
These sortases operate through a similar mechanism; however, they link pilin subunits together
79
by joining a lysine amino nucleophile located in one protein to the CWSS of another. A
molecular-level understanding of sortase function could lead to new therapeutics for the
treatment of bacterial infections, as many clinically significant pathogens are attenuated in their
bacteria can be grouped into four distinct families, called class A–D enzymes.(7, 11, 12) Class A
enzymes are most closely related to SaSrtA. They appear to perform a housekeeping role in
different species of bacteria as members of this group have been shown to anchor a large number
of functionally distinct proteins to the cell wall that generally contain an LPXTG motif within
their CWSS. In contrast, class B–D enzymes appear to have more specialized functions as
representative members of these groups display a limited number of proteins that frequently
contain noncanonical sorting signals. Class B enzymes are present in Firmicutes and can have
distinct functions. Some members of this group attach heme receptors to the peptidoglycan,(13)
while others function as polymerases that assemble pili. Class C enzymes are also broadly
distributed in Gram-positive bacteria and function as pilin polymerases.(14, 15) Finally, class D
enzymes are present in many bacilli (Bacillus cereus, Bacillus anthracis, and Bacillus
thuringiensis). A single member of this family has been characterized, the class D enzyme from
B. anthracis, which has been shown to attach proteins to the cell wall that facilitate
sporulation.(16, 17) Atomic structures of representative class A–C enzymes have been
determined, revealing a conserved eight-stranded β-barrel fold that houses three essential active
site residues: His120, Cys184, and Arg197 (SaSrtA numbering).(1, 2, 7) The roles of these
residues in catalysis have been defined for SaSrtA and are likely conserved in other sortases:
Cys184 acts as a nucleophile that attacks the carbonyl atom in the scissile peptide bond of the
80
CWSS, Arg197 stabilizes the binding of the sorting signal substrate, and His120 may act as a
disease.(18) Like other bacterial species within the genus Bacillus, it forms dormant endospores
(spores) that are capable of surviving for long periods of time under harsh conditions. In humans,
anthrax caused by the inhalation of aerosolized B. anthracis spores has a high rate of mortality
and has led to their use as a bioterrorism agent.(19) Efficient sporulation of B. anthracis is
dependent upon the activity of SrtC (BaSrtC), a class D sortase that attaches the BasH and BasI
predivisional cells, while BasH is exclusively attached to the forespore presumably by BaSrtC
inherited from the mother cell before the polar division takes place. B. anthracis also encodes a
housekeeping class A enzyme (BaSrtA) that anchors a different set of proteins to the cell
wall.(20, 21) Interestingly, BaSrtC and BaSrtA specifically recognize very closely related sorting
signals.(16, 17) BaSrtA recognizes a canonical LPXTG-type sorting signal present in seven B.
anthracis proteins (three of these proteins are involved in collagen adhesion, while the functions
of the other proteins are not known). In contrast, BaSrtC anchors BasH and BasI that contain
LPNTA sorting signals. B. anthracis also encodes a third sortase, BaSrtB.(22) This class B
enzyme anchors the IsdC heme binding protein to the cell wall by recognizing its unique
NPQTN sorting signal. While the sortase enzymes in B. anthracis anchor different proteins to the
cell wall by recognizing distinct sorting signals, they are believed to attach these proteins to the
same chemical group, the m-diaminopimelic acid (m-DAP) side chain of the cross-bridge
peptide.(17, 23) In this paper, we describe the structure and dynamics of the B. anthracis BaSrtC
enzyme, which is the first reported structure of a class D enzyme. The structure provides insight
81
into the mechanism of protein anchoring in pathogenic B. anthracis, as well as the evolutionary
relationship between different types of enzymes within the sortase enzyme superfamily.
We studied the B. anthracis BaSrtC enzyme to gain insight into the molecular basis
through which it selectively anchors proteins to the cell wall during bacterial sporulation. BaSrtC
is 198 amino acids in length and contains a nonpolar N-terminal region (residues Met1–Tyr19)
that likely functions to embed the protein in the membrane. This is followed by a C-terminal
region (residues Lys20–Lys198) that has been shown to mediate the in vitro cleavage of an
LPNTA peptide that contains the amino acid sequence of the CWSS found in its BasI protein
substrate.(16) Inspection of the amino acid sequence of BaSrtC reveals that the C-terminal region
contains residues that are homologous to the catalytic domain of SaSrtA as well as a less
conserved polypeptide segment that connects the presumed catalytic domain to the N-terminal
nonpolar region (Figure 1A). Structural studies have revealed that many sortases can contain
ordered appendages that either precede or follow the amino acid sequence of the catalytic
domain. We therefore used NMR to study two polypeptides, BaSrtCΔ19 (residues Lys20–
Lys198), which contains the entire C-terminal region, and BaSrtCΔ55 (residues Ser56–Lys198),
which contains only amino acids within the presumed catalytic domain. The 1H–15N HSQC
spectra of these proteins reveal a similar set of well-dispersed backbone amide chemical shifts
indicating that these fragments adopt similar three-dimensional structures (Figure 1B). However,
the spectrum of BaSrtCΔ19 also contains an additional ∼30 cross-peaks that exhibit narrow line
widths and random coil chemical shifts. Surprisingly, this indicates that unlike sortases from
82
other classes, the linker preceding the catalytic domain in BaSrtC is unstructured in the isolated
protein.
Previous studies have shown that the intact C-terminal region of BaSrtC selectively
cleaves an LPNTA peptide derived from the CWSS of its BasI protein substrate.(16) To
ascertain whether only the conserved catalytic domain is required for this enzymatic activity, the
ability of BaSrtCΔ55 to cleave a peptide containing the LPNTA sequence was determined.
MALDI-TOF analysis of a reaction mixture containing BaSrtCΔ55 and the peptide reveals that
the isolated catalytic domain cleaves the peptide between the threonine and alanine residues
(Figure 2). Importantly, BaSrtCΔ55 exhibits specificity for the BasI sorting signal, as it is unable
to cleave LPATG and LPETG peptides derived from proteins that are anchored to the cell wall
by the BaSrtA sortase (data not shown and ref 16). Combined, these data indicate that in vitro the
catalytic domain within BaSrtC is folded and sufficient for hydrolytic activity.
and simulated annealing methods. A total of 2584 experimental restraints were employed,
including 2330 interproton distance, 196 dihedral angle, and 58 hydrogen bond restraints. Figure
exhibits good covalent geometry and has no NOE or dihedral angle violations greater than 0.5 Å
or 5°, respectively (structural and restraint statistics are listed in Table 1). The structure of
BaSrtCΔ55 is generally well defined by the NMR data; the backbone and heavy atom coordinates
mean-square deviations (rmsd's) from the average structure of 0.42 ± 0.07 and 0.82 ± 0.05 Å,
83
respectively. The structure of BaSrtCΔ55 also contains three disordered surface loops that
surround the active site (colored red and orange in Figure 3A), two of which (red) likely mediate
BaSrtCΔ55 adopts an eight-stranded β-barrel fold that contains two short helices
positioned on opposite sides of the protein (Figure 3B). The structure is initiated by strand β1
(Val67–Ile72), which after a short reverse turn interacts with strand β2 (Lys77–Tyr82) in an
antiparallel manner. A poorly ordered 11-amino acid loop then leads into strand β3 (Val94–
respectively. The active site His116 residue is located immediately after strand β4 and is
the opposite face of the protein. Strand β5 (Thr132–Glu136) then joins with β1 in an antiparallel
manner before the peptide reverses direction to form strand β6 (Thr141–Thr152). Strand β6
wraps around the enzyme toward the active site and contains a sharp point of curvature created
by a β-bulge at residues Gln147 and Lys148. A long loop containing helix H2 then leads into
strand β7 (Ile167–Thr172), which lies parallel with respect to strand β4. A large loop then
reverses the directionality of the chain before forming strand β8 (Arg185–Tyr197) that lies
antiparallel with respect to strands β6 and β7. Strand β8 is extensive as it contains a β-bulge at
residues Thr193 and Gly194 that allows it to form a continuous set of hydrogen bonding
interactions with strand β6 that together contribute residues to both faces of the protein structure.
84
3.3.3 Disordered Loops Positioned near the Active Site Might Mediate in vitro
Dimerization
On the basis of primary sequence homology, the active site in BaSrtCΔ55 is formed by
three spatially adjacent amino acids: His116, Cys173, and Arg185. These residues are conserved
in all sortases, and in BaSrtCΔ55, they are located at the end of a large cleft formed by strands β4,
β7, and β8. Cys173 is located at the end of strand β7, and in the prototypical SaSrtA enzyme, the
analogous residue mediates the nucleophilic attack on the threonine carbonyl carbon within the
sorting signal.(38, 39) Surrounding the thiol are the side chains of Arg185 (strand β8) and
His116 (strand β4), which in SaSrtA may stabilize binding of the sorting signal substrate and
sortases, BaSrtCΔ55 contains three disordered surface loops that are positioned immediately
adjacent to the active site (loops colored red and orange in Figure 3A). The loops are formed by
residues that immediately follow the active site cysteine so as to connect strand β7 to β8 (the
β7−β8 loop, residues Tyr173–Asp184), residues immediately following the active site histidine
that connect strand β4 to H1 (the β4–H1 loop, residues His116–Asp123), and residues
immediately adjacent to His116 that connect strands β2 and β3 (the β2−β3 loop, residues Trp83–
Gly93). The loops are structurally disordered because an insufficient number of interproton
distance restraints were identified in the NOESY spectra to define their position. For the β2−β3
and β4–H1 loops, few NOEs were observed because many of their residues exhibit resonance
line broadening. The broadening is most severe for residues Leu90–Lys92 (β2−β3 loop) and
His116–Phe121 (β4–H1 loop), which are completely absent from the NMR spectra. This
suggests that residues within the β2−β3 and β4–H1 loops experience fluctuations in their
85
magnetic environments that occur on a micro- to millisecond time scale, presumably because
they are flexible or because they reside within a surface that mediates protein aggregation.
To gain insight into the origin of structural disorder in the three active site loops, we
15 15
quantitatively probed N–H bond motions in BaSrtCΔ55 by measuring N spin–spin (T2), N
spin–lattice (T1), and {1H}–15N heteronuclear NOE relaxation parameters (Figure S1 of the
Supporting Information). In general, the relaxation data are compatible with the structure of
BaSrtCΔ55 as residues within regular secondary structural elements whose coordinates are well
defined in the ensemble of conformers have an average {1H}–15N NOE value of 0.79, indicating
that they undergo only small amplitude motions on the picosecond time scale. Interestingly, an
analysis of the relaxation data suggests that the structurally disordered β2−β3, β4–H1, and β7−β8
loops may participate in transient protein oligomerization. As the chemical shifts of many
residues within the disordered β2−β3 and β4–H1 loops are unassigned, the relaxation data do not
directly report on their dynamic status. However, several assigned surrounding residues within
the β2−β3 and β4–H1 loops, as well as in the adjacently positioned β7−β8 loop, exhibit
substantially shortened T2 values compared to those of other residues in the protein (Figure S1A
of the Supporting Information). Intriguingly, this indicates that their backbone amide atoms
experience fluctuating magnetic environments on the micro- to millisecond time scale that could
result from motions within the loops themselves and/or because the loops reside within or near a
A more detailed analysis of the NMR relaxation data indicates that active site loops likely
form a surface that mediates protein oligomerization. Initially, we attempted to model the
dynamics of BaSrtC using the Lipari–Szabo model-free formalism, which yields for each
backbone N–H bond vector a generalized order parameter (S2).(35) S2 ranges from 0 to 1 and
86
describes the rigidity of the amide bond on the picosecond time scale, with a value of 1
indicating that it is completely immobilized. Surprisingly, fitting of our data consistently yielded
many S2 values of precisely 1.0, which is suggestive of protein aggregation. This is illustrated in
Figure 4A, which shows a scatter plot of experimentally derived T1 and T2 values for each
residue overlaid with model-free predicted values of T1 and T2 for a range of S2 values [colored
lines indicate values of T1 and T2 expected for backbone amide nitrogens containing S2 values
ranging from 0.5 to 1 (see the key)]. The plot reveals that data for most residues are inconsistent
with all reasonable values of S2 because their T2 and T1 times are conjointly too short; this results
in the software attempting a “best fit” by assigning values of 1 to S2. We hypothesized that these
dependent, we remeasured the T1 and T2 times using a sample of BaSrtCΔ55 that was diluted to
0.125 mM. As shown in Figure 4B, the data for the dilute sample are in better agreement with
values predicted by the model-free formalism primarily because the residues now have longer T2
values. This is most likely because in the dilute sample a smaller fraction of the protein forms
sedimentation equilibrium ultracentrifugation experiments (Figure 4C). The resultant data are
best fit by a monomer–dimer equilibrium model that yields a dimer dissociation constant (KD) of
89 μM. Thus, at the protein concentrations used in our NMR studies, ∼90% of BaSrtCΔ55 likely
exists in its dimeric form. This explains our difficulty in fitting the relaxation data with the
model-free formalism and why analysis of the relaxation data yielded an overall molecular
correlation time of ∼16 ns, a value that is significantly longer than expected for a protein with
BaSrtCΔ55’s molecular mass. Taken together, the NMR and centrifugation data indicate that
87
residues within the β2−β3 and β4–H1 loops likely form a dimerization interface and that these
residues are broadened beyond detection in the NMR data because of monomer–dimer exchange
that occurs at a rate that is intermediate on the chemical shift time scale. This notion is
compatible with the positioning of many residues that exhibit anomalously short T2 times at high
protein concentrations as they are positioned proximal to the β2−β3 and β4–H1 loops (orange
points in Figure 4B and orange spheres in Figure 4D). Apparently, these residues also experience
them. In addition to forming a dimer interface, residues within the three disordered loops may
also exhibit elevated mobility as several residues within the β2−β3 and β7−β8 loops exhibit
depressed {1H}–15N heteronuclear NOEs indicating that they undergo faster picosecond motions.
The biological significance of the weak in vitro dimerization surface detected by NMR
and centrifugation is unclear. As BaSrtC is embedded in the membrane, it may indeed dimerize
on the cell surface because it undergoes limited two-dimensional diffusion. If this is the case,
then class D enzymes might be unique, as previously reported structures of class A–C sortases
have shown that they are monomeric(40-55) (it should be noted that biochemical studies have
shown that the S. aureus SaSrtA enzyme dimerizes with weak affinity, but this finding is
observed occurs only because BaSrtC is being studied in isolation. In this scenario, dimerization
via the large structurally disordered surface occurs because the appropriate binding partner that
normally interacts with this surface is missing. On the cell surface, the missing binding partner
could be, among others, protein factors involved in secretion (e.g., the SEC translocon), a
component of the cell wall, or BaSrtC’s intact lipid II and protein substrates. This is supported by
microscopy studies that have shown that the SrtA and SecA proteins in Streptococcus pyogenes
88
colocalize at the cross wall compartment where the cell wall is synthesized, and at polar sites
where surface protein anchoring also occurs.(57) It would also explain the in vitro catalytic
properties of BaSrtC as we and others have been able to demonstrate only that it can hydrolyze
its sorting signal (Figure 2 and ref 16). The full transpeptidation reaction using m-DAP as a
nucleophile has never been demonstrated biochemically. This is presumably because to be fully
active, BaSrtC needs to either associate with other factors on the cell surface or bind to its intact
lipid II and protein substrates. The notion that the disordered surface mediates substrate
recognition on the cell surface is supported by NMR and X-ray studies of SaSrtA and SaSrtB
that have implicated the β7−β8 loop in binding the cross-bridge peptide.(49, 53) Furthermore, it
is interesting to note that of all the sortase enzymes whose structures have been determined only
the sortases from B. anthracis (BaSrtA, BaSrtB, and BaSrtC) contain a disordered β7−β8 loop
and only these enzymes attach proteins to lipid II molecules with a m-DAP group.(50, 51) Thus,
on the cell surface, the large disordered loop regions identified by NMR may play a key role in
associating BaSrtC with protein factors and substrates that facilitate cell surface protein
anchoring.
Structures of representative class A–C sortases have been determined. Despite a limited
degree of sequence homology, BaSrtC is most structurally similar to class A enzymes based on a
DALI analysis. Three structures of class A enzymes have been reported: SaSrtA, BaSrtA, and
the SrtA enzyme from S. pyogenes (SpSrtA).(41, 48-50, 52) The sequence of BaSrtC is only 23–
24% identical with the sequences of these enzymes; however, the coordinates of its carbon atoms
that participate in regular secondary structures can be superimposed with rmsd's of 2.4, 1.9, and
89
2.0 Å, respectively. Notably, both class A and D sortases contain a short 310-helix located within
the β6−β7 loop (Figure 5, yellow). In SaSrtA, NMR studies have shown that this loop interacts
with the leucine-proline portion of the LPXTG sorting signal, suggesting that BaSrtC will also
recognize its distinct LPNTA sorting signal through a generally similar mechanism (vide
infra).(49) Interestingly, class A enzymes also exhibit structural heterogeneity in their N-termini,
because BaSrtA possesses a long N-terminal appendage that contacts the active site histidine.
This N-terminal appendage is disordered in the structures of BaSrtC, SaSrtA, and SpSrtA (Figure
5, green). Although the BaSrtC enzyme adopts a canonical sortase fold, its structure differs
markedly from class B and C enzymes. For example, in contrast to class B enzymes that possess
a large structured β6−β7 loop that contains an α-helix, the analogous loop in BaSrtC is
substantially shorter and lacks a similarly positioned helix (Figure 5, yellow).(42, 51, 53) Unlike
class B enzymes, BaSrtCΔ55 is also missing the N-terminal α-helices that precede the catalytic
domain (Figure 5, green). The structure of BaSrtCΔ55 also differs substantially from that of class
C enzymes as it is missing an N-terminal “lid” appendage that has been proposed to regulate
access to the active site in several members of this group.(44, 46, 54, 58)
To gain insight into how BaSrtCΔ55 recognizes its sorting signal, we modeled the
structure of its substrate complex by superimposing the backbone coordinates of BaSrtCΔ55 onto
the coordinates of our previously determined solution structure of the SaSrtA enzyme covalently
bound to an LPAT peptide analogue.(49, 59) Inspection of the model reveals that BaSrtC and
SaSrtA can bind to their respective sorting signals in a similar manner as there is minimal atomic
overlap between the peptide and BaSrtC atoms in the model of the BaSrtC–peptide complex
90
(Figure 6A). In the model, the sorting signal peptide adopts a kinked structure and rests in a
groove whose base is formed by residues in strands β4 and β7 and whose walls are formed by the
β6−β7, β7−β8, β3−β4, and β2–H2 surface loops. Because the experimentally determined
structure of the apo form of BaSrtC can readily accommodate the peptide, this suggests that the
enzyme may recognize its sorting signal substrate through a lock-and-key mechanism. This is in
marked contrast to SaSrtA, which undergoes major changes in the structure and dynamics of the
β6−β7 loop upon signal binding.(49, 60) This can be seen in Figure 6B. The left panel shows
apo-SaSrtA (red) aligned with our BaSrtC structure (green). While the active site residues of
apo-SaSrtA (pink) align well with the active site residues of BaSrtC (gray), the agreement is
significantly better than that of the holo-SaSrtA (SaSrtA–LPAT*) structure in Figure 6B [middle
panel (blue)]. Furthermore, the BaSrtC active site residues more closely resemble the position of
holo-SaSrtA than that of the closely related apo form of the class A sortase from B. anthracis,
apo-BaSrtA (Figure 6B, right panel, orange). This result further suggests that BaSrtC exists in an
acids within the signal as P and P′ if they precede and follow the scissile peptide bond,
respectively [e.g., the LPNTA signal-recognized BaSrtC is Leu (P4)-Pro (P3)-Asn (P2)-Thr (P1)-
Ala (P′1)]. Corresponding binding sites on the enzyme for these residues are termed S and S′,
respectively. The model clearly defines enzyme subsites that can accommodate the leucine (P4)
and proline (P3) residues in the LPNTA signal. The leucine residue is recognized by a small
pocket on the enzyme that is formed by the β6−β7 loop and residues in the underlying β-sheet.
Within the S4 subsite, the leucine is contacted by the side chains of Val166, Pro168, Val173, and
Val174 on the β6−β7 loop and Val198 on strand β8. The S3 subsite, which recognizes the
91
LPNTA signal’s proline residue, is formed by residues in the underlying β-sheet [Ala124 (β4)
and Ile185 (β7)] and Val110 within the adjacent β3−β4 loop. The structures of the class A
BaSrtA and SpSrtA enzymes also contain similarly shaped nonpolar subsites,(48, 50) which is
compatible with bioinformatics studies that have predicted that the majority of sorting signals
processed by class A and D enzymes contain proline and leucine amino acids at positions P3 and
P4, respectively.(11) In B. anthracis, the class A BaSrtA and class D BaSrtC enzymes anchor
proteins to the cell wall that contain closely related LPXTG and LPNTA sorting signals,
respectively. Our modeling studies suggest both enzymes recognize the conserved leucine-
proline element of these signals through structurally conserved subsites that do not require
substrate binding to form. How each distinguishes differences that occur at the P1′ position
cannot be predicted as the coordinate precision of amino acids that contact this residue is poorly
defined in the NMR ensembles of both enzymes. However, BaSrtC may in part distinguish the
unique signal present in its BasH and BasI substrates by interacting with the P2 amino acid that
is invariably an asparagine, as the bottom of BaSrtC’s S2 subsite contains Ser114, which can
presumably favorably interact with the asparagine amide group through hydrogen bonding
At present, more than 800 sortase genes have been identified in nearly 300 species of
bacteria. The functions of the vast majority of these enzymes are not known because most
bacteria contain multiple sortases with unknown substrate specificities. Work described here lays
the foundation for more detailed structure–function studies that will elucidate the molecular basis
of substrate specificity and could facilitate future efforts to predict sortase function and to
discover therapeutically useful sortase inhibitors. Structural data could aid inhibitor discovery
efforts in several ways.(61) First, atomic structures can be used to discover new chemical
92
scaffolds that are likely to bind and inhibit sortases. In this approach, an in silico screen of
compound libraries could be performed to identify small molecules that have the appropriate
physicochemical properties to efficiently interact with the enzyme active site. These molecules
serve as new leads for further development after their inhibitory properties have been confirmed
experimentally. Second, the structures can be used to optimize the binding affinity and
selectivity of established sortase inhibitors. In this approach, computational methods are used to
dock the lead inhibitor molecule to the structure of the enzyme. Analogues that are likely to have
increased potency are then identified in silico by modifying the bound inhibitor and calculating
an algorithm-dependent binding energy. The most promising of these analogues are then tested
biochemically to ascertain whether they have improved activity. Finally, if the structure of the
molecules with physicochemical features similar to those of the substrate. These molecules
presumably bind and inhibit the enzyme that can be tested experimentally. As several sortases
with known structures recognize sorting signals with distinct amino acid sequences, their unique
active site features in combination with the aforementioned computational methods may allow
class- and/or bacterial species-specific inhibitors to be developed. Such molecules are urgently
needed as infections caused by MRSA and other multidrug-resistant bacteria are a major health
concern.
BaSrtCΔ19 and BaSrtCΔ55 were amplified via polymerase chain reaction (PCR) from
genomic B. anthracis DNA (Sterne strain) with primers that placed an NdeI restriction site and a
93
BamHI restriction site on the 5′ and 3′ ends of the PCR product, respectively. Each PCR product
was digested with NdeI and BamHI restriction enzymes, as was empty vector pET15B (Qiagen).
Digested PCR products and pET15B plasmid were ligated together and transformed into
Escherichia coli XL-1 cells (Stratagene). Successful transformants were confirmed by DNA
sequencing. Plasmids where then transformed into E. coli BL21(DE3) (Stratagene) for
expression.
Protein for enzymatic assays was expressed in BL21(DE3) cells in standard Luria-Bertani
broth (LB), at 37 °C. Isotopically labeled protein for nuclear magnetic resonance (NMR) studies
15
was expressed in BL21(DE3) cells grown in M9 medium supplemented with NH4Cl and/or
centrifugation at 6000g and stored at −80 °C. Cells grown in M9 media where shifted to 22 °C
after induction, with induction allowed to progress overnight. Pellets were resuspended in
benzamidine (the final concentration of each was 2 mM) and allowed to lyse at room
temperature for 30 min. The lysate was cleared by centrifugation at 13000g for 30 min at 4 °C.
The soluble fraction was incubated with TALON His-affinity resin (Clontech). The resin was
washing with thrombin cleavage buffer [20 mM Tris (pH 8.0), 150 mM NaCl, and 2.5 mM
CaCl2]. Resin was then treated with thrombin at 37 °C for 2 h, releasing untagged BaSrtC
proteins. Protein was further purified by Sephacryl-100 gel filtration in either NMR buffer [20
94
3.4.2 NMR Spectroscopy and Structure Calculations
D2O]. In addition, PMSF and ethylenediaminetetraacetic acid (EDTA) were added to final
buffer and resolubilizing it with an equal volume of 99.999% D2O. NMR spectra were recorded
at 298 K on Bruker Avance 500, 600, and 800 MHz spectrometers equipped with triple-
resonance cryogenic probes. All NMR spectra were processed using NMRPipe(24) and analyzed
using CARA (version 1.8.4).(25) Chemical shift assignments (1H, 13C, and 15N) were obtained by
and ψ dihedral angle restraints. Additional φ and ψ dihedral angle restraints were determined by
measuring relative HNHAi and HNHAi–1 NOE intensities. Distance restraints were obtained
15 13
from three-dimensional N- and C-edited NOESY spectra. The collection of residual dipolar
couplings (RDCs) was attempted in a PEG/hexanol mixture; however, protein alignment was not
achieved. Initial NOE assignments were determined using ATNOS/CANDID(29, 30) and
back to CARA for validation and additional manual picking of NOE assignments. Hydrogen
bonds were identified initially on the basis of calculated structures and NOE patterns, verified
using deuterium exchange of backbone amides, and included in NIH-XPLOR calculations using
the HBDB algorithm.(32) A total of 200 final structures were calculated. Structures with no NOE
or dihedral angle violations greater than 0.5 Å or 5°, respectively, were selected with the 20
lowest-energy conformers presented. The BaSrtC–LPAT* complex was modeled by aligning the
95
SaSrtA–LPAT* structure [Protein Data Bank (PDB) entry 2KID] with the BaSrtC structure in
PyMOL.(33) The enzyme portion of the SaSrtA–LPAT* complex was removed, leaving behind
the substrate only. Manual adjustment of the remaining LPAT* substrate was done to remove
atom clashes. The NMR structure of the apoenzyme has been deposited as PDB entry 2LN7.
Bruker Avance 600 MHz spectrometer equipped with a triple-resonance cryogenic probe at an
initial protein concentration of 1 mM. Relaxation data and parameters were analyzed in
SPARKY,(34) which generated raw relaxation and NOE parameters. Relaxation parameters were
analyzed using programs kindly provided by Arthur G. Palmer III (Columbia University, New
York, NY) in an attempt to perform “ModelFree” analysis.(35) Using the data from a protein
However, we noticed that the R2R1_tm software that utilizes the R2/R1 ratio to calculate an
approximate correlation time (τc) resulted in an unusually high number for a protein of ∼16.5
additional protein concentration 0.125 mM. To compare our relaxation data to the ModelFree
formalism of Lipari and Szabo,(36, 37) we used eqs 7a and 7b from ref 36 to plot T1 versus T2
96
3.4.4 Analytical Ultracentrifugation and Enzyme Assays
HEPES (pH 6.0) and 100 mM NaCl with the addition of 5 mM TCEP. Absorption was
monitored at 280 nm at sample concentrations of 0.03, 0.015, and 0.0075 mM. Sedimentation
equilibrium profiles were measured at 40000 and 50000 rpm. The data were initially fit with a
nonlinear least-squares exponential fit for a single ideal species using the Beckman Origin-based
software (version 3.01). Analysis of the association behavior used the global analysis software
(the multifit option of the software mentioned above) to analyze four scans simultaneously,
corresponding to protein at a concentration of 0.03 mM at 40000 and 50000 rpm and a protein at
a concentration of 0.015 mM at 40000 and 50000 rpm. Partial specific volumes were calculated
Peptides were synthesized by NEO Biosciences and used without further purification.
BaSrtC (10 μM) was incubated with 200 μM peptide substrate in 5 mM CaCl2 and 20 mM
HEPES (pH 7.5) at room temperature. Samples (1 μL) were removed and plated with 0.5 μL of
2,5-dihydroxybenzoic acid (DHB) matrix (Acros Organics) dissolved in 50% ethanol and 0.1%
trifluoroacetic acid (TFA) and analyzed by MALDI-TOF immediately upon addition of enzyme,
and again after 24 h. Each reaction was performed in triplicate, and all samples were spotted and
analyzed in triplicate.
97
3.5 Figures
(A) Sequence alignment of B. anthracis (Ames strain) class A sortase (BaSrtA), class B sortase
(BaSrtB), and class D sortase (BaSrtC). The sequence alignment was performed by ClustalW.63
Conserved residues are colored orange, while identical residues are colored red. The predicted
98
transmembrane region (TM) is indicated by a cylinder labeled TM. Secondary structure features
of BaSrtC are indicated by cylinders or arrows above the sequence. (B) 1H–15N HSQC spectra of
BaSrtCΔ19 (left) and BaSrtCΔ55 (right). BaSrtCΔ19 contains approximately 30 more peaks, the
majority of which reside toward the center of the spectrum. The BaSrtCΔ55 HSQC spectrum has
99
Figure 3.2 MALDI-TOF demonstrating BaSrtCΔ55 can cleave the LPNTA-containing
peptide.
temperature. MALDI spectra of the peptide were taken immediately upon addition of the enzyme
100
Figure 3.3 Structures of B. anthracis class D sortase.
(A) Stereoview of the overlay of the ensemble of NMR structures of BaSrtCΔ55 (PDB entry
2LN7). The unassigned β2–β3 and β4–H1 loops are colored red. The assigned but poorly ordered
β7–β8 loop is colored orange. (B) Ribbon diagram with secondary structure elements labeled.
Loop structures are labeled with text. Active site residue side chains for Arg185, Cys173, and
101
Table 3.1 Structural Statistics for the Solution Structure of BaSrtCΔ55a
⟨SA⟩ ( SA )r
a
The notation of the NMR structures is as follows. ⟨SA⟩ represents the final 20 simulated
>5°.
102
c
The experimental dihedral angle restraints were as follows: 95 ϕ, 94 ψ, and39 χ1 angular
restraints.
d
Determined using PROCHECK. (62)
e
The coordinate precision is defined as the average atomic rmsd of the 20 individual SA
structures and their mean coordinates. These values are for residues 63−85, 94−113, 124−172,
and 182−198 of BaSrtCΔ55. Backbone atoms are N, Cα, and C′. Assignments were made for
103
Figure 3.4 Dynamics and ultracentrifugation analysis.
(A) Plot of experimentally determined T2 vs T1 data (black dots) for 1 mM BaSrtCΔ55 (assignable
backbone 15N resonances). Also plotted are lines that correspond to calculated T2 and T1 values
for various S2 order parameter values and correlation times (correlation times increase along each
line in a clockwise fashion). (B) Same as panel A, but T2 and T1 data were collected at a protein
concentration of 0.125 mM. Residues with T2 times of <0.07 s are colored orange. (C)
104
of 89 µM. (D) Structure of BaSrtC with unassigned residue backbone nitrogens colored red.
Residue backbone nitrogens with short T2 times identified in panel B are colored orange.
105
Figure 3.5 Comparison of representative structures from class A–D sortases.
B. anthracis class D enzymes are structurally similar to class A enzymes. BaSrtC lacks the long,
structured, β6–β7 loop (yellow) present in class B enzymes, but the appearance of a short 310-
helix in this region likely indicates the existence of a rigid substrate-binding pocket, similar to
BaSrtA. BaSrtC also lacks the structured N-terminal lid present in class C enzymes (green).
106
Figure 3.6 Substrate binding and active site architecture.
(A) Model of the CWSS peptide structure from the SaSrtA–LPAT* complex on the substrate
binding surface of BaSrtC. The peptide (green) is taken from the SaSrtA–LPAT* structure
without modification. Active site residues are indicated along with the potential Ser114 binding
site. The hydroxyl oxygen of Ser114 (Ser114 is colored gray, while its OH group is colored
white) comes into close contact with the side chain of position X. (B) Close-up overlay of the
BaSrtC structure (green) with apo-SaSrtA (red, left), holo-SaSrtA (blue, center), and apo-BaSrtA
(orange, right). Active site arginine, cysteine, and histidine residues are denoted. The active site
107
3.6 References
Insights into mechanism, substrate specificity, and inhibition Biopolymers 94, 385– 396.
2. Hendrickx, A. P., Budzik, J. M., Oh, S. Y., and Schneewind, O. (2011) Architects at the
bacterial surface: Sortases and the assembly of pili with isopeptide bonds Nat. Rev.
3. Mandlik, A., Swierczynski, A., Das, A., and Ton-That, H. (2008) Pili in Gram-positive
5. Marraffini, L. A., Dedent, A. C., and Schneewind, O. (2006) Sortases and the art of
anchoring proteins to the envelopes of Gram-positive bacteria Microbiol. Mol. Biol. Rev.
6. Scott, J. R. and Barnett, T. C. (2006) Surface proteins of Gram-positive bacteria and how
7. Spirig, T., Weiner, E. M., and Clubb, R. T. (2011) Sortase enzymes in Gram-positive
108
8. Frankel, B. A., Kruger, R. G., Robinson, D. E., Kelleher, N. L., and McCafferty, D. G.
(2005) Staphylococcus aureus sortase transpeptidase SrtA: Insight into the kinetic
mechanism and evidence for a reverse protonation catalytic mechanism Biochemistry 44,
11188– 11200.
9. Huang, X., Aulabaugh, A., Ding, W., Kapoor, B., Alksne, L., Tabei, K., and Ellestad, G.
11307– 11315.
10. Suree, N., Jung, M. E., and Clubb, R. T. (2007) Recent advances towards new anti-
infective agents that inhibit cell surface protein anchoring in Staphylococcus aureus and
11. Comfort, D. and Clubb, R. T. (2004) A comparative genome analysis identifies distinct
12. Dramsi, S., Trieu-Cuot, P., and Bierne, H. (2005) Sorting sortases: A nomenclature
proposal for the various sortases of Gram-positive bacteria Res. Microbiol. 156, 289–
297.
13. Mazmanian, S. K., Ton-That, H., Su, K., and Schneewind, O. (2002) An iron-regulated
14. Ton-That, H., Marraffini, L. A., and Schneewind, O. (2004) Protein sorting to the cell
wall envelope of Gram-positive bacteria Biochim. Biophys. Acta 1694, 269– 278.
109
15. Ton-That, H. and Schneewind, O. (2003) Assembly of pili on the surface of
16. Marraffini, L. A. and Schneewind, O. (2006) Targeting proteins to the cell wall of
17. Marraffini, L. A. and Schneewind, O. (2007) Sortase C-mediated anchoring of BasI to the
18. Koch, R. (1876) Die Ätiologie der Milzbrand-Krankheit, begründet auf die
19. Brookmeyer, R. and Blades, N. (2002) Prevention of inhalational anthrax in the U.S.
20. Gaspar, A. H., Marraffini, L. A., Glass, E. M., Debord, K. L., Ton-That, H., and
containing surface proteins to the cell wall envelope J. Bacteriol. 187, 4646– 4655.
21. Aucher, W., Davison, S., and Fouet, A. (2011) Characterization of the sortase repertoire
22. Maresso, A. W., Chapa, T. J., and Schneewind, O. (2006) Surface protein IsdC and
Sortase B are required for heme-iron scavenging of Bacillus anthracis J. Bacteriol. 188,
8145– 8152.
23. Budzik, J. M., Oh, S. Y., and Schneewind, O. (2008) Cell wall anchor structure of BcpA
110
24. Delaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J., and Bax, A. (1995)
25. Keller, R. (2004) The Computer Aided Resonance Assignment Tutorial, 1st ed.,
26. Palmer, A. G., III (1995) Protein NMR Spectroscopy, Academic Press, San Diego.
27. Teng, Q. (2005) Structural Biology: Practical NMR Applications, Springer, Berlin.
28. Shen, Y., Delaglio, F., Cornilescu, G., and Bax, A. (2009) TALOS+: A hybrid method
for predicting protein backbone torsion angles from NMR chemical shifts J. Biomol.
29. Herrmann, T., Guntert, P., and Wuthrich, K. (2002) Protein NMR structure determination
with automated NOE-identification in the NOESY spectra using the new software
30. Herrmann, T., Guntert, P., and Wuthrich, K. (2002) Protein NMR structure determination
with automated NOE assignment using the new software CANDID and the torsion angle
31. Schwieters, C. D., Kuszewski, J. J., Tjandra, N., and Clore, G. M. (2003) The Xplor-NIH
NMR molecular structure determination package J. Magn. Reson. 160, 65– 73.
111
32. Grishaev, A. and Bax, A. (2004) An empirical backbone-backbone hydrogen-bonding
potential in proteins and its applications to NMR structure refinement and validation J.
33. Delano, W. L. (2006) The PyMOL Molecular Graphics System, version 0.99, DeLano
34. Goddard, T. D. and Kneller, D. G. (2001) Sparky NMR Analysis Software, University of
35. Palmer, A. G., III (2001) NMR probes of molecular dynamics: Overview and comparison
with other techniques Annu. Rev. Biophys. Biomol. Struct. 30, 129– 155.
36. Lipari, G. and Szabo, A. (1982) Model-free approach to the interpretation of nuclear
37. Lipari, G. and Szabo, A. (1982) Model-free approach to the interpretation of nuclear
38. Liew, C. K., Smith, B. T., Pilpa, R., Suree, N., Ilangovan, U., Connolly, K. M., Jung, M.
E., and Clubb, R. T. (2004) Localization and mutagenesis of the sorting signal binding
site on sortase A from Staphylococcus aureus FEBS Lett. 571, 221– 226.
39. Ton-That, H., Liu, G., Mazmanian, S. K., Faull, K. F., and Schneewind, O. (1999)
112
proteins of Staphylococcus aureus at the LPXTG motif Proc. Natl. Acad. Sci. U.S.A. 96,
12424– 12429.
40. Cozzi, R., Malito, E., Nuccitelli, A., D’Onofrio, M., Martinelli, M., Ferlenghi, I., Grandi,
G., Telford, J. L., Maione, D., and Rinaudo, C. D. (2011) Structure analysis and site-
directed mutagenesis of defined key residues and motives for pilus-related sortase C1 in
41. Ilangovan, U., Ton-That, H., Iwahara, J., Schneewind, O., and Clubb, R. T. (2001)
Structure of sortase, the transpeptidase that anchors proteins to the cell wall of
Staphylococcus aureus Proc. Natl. Acad. Sci. U.S.A. 98, 6056– 6061.
42. Kang, H. J., Coulibaly, F., Proft, T., and Baker, E. N. (2011) Crystal structure of
Spy0129, a Streptococcus pyogenes class B sortase involved in pilus assembly PLoS One
6, e15969.
43. Lu, G., Qi, J., Gao, F., Yan, J., Tang, J., and Gao, G. F. (2011) A novel 'open-form'
structure of sortaseC from Streptococcus suis Proteins: Struct., Funct., Bioinf. 79, 2764–
2769.
44. Manzano, C., Contreras-Martel, C., El Mortaji, L., Izore, T., Fenel, D., Vernet, T.,
Schoehn, G., Di Guilmi, A. M., and Dessen, A. (2008) Sortase-mediated pilus fiber
45. Manzano, C., Izore, T., Job, V., Di Guilmi, A. M., and Dessen, A. (2009) Sortase activity
113
46. Neiers, F., Madhurantakam, C., Falker, S., Manzano, C., Dessen, A., Normark, S.,
pilus sortase C provide novel insights into catalysis and substrate specificity J. Mol. Biol.
47. Persson, K. (2011) Structure of the sortase AcSrtC-1 from Actinomyces oris Acta
48. Race, P. R., Bentley, M. L., Melvin, J. A., Crow, A., Hughes, R. K., Smith, W. D.,
Sessions, R. B., Kehoe, M. A., McCafferty, D. G., and Banfield, M. J. (2009) Crystal
49. Suree, N., Liew, C. K., Villareal, V. A., Thieu, W., Fadeev, E. A., Clemens, J. J., Jung,
M. E., and Clubb, R. T. (2009) The structure of the Staphylococcus aureus sortase-
substrate complex reveals how the universally conserved LPXTG sorting signal is
50. Weiner, E. M., Robson, S., Marohn, M., and Clubb, R. T. (2010) The Sortase A enzyme
that attaches proteins to the cell wall of Bacillus anthracis contains an unusual active site
51. Zhang, R., Wu, R., Joachimiak, G., Mazmanian, S. K., Missiakas, D. M., Gornicki, P.,
aureus and Bacillus anthracis reveal catalytic amino acid triad in the active site Structure
114
52. Zong, Y., Bice, T. W., Ton-That, H., Schneewind, O., and Narayana, S. V. (2004) Crystal
structures of Staphylococcus aureus sortase A and its substrate complex J. Biol. Chem.
53. Zong, Y., Mazmanian, S. K., Schneewind, O., and Narayana, S. V. (2004) The structure
54. Khare, B., Fu, Z. Q., Huang, I. H., Ton-That, H., and Narayana, S. V. (2011) The crystal
structure analysis of group B Streptococcus sortase C1: A model for the “lid” movement
55. Khare, B., Krishnan, V., Rajashankar, K. R., I-Hsiu, H., Xin, M., Ton-That, H., and
housekeeping and pilus-specific sortases: SrtA and SrtC1 PLoS One 6, e22995.
56. Lu, C., Zhu, J., Wang, Y., Umeda, A., Cowmeadow, R. B., Lai, E., Moreno, G. N.,
Person, M. D., and Zhang, Z. (2007) Staphylococcus aureus sortase A exists as a dimeric
57. Raz, A. and Fischetti, V. A. (2008) Sortase A localizes to distinct foci on the
Streptococcus pyogenes membrane Proc. Natl. Acad. Sci. U.S.A. 105, 18549– 18554.
58. Lu, G., Qi, J., Gao, F., Yan, J., Tang, J., and Gao, G. F. (2011) A novel “open-form”
115
59. Jung, M. E., Clemens, J. J., Suree, N., Liew, C. K., Pilpa, R., Campbell, D. O., and
analogue for covalent inhibition of sortases Bioorg. Med. Chem. Lett. 15, 5076– 5079.
60. Naik, M. T., Suree, N., Ilangovan, U., Liew, C. K., Thieu, W., Campbell, D. O., Clemens,
transpeptidase. Calcium promotes sorting signal binding by altering the mobility and
61. Taboureau, O., Baell, J. B., Fernández-Recio, J., and Villoutreix, B. O. (2012)
62. Laskowski, R. A., Rullmannn, J. A., MacArthur, M. W., Kaptein, R., and Thornton, J. M.
(1996) AQUA and PROCHECK-NMR: Programs for checking the quality of protein
63. Thompson, J. D., Gibson, T. J., and Higgins, D. G. (2002) Multiple sequence alignment
116
3.7 Supplementary Materials
117
Figure S3.1 Residue by residue T2, T1 and 15N{1H} Heteronuclear NOE data for
BaSrtCΔ55.
Panels C and D are T1 relaxation at 1 mM and 0.125 mM respectively. Panel contains the
15
N{1H} heteronuclear NOE data. Labeled bars indicate flexible loops. Absence of data
118
Chapter 4
Oxyanion Hole
119
The work described in this chapter has been reproduced with permission from:
Jacobitz, A. W., Wereszczynski, J., Yi, S. W., Amer, B. R., Huang, G. L., Nguyen, A. V.,
Sawaya, M. R., Jung, M. E., McCammon, J. A., Clubb, R. T. (2014). Structural and
http://doi.org/10.1074/jbc.M113.509273
Copyright 2014
I was the primary author of this work. I designed experiments, performed biochemical assays,
120
4.1 Overview
Sortase cysteine transpeptidases covalently attach proteins to the bacterial cell wall or
assemble fiber-like pili that promote bacterial adhesion. Members of this enzyme superfamily are
widely distributed in Gram-positive bacteria that frequently utilize multiple sortases to elaborate
their peptidoglycan. Sortases catalyze transpeptidation using a conserved active site His-Cys-Arg
triad that joins a sorting signal located at the C terminus of their protein substrate to an amino
nucleophile located on the cell surface. However, despite extensive study, the catalytic
mechanism and molecular basis of substrate recognition remains poorly understood. Here we
report the crystal structure of the Staphylococcus aureus sortase B enzyme in a covalent complex
with an analog of its NPQTN sorting signal substrate, revealing the structural basis through
which it displays the IsdC protein involved in heme-iron scavenging from human hemoglobin.
The results of computational modeling, molecular dynamics simulations, and targeted amino
acid mutagenesis indicate that the backbone amide of Glu224 and the side chain of Arg233 form
an oxyanion hole in sortase B that stabilizes high energy tetrahedral catalytic intermediates.
Surprisingly, a highly conserved threonine residue within the bound sorting signal substrate
facilitates construction of the oxyanion hole by stabilizing the position of the active site arginine
residue via hydrogen bonding. Molecular dynamics simulations and primary sequence
conservation suggest that the sorting signal-stabilized oxyanion hole is a universal feature of
121
4.2 Introduction
Surface proteins in bacteria play key roles in the infection process by promoting
microbial adhesion to host tissues, nutrient acquisition, host cell entry, and the suppression of the
immune response. In Gram-positive bacteria, virulence factors are displayed by sortase enzymes,
a superfamily of cysteine transpeptidases that join proteins bearing a cell wall sorting signal to
the cell wall or to other proteins to construct pili (1–6). Sortases have proven to be useful
molecular biology tools to site-specifically attach proteins to a variety of biomolecules and are
considered a potential drug target because they display virulence factors. The clinically
important pathogen, Staphylococcus aureus, displays surface proteins using two sortase
enzymes, sortase A (SrtA)(2) and sortase B (SrtB). srtA− strains of S. aureus are significantly
attenuated in virulence, whereas srtB− strains establish less persistent infections (7–9). SrtA
plays a “housekeeping” role in the cell, covalently mounting a variety of proteins to the cell wall,
whereas SrtB anchors the heme transporter IsdC, a key component of the iron-regulated surface
determinant system that captures heme-iron from hemoglobin (10–12). The mechanism of
catalysis is best understood for SrtA, the archetypal member of the sortase superfamily. Proteins
anchored by SrtA possess a C-terminal cell wall sorting signal that consists of an LPXTG motif
positively charged C-terminal tail (13). SrtA operates through a ping-pong mechanism that
begins when its active site cysteine residue nucleophilically attacks the carbonyl carbon of the
threonine in the LPXTG motif. This results in a tetrahedral intermediate that, after cleavage of
intermediate (14, 15). The protein is then transferred by SrtA to the cell wall precursor, lipid II,
when the amino group in this molecule nucleophilically attacks the thioacyl linkage creating a
122
second tetrahedral intermediate that collapses to form the covalently linked protein-lipid II
product (16–18). Transglycosylation and transpeptidation reactions that synthesize the cell wall
then incorporate this product into the peptidoglycan. SrtB anchors the IsdC protein to the cell
wall through a similar mechanism. However, unlike SrtA, SrtB recognizes a unique NPQTN
sorting signal, and it attaches IsdC to un-cross-linked peptidoglycan instead of heavily cross-
Sortase enzymes adopt an α/β sortase fold that contains three proximally positioned
active site residues: His130, Cys223, and Arg233 (SrtB numbering). Although the nucleophilic
role of the cysteine residue is well established, various catalytic functions have been proposed
for the histidine and arginine residues. His130 was originally thought to activate the cysteine by
forming a histidine-cysteine ion pair (21), but more recent data suggest that it instead functions
as a general acid/base (14, 22). Arg233 has been proposed to either stabilize substrate binding
(23–26), function as a general base (27), or directly stabilize the tetrahedral catalytic
intermediates (14, 22, 23). Other residues within the active sites of sortases have also been
remained poorly understood because the tetrahedral and acyl intermediates of catalysis are too
short lived to be characterized by either NMR or x-ray crystallography. Several sortase structures
have been determined in the absence of their substrates (28–32) or covalently bound to generic
sulfhydryl modifiers (33). However, only a single structure of a sortase enzyme covalently bound
to its sorting signal substrate has been reported (the NMR structure of SrtA bound to an LPAT
substrate analog) (1, 24). This structure revealed that the active site in this enzyme undergoes
substantial changes in its structure and dynamics that facilitate specific recognition of the sorting
123
signal, but it did not provide an atomic level view of the positioning of atoms within the active
site because their coordinates were not well defined in the NMR structure because of resonance
line broadening. Thus, the structural features utilized by sortases to stabilize key tetrahedral and
Here we report the 2.5 Å crystal structure of SrtB covalently bound to an analog of its
NPQTN sorting signal. The structure of the complex closely resembles the thioacyl intermediate
formed during catalysis, laying the groundwork for MD simulations to investigate the catalytic
mechanism. The results of these simulations and in vitro transpeptidation measurements suggest
that Arg233 and the backbone amide of Glu224 form an oxyanion hole that stabilizes high
within the sorting signal actively participates in constructing the oxyanion hole by hydrogen
bonding to the active site arginine residue. MD simulations of SrtA, as well as primary sequence
conservation, suggest that all sortases will use a similar substrate-stabilized mechanism to anchor
4.3 Results
4.3.1 Crystal Structure of the SrtB-NPQT* Complex and Computational Modeling of the
Thioacyl Intermediate
Sortase catalyzed transpeptidation reactions occur via covalent enzyme-substrate acyl and
tetrahedral intermediates that are too short-lived to be resolved by NMR or x-ray crystallography
(2, 14, 43). To overcome this problem, we synthesized a Cbz-NPQT* sorting signal analog,
where Cbz is a carbobenzyloxy protecting group, and T* is a threonine derivative that replaces
124
the sorting signal sequence recognized by the S. aureus SrtB sortase and forms a disulfide bond
via its T* moiety to the thiol of Cys223, generating a SrtB-NPQT* complex that structurally
mimics the thioacyl catalytic intermediate (Fig. 1a). A soluble version of SrtB lacking its 30-
amino acid N-terminal membrane anchor (SrtB, residues 31–244) was disulfide-bonded to the
NPQT* substrate analog, and a 2.5 Å crystal structure of the SrtB-NPQT* complex was
determined (Table 1). The data were refined to 2.5 Å resolution in accordance compatible with
CC1/2 statistics (36, 37), although the data extended only to 2.9 Å using more conventional
statistics such as Rmerge and I/σ. In the complex, SrtB adopts an α/β sortase fold containing eight
β-strands that are flanked by five α-helices (Fig. 1b). The positioning of the sorting signal
substrate is well defined, as evidenced by an Fo − Fc omit map of the complex (Fig. 2a). The
solvent-exposed Cbz group at the N terminus of the peptide is only partially modeled in the
structure of the complex because its electron density was not strong enough to fully define its
position. The signal adopts an L-shaped structure and is connected via the sulfhydryl of the T*
moiety to the active site cysteine, Cys223. It is nestled within a narrow groove whose base is
formed by residues within strands β4 and β5 and whose walls are formed by residues projecting
from loops connecting strands β6 to β7 (the β6/β7 loop), β7 to β8 (the β7/β8 loop), and β2 to β3
(the β2/β3 loop) (Fig. 1, b and c). To facilitate a discussion of the molecular basis of substrate
recognition, we henceforth utilize the nomenclature developed by Schechter and Berger (56),
where P and P′ refer to amino acids on the N-terminal and C-terminal sides of the scissile peptide
bond of the sorting signal, respectively. For the NPQTN sorting signal, the N-terminal N is P4, P
is P3, Q is P2, T is P1, and the C-terminal N, occurring after the scissile bond, is P1′. The base of
the NPQT binding pocket is defined by residues Asn92, Tyr128, Tyr181, Ile182, and Ser221,
with the walls defined by residues Leu96, Thr177, Cys223, Glu224, and Arg233 (Fig. 2, b and
125
c). The side chain of the disulfide linked T* residue is buried inside of a deep groove where its
methyl group contacts the side chains of Tyr128 and Ile182. This positions the hydroxyl oxygen
on the threonine residue to accept two hydrogen bonds from the side chain of Arg233 in the
active site: a 3.0 Å hydrogen bond to the ϵ-nitrogen atom and a 3.3 Å hydrogen bond to the η-
nitrogen atom in the guanidino group (Fig. 2, b and c). The side chain of the glutamine residue at
position P2 points out of the binding pocket where it is packed against the side chain of Leu96
and donates a hydrogen bond to the backbone carbonyl group of Glu224. Additional enzyme
interactions to the backbone of P2 Gln residue stabilize the positioning of the sorting signal. The
backbone carbonyl group is held in place by a hydrogen bond from the side chain amide nitrogen
of Asn92, and its backbone nitrogen donates a hydrogen bond to a sulfate ion, which in turn is
coordinated by the side chain of His93 and the backbone amide of Asn92. The P3 Pro residue
rests on top of Ile182 and forms a kink that causes the bound peptide to adopt an L-shaped
structure that positions the side chain of the P4 Asn to donate a hydrogen bond to the backbone
carbonyl of Thr177 within the β6/β7 loop. Substrate binding induces only small changes in the
structure of SrtB as the Cα coordinates of the SrtB-NPQT* complex can be superimposed with
the previously determined structure of the unmodified enzyme (28) with a root mean square
deviation of 0.44 Å.
model of the thioacyl complex was generated by replacing the disulfide link in the structure of
the SrtB-NPQT* complex with a thioacyl bond (Fig. 3). This required only small changes
including the removal of the methylene group in between the Cys223 thiol and the P1 Thr
residue of the sorting signal that decreased their separation by ∼1 Å. The coordinates of the
thioacyl intermediate were then energy-minimized while the restraints on the initial atom
126
positions were gradually removed. To verify that the final model contained the proper orientation
of the thioacyl bond, two initial models of the intermediate were energy-minimized, one in which
the carbonyl oxygen pointed toward Arg233 and a second model in which the thioacyl bond was
rotated by 180° (carbonyl oxygen pointing away from Arg233). After minimization, both starting
models converged to nearly identical structures (root mean square deviation = 0.71 Å for all Cα)
that closely resembled the SrtB-NPQT* complex. Importantly, in both refined models, the
thioacyl linkage adopts a similar conformation in which the P1 threonine residue in the signal is
poised to accept hydrogen bonds from the active site Arg233 residue and the backbone amide
from Glu224 (Fig. 3). This key sorting signal-enzyme interaction may hold the active site
Arg233 side chain in a catalytically competent conformation that also stabilizes higher energy
The catalytic importance of active site and sorting signal amino acids was investigated in
vitro using SNKDKVENPQTNAGT (sorting signal in bold) and GGGGG peptides that mimic
the sorting signal and secondary lipid II substrates, respectively. HPLC separation of the reaction
products indicates that SrtB used for crystallographic studies is fully functional in vitro (kcat and
Km values for SrtB of 1.0 × 10−4 s−1 and 1.8 mM, respectively) (Fig. 4a) (57). Based on sequence
conservation, residues His130, Cys223, and Arg233 in SrtB have been postulated to form a triad
that mediates catalysis. In addition, it has been proposed that the side chain of Asp225,
positioned near the active site, may also play a critical role in catalysis by stabilizing and
activating His130 (28). To investigate the relative importance of these residues, we purified four
single amino acid mutants of SrtB and assayed them for their ability to catalyze transpeptidation.
127
H130A, C223A, and R233A mutant enzymes had no detectable activity after 24 h, whereas
D225A exhibited nonspecific proteolytic activity (Fig. 4b). This suggests that unlike the
conserved active site residues (His130, Cys223, and Arg233), Asp225 is not required for early
steps in transpeptidation that involve the formation of the first thioacyl intermediate. The D225A
mutation may disrupt the active site architecture of SrtB, allowing recognition of various
sequences as primary substrates, but it does not appear to play a direct role in the catalytic
mechanism. Interestingly, similar promiscuous activity has been observed in several SrtA
mutants (57, 58), as well as the wild-type SrtB enzyme from Listeria monocytogenes (59).
complex was also tested using the HPLC assay. Sorting signal peptides containing alanine
substitutions at sites P1 (NPQAN), P3 (NAQTN), and P4 (APQTN) were unreactive (Fig. 4c),
compatible with extensive enzyme contacts to the side chains of these residues in the structure of
the complex. In contrast, peptides containing an alanine substitution at either site P2 (NPATN) or
P1′ (NPQTA) could be processed by the enzyme to yield transpeptidation products, although
they were less reactive than the native signal. The ability of SrtB to process the P2 mutant
peptide (NPATN) is compatible with the structure of the complex as the side chain of residue P2
projects out of the binding pocket toward the β7/β8 loop. The SrtA enzyme also processes
signals containing a range of amino acids at site P2 in its LPXTG sorting signal (60), suggesting
The structure of the SrtB-NPQT* complex and computational model of the thioacyl
intermediate reveal that the threonine side chain within the sorting signal likely plays a key role
in catalysis by stabilizing the positioning of Arg233 through hydrogen bonding (Figs. 2b and 3).
To determine the importance of this interaction in catalysis, we tested how efficiently NPQSN
128
and NPQVN peptides were used as transpeptidation substrates. These peptides are identical to
the native sorting signal peptide but contain threonine to serine and threonine to valine mutations
at the P1 position of the signal, respectively. The threonine to serine substitution preserves the
hydroxyl group that hydrogen bonds to Arg233 but removes the methyl group of threonine that
interacts with the side chains of Tyr128 and Ile182. In contrast, introduction of a valine
substitution eliminates the hydroxyl group but does not significantly alter the size or shape of the
P1 residue. Both peptides were unreactive, suggesting that each type of substrate-enzyme
4.3.3 Molecular Dynamics Simulations of SrtA and SrtB Reveal a Conserved Mechanism
through which the Substrate Stabilizes the Positioning of the Active Site Arginine
Residue
The atomic structures of only two sortases covalently bound to their substrates have been
determined: the NMR structure of SrtA bound to an LPAT* peptide (24) and the structure of the
SrtB complex reported here (Fig. 5). A comparison reveals a generally similar mode of binding
in which the signals adopt an L-shaped structure enabling extensive enzyme contacts to the side
chains of residues located at positions P4 and P3 (Leu-Pro and Asn-Pro in the SrtA and SrtB
substrates, respectively). Interestingly, although both bound sorting signals contain a conserved
threonine residue at the P1 position, the side chain of this amino acid is oriented differently in the
structures of the SrtA-LPAT* and SrtB-NPQT* complexes. As described above, in SrtB the side
chain of the P1 Thr residue faces “in” and interacts with the active site arginine, and the side
chain of the P2 Gln residue projects “out” toward the β7/β8 loop. In contrast, in the SrtA-LPAT*
structure the P1 Thr points out toward the solvent and the P2 Ala residue points in toward the
129
bottom of the binding pocket. This conformational difference occurs because the P1 and P2
residues in each signal have distinct backbone torsional angles; the P2 ϕ and ψ angles are 51 and
98° for SrtA-LPAT*, and −75 and 127° for SrtB-NPQT*, respectively, and the P1 T* pseudo ϕ
and ψ angles are −71 and −17° for SrtA-LPAT* and −119 and 25° for SrtB-NPQT*,
respectively. This key structural difference is not caused by a lack of coordinate precision in the
NMR structure of the SrtA-LPAT* complex because several NOEs define the positioning of the
It is possible that the sorting signals bound to SrtA and SrtB are flexible and thus capable
of undergoing motions in which the P1 Thr side chain moves into, and out of, the active site. To
investigate this issue, we performed MD simulations of both complexes using the method of
umbrella sampling with Hamiltonian replica exchange (54). The free energy profile for
transitions between the Thr-in (SrtB-like) and Thr-out (SrtA-like) states for the bound signals in
each complex was then calculated. MD calculations using the coordinates of the SrtA-peptide
complex indicate that the peptide can transition from the Thr-out conformation observed in the
NMR structure of the SrtA-LPAT* complex (Fig. 6a, middle panel) to a Thr-in orientation
observed in the structure of the SrtB-NPQT complex (Fig. 6a, left panel). To investigate the
energetics associated with this transition, we calculated the free energies of intermediate
structures on this pathway, which are represented by a two-dimensional coordinate system (Fig.
6b). The first coordinate (x axis) reports on the positioning of the P1 and P2 residues, and the
second reaction coordinate (y axis) reports on the positioning of the remainder of the sorting
signal relative to the enzyme. The former is defined as a collective coordinate that describes the
structure of residues P1 and P2 relative to the catalytic cysteine, and the latter is defined as the
radius of gyration of residues P3 and P4 with select atoms in the β sheet of the sortase molecule
130
(described further under “Experimental Procedures”). In the free energy profile for the covalent
SrtA-LPAT complex, there are three dominant energy wells all with minima within 0.4 kcal/mol
of one another. The region we refer to as well 1 corresponds to a sorting signal structure similar
to that observed in the SrtB-NPQT* complex in which the sorting signal contacts the arginine
(Thr-in) (Fig. 6a, left panel), whereas well 2 conformers resemble the NMR structure of the
SrtA-LPAT* complex in which the P1 threonine side chain points away from the active site
(Thr-out) (Fig. 6a, middle panel). Interestingly, this analysis reveals that a low energy pathway
exists between these two states (Fig. 6b), suggesting that the P1 Thr and P2 Ala residues can
alter their conformation within the active site of SrtA. This structural transition is documented in
supplemental Video S1, which shows select snapshots from the MD trajectory in which the P2
Thr transitions from the Thr-out to Thr-in state where it engages the active site arginine residue.
Interestingly, the Thr-out to Thr-in transition can also occur through a third low energy
intermediate (well 3) in which the hydrophobic residues P3 and P4 do not contact SrtA but are
instead exposed to solution. This entropically stabilized state is presumably not significantly
populated in vivo when the enzyme contacts larger protein substrates that contain a full cell wall
sorting signal.
narrow free energy minimum in which the threonine remains projected into the active site (Thr-
in) where it contacts Arg233 (Fig. 6, a, right panel, and c). This indicates that the Thr-out
conformer observed in SrtA is disfavored in SrtB. The larger size of the P2 residue in the SrtB
sorting signal could, in principle, cause this difference. In the SrtA-LPAT complex, the P2 Ala
residue adopts a positive ϕ angle, and the side chain projects toward the base of the binding
pocket, whereas the larger Gln P2 residue in the SrtB bound peptide adopts a negative ϕ angle
131
and rests on the surface of the enzyme. Because reduced steric stress enables amino acids with
smaller side chains to more readily adopt positive ϕ angles, it is possible that the smaller size of
the alanine side chain in the LPAT signal facilitates formation of the Thr-out state. To test this
hypothesis, a third free energy profile was computed for SrtB bound to NPAT, which changes its
P2 residue to alanine (Fig. 6d). This change expanded the range of conformers accessible to the
peptide bound to SrtB (Fig. 6, compare c and d), but it did not encourage sampling of the Thr-out
state observed in SrtA. This suggests that features of the SrtB-peptide complex other than the
identity of its P2 residue are important for dictating how the P1 Thr residue is positioned
(described below). In sum, our MD simulations indicate that sorting signals bound to both SrtA
and SrtB can adopt conformations in which the P1 Thr side chain is located within the active site
for stabilizing interactions with the active site arginine. In SrtB this is the predominant state of
the signal, whereas in SrtA, it is one of two possible low energy binding conformations.
4.4 Discussion
attach proteins to the bacterial cell wall or assemble pili (1–3). At present, over a thousand
sortase enzymes have been identified that, based on their primary sequences and functions, can
be partitioned into at least six distinct families (called class A to F enzymes) (1). The structure of
the SrtB-NPQT* complex reveals how class B sortases recognize their substrates. In bacterial
pathogens, class B enzymes typically anchor heme-binding proteins to the cell wall that enable
the bacterium to utilize host derived heme-iron as a nutrient (1, 10–12). Our results indicate that
SrtB recognizes its NPQTN sorting signal via a large groove adjacent to the active site formed by
residues in the β6/β7 loop and residues within strands β4 and β7. Structurally, class B enzymes
132
are distinguished from other members of the sortase superfamily by the presence of a large β6/β7
loop that contains a long α-helix. In the complex, this loop plays a major role in signal
recognition because it contacts the Asn (P4), Pro (P3), and Thr (P1) residues of the peptide,
which are highly conserved in sorting signals recognized by class B enzymes. These substrate-
enzyme contacts are important for catalysis because alanine substitutions at these sites in the
sorting signal disrupt transpeptidation (Fig. 4c). As in the prototypical SrtA enzyme, site P2 in
the SrtB sorting signal is tolerant to alanine substitution, which is consistent with the positioning
of the glutamine residue at this site, which rests on the surface of the enzyme. Although not
visualized in our structure, the P1′ residue following the scissile peptide bond is also tolerant to
alanine mutation (Fig. 4c), distinguishing SrtB from the prototypical class A SrtA enzyme, which
only recognizes glycine at the P1′ position (61). Given that the signal sequences and known
structures of class B enzymes are highly similar, it is likely that they all recognize their
substrates in a fashion similar to what is seen in the SrtB-NPQT* complex. Interestingly, our
results indicate that SrtA and SrtB enzymes recognize their cognate sorting signals in a generally
similar manner, despite the fact that the enzymes share only 26% sequence identity and that they
recognize distinct LPXTG and NPQTN sorting signals, respectively. A comparison of the SrtB-
NPQT* and previously reported NMR structure of the SrtA-LPAT* complex reveals that the
bound sorting signals both adopt an L-shaped conformation in which the proline residue at
position P3 redirects the polypeptide to position the side chain of the P4 residue so that it
contacts the β6/β7 loop. This general mode of binding is likely a conserved feature of substrate
these enzymes shows that over 90% of them contain a proline at P3 (60).
133
Computational modeling of the thioacyl enzyme-substrate reaction intermediate suggests
that SrtB facilitates catalysis by forming a substrate-stabilized oxyanion hole. In both the
experimental and energy-minimized model of this reaction intermediate, the conserved active
site arginine (Arg233) is held in place near the scissile bond by donating a hydrogen bond from
its ϵ-nitrogen to the hydroxyl group of the threonine located at the P1 position in the sorting
signal (Figs. 2b and 3). In the model of the thioacyl complex, this interaction positions the
arginine so that its guanidino group and the backbone amide of Glu224 donate hydrogen bonds
to the oxygen atom in the thioacyl bond. During catalysis, two oxyanionic transition states form;
the first precedes the generation of the thioacyl intermediate emulated by the SrtB-NPQT*
structure, and the second occurs after nucleophilic attack of the thioacyl bond by the amino
group present in lipid II. It seems likely that Arg233 and Glu224 form an oxyanion hole that
stabilizes these high energy reaction intermediates because only small changes in the positioning
of the oxygen atom within the thioacyl bond are expected to occur when the carbonyl carbon
transitions from its planar sp2 configuration to its tetrahedral sp3 state. This is distinct from the
oxyanion hole discovered in penicillin binding proteins that perform a similar transpeptidation
reaction but use two backbone amides to stabilize the tetrahedral intermediate (62). The
threonine residue in the sorting signal appears to play a significant role in stabilizing this
oxyanion hole because even conservative mutations to either serine or valine disrupt
transpeptidation (Fig. 4c). Its catalytic role is also substantiated by the high degree of sequence
conservation at site P1 in all known SrtB substrates. The putative oxyanion hole reported here is
compatible with results obtained by McCafferty and co-workers (22, 23), who demonstrated that
the analogous arginine in the SrtA enzyme is intolerant to mutation, except when substituted
134
MD simulations indicate that the SrtA enzyme can also form a substrate stabilized
oxyanion hole that could facilitate catalysis. In the structures of the SrtB-NPQT* and SrtA-
LPAT* complexes, the bound sorting signals adopt similar L-shaped conformations. However,
the positioning of the P1 Thr residue in each sorting signal differs substantially; in SrtB-NPQT*,
the threonine side chain interacts with the active site arginine residue (Thr-in conformation)
(described above), whereas in the SrtA complex, the analogous threonine residue projects away
from the active site (Thr-out conformation). Because the threonine in the SrtA complex is not
positioned to form stabilizing interactions with the active site arginine residue, we wondered
whether the Thr-in conformer observed in the SrtB complex could also be sampled by the bound
SrtA sorting signal to stabilize its tetrahedral reaction intermediates. To investigate this issue,
MD simulations of the SrtA thioacyl complex were performed, revealing that the P1 threonine
residue can transition from the Thr-out to the Thr-in state presumably needed to construct the
oxyanion hole (supplemental Video S1 and Fig. 6). Conformations of the SrtA-LPAT complex in
which the threonine side chain of the sorting signal forms interactions with the active site
arginine were obtained without major rearrangement of the structure of the enzyme.
Interestingly, a previous MD study based on the SrtA-LPAT* NMR structure reports that the
active site arginine residue functions only to position the sorting signal substrate by hydrogen
bonding to its backbone carbonyl atoms at positions P2 and P4 (25). Our work is compatible
with this conclusion but suggests that this stabilizing interaction will only occur when the P1 Thr
samples the out position that is presumably not catalytically active. Thus, based on primary
sequence conservation and the demonstrated importance of the P1 Thr in catalysis in both SrtA
and SrtB, we conclude that the Thr-in conformation observed in the SrtB structure, and
135
accessible to SrtA, represents a catalytically competent form of the peptide that is essential for
The active site of SrtB appears to be more conformationally restrictive than SrtA because
MD simulations of the SrtB-NPQT thioacyl complex reveal that only the Thr-in conformer is
accessible to the bound peptide. This is compatible with structural and NMR relaxation data,
which have shown that the active site of SrtA contains a flexible β6/β7 loop that becomes
ordered upon substrate binding (24, 63), whereas SrtB contains a preformed, rigid binding pocket
for its sorting signal substrate. Interestingly, inspection of the MD data suggests that SrtA can
form unique contacts to the sorting signal that may stabilize the Thr-out conformer. In the NMR
structure of the SrtA-LPAT* structure, the conserved active site His101 residue is positioned to
form a 3.1 Å hydrogen bond from its ϵ-nitrogen to the P1 Thr hydroxyl. In contrast, although the
SrtB active site histidine is about the same distance from the active site cysteine as the analogous
residues in SrtA (∼5.2 Å from His δ-N to Cys S), this potential stabilizing interaction for the
Thr-out conformer is obstructed by the side chain of Leu96, which is inserted between His130
and Cys223 in SrtB. Thus, the more restrictive active site of SrtB and the lack of stabilizing
interactions may prevent the signal bound to SrtB from adopting the catalytically nonproductive
Thr-out conformer. It is unlikely that the ability of the SrtA peptide to sample the less reactive
Thr-out state impacts the kinetics of transpeptidation because the half-life of the long-lived
thioacyl intermediate presumably far exceeds the time needed for the threonine residue to
It seems likely that nearly all members of the sortase superfamily will employ a substrate
stabilized oxyanion hole to anchor proteins to the cell wall or to assemble pili. Based on
structural and mutagenesis data, the transpeptidation reaction will be initiated when the sorting
136
signal of the partially secreted protein substrate binds to the groove on sortase formed by
residues in the β6/β7 loop and residues within strands β4 and β7. Most sorting signals can be
expected to adopt an L-shaped structure when bound to the enzyme because they contain a
proline residue at position P3 (∼90% of predicted sorting signals) (60). For catalysis to proceed,
the sortase must contain a properly charged active site in which the cysteine and histidine
residues are in their thiolate and imidazolium ionization states, respectively. In isolation, only a
small fraction of the enzyme may be properly ionized based on pKa measurements of the SrtA
enzymes from S. aureus and Bacillus anthracis (14, 29, 64). If properly ionized, the cysteine
thiolate attacks the carbonyl carbon of the P1 residue, forming the first tetrahedral intermediate.
Our results suggest that the oxyanion in this intermediate is stabilized by hydrogen bonding from
the active site arginine residue and a backbone amide group located in the β7/β8 loop (in SrtB
Arg233 and the backbone amide of Glu224) (Fig. 7). The threonine residue within the sorting
signal, at position P1, plays a key role in constructing the oxyanion hole by stabilizing the
positioning of the arginine side chain via hydrogen bonding. This oxyanion hole is presumably
used by most sortase enzymes to facilitate catalysis because they all contain conserved active site
arginine residues, and ∼95% of their predicted sorting signal substrates contain a threonine at the
P1 position (60). Breakage of the scissile bond is then facilitated by protonation of the amide
group by His130, resulting in a semistable thioacyl intermediate. It remains unclear where the
amino nucleophile on lipid II enters the active site in SrtB. A crystal structure of SrtB
noncovalently bound to a triglycine peptide, meant to mimic the lipid II substrate, localized the
binding site for the peptide to the β7/β8 loop (33). However, the specificity of this interaction is
suspect because the peptide is expected to bind weakly (Km for GGGGG binding is 140 μM (43))
and because the complex was not co-crystallized (the peptide was soaked into the crystal). It is
137
also important to note that in this and all crystal structures of SrtB solved to date, side chains
from residues Asp225–Tyr227 in the β7/β8 loop are involved in crystal lattice contacts to
symmetry related molecules in the crystal. Although it is possible that these contacts simply
reinforce the existing, predominant conformation of this loop, it is also possible that they have
captured one of many possible conformations that could be involved in mediating substrate
access to the active site. Alternatively, as originally proposed by Joachimiak and co-workers
(28), the pentaglycine peptide in lipid II may enter the active site via a groove located between
the β7/β8 and β2/β3 loops. This is compatible with NMR chemical shift perturbation studies of
SrtA (24), high resolution crystal structures of other sortase enzymes, which also contain a
similarly positioned groove (30, 32), and the presence of the highly conserved histidine residue
in this groove that has been proposed to function as a general base that deprotonates lipid II. Our
model of the thioacyl intermediate does not rule out either of these entry points. However, it is
most compatible with lipid II entering via the groove between the β7/β8 and β2/β3 loops because
from this direction, attack of the carbonyl carbon by the amino nucleophile will generate a
hole formed by Arg233. The transpeptidation reaction would then be completed by the collapse
of the second tetrahedral intermediate into the final, covalently linked, protein-lipid II product.
to quantitatively investigate the role of the oxyanion hole in catalysis. Beyond providing
fundamental insight into the process of protein display and pilin assembly in bacteria, the new
mechanistic insights reported in this paper could guide the rational design of therapeutically
useful transition state analog inhibitors of sortases and facilitate protein engineering efforts to
138
4.5 Experimental Procedures
DNA encoding SrtB (residues 31–244) was amplified by PCR from S. aureus genomic
DNA, cloned into a pE-SUMO vector (LifeSensors) and transformed into Escherichia coli
Rosetta (DE3) pLysS cells (Novagen). Protein expression was induced by addition of 1 mM
purified by affinity purification using HisPur cobalt resin (Thermo) per the manufacturer's
instructions. The His6-SUMO tag was then cleaved by incubating the protein overnight at 4 °C
with recombinant ULP1 protease and removed by reapplying the protein mixture to the HisPur
carbobenzyloxy protecting group) was synthesized as in Ref. 34 and added to purified SrtB in
modification buffer (10 mM Tris-HCl, pH 7.0, 20 mM NaCl, 1 M L-proline) at a ratio of 10:1 for
a final concentration of 1 mM Cbz-NPQT* to 100 μM SrtB. The reaction was first reduced with
1 mM DTT for 4 h, then oxidized by addition of 10 mM CuCl2, and allowed to rock gently at
room temperature for 7 days. Production of stable complex was confirmed by MALDI mass
spectrometry.
Crystals of the SrtB-NPQT* complex were produced from a stock of 150 μM SrtB-
NPQT* in 10 mM Tris-HCl, pH 7.0, 20 mM NaCl. Crystals were grown using the hanging drop,
vapor diffusion method in 2.8 M ammonium sulfate, 70 mM sodium citrate, pH 5.0. Data were
collected on Beamline 24-ID-C at 100 K at the Advanced Photon Source (λ = 0.964 Å). Three
data sets were scaled, integrated, and merged using XDS and XSCALE (35). Using conventional
criteria, the resolution boundary for the data set might have been drawn at 2.9 Å given that I/σ in
139
this shell (2.98–2.90 Å) is 2.0, and Rmerge is 72%, with a completeness of 95% and a multiplicity
of 5.2. However, recent studies from Karplus and Diederichs (36) have indicated that the CC1/2
statistic has superior properties as an indicator of data precision compared with Rmerge. Moreover,
in their study, and in the following work (37), the authors show that high resolution data
typically discarded because of high Rmerge values (i.e., over the conventionally acceptable
threshold value of ∼60–80%), and low I/σ values (i.e., under the conventionally acceptable
threshold value of 2) actually contain information that can improve the quality of the model if
used in refinement. Given these results, we thought our model would improve if we used the
more generous resolution cutoff (2.5 Å) indicated by the CC1/2 statistic (50.4% in the 2.5 Å
shell), rather than a conventionally accepted limit (2.9 Å) indicated by the Rmerge and I/σ
statistics. Thus, data extending to 2.5 Å resolution were used for the refinement process,
resolution.
Phases were determined by molecular replacement using the unmodified SrtB structure
(28) (Protein Data Bank code 1NG5) as a search model in the program Phaser (38). The NPQT*
modifier was modeled into positive density using COOT (39, 40), and the model was prepared
through successive iterations of manual adjustment in COOT and refinement in BUSTER (41,
42).
Active site mutants were produced using the QuikChange site-directed mutagenesis kit
expressed and purified as described for the wild-type protein. In vitro transpeptidation reactions
140
were performed based on the method developed by Kruger et al. (43). 100 μM SrtB (wild-type or
mutant) was incubated with 2 mM GGGGG and 200 μM peptide substrate in 100 μl of assay
buffer (300 mM Tris-HCl and 150 mM NaCl) at 37 °C for 24 h. The reactions were quenched by
adding 50 μl of 1 M HCl and injected onto a Waters XSelect HSS C18 reversed phase HPLC
column. Peptides were eluted by applying a gradient from 3 to 23% acetonitrile (in 0.1%
trifluoroacetic acid) over 25 min at a flow rate of 1 ml/min. Elution of the peptides was
monitored by absorbance at 215 nm. Peak fractions were collected, and their identities were
Molecular dynamics simulations were performed with NAMD (44), using the
AMBER99SB-ILDN force field (45), a 2-fs time step, and the SHAKE algorithm to constrain all
hydrogen containing bonds (46). Nonbonded interactions were truncated at 10 Å, with the use of
a smoothing function beginning at 9 Å, and long range electrostatics were handled with the
particle mesh Ewald method using a maximum grid spacing of 1 Å and a cubic B spline (47).
Parameters for the Cys-Thr linkage were generated with GAFF (48, 49), with the charges derived
from a RESP fit (48). Constant temperature was maintained through the use of Langevin
dynamics with a damping coefficient of 2 ps−1, whereas the barostat was controlled through a
Nosè-Hoover method with a target pressure of 1 atm, a piston period of 100 fs, and a damping
Models of the thioacyl intermediate were originally constructed from the SrtB-NPQT*
structure by replacing the disulfide bond with a thioester in PyMOL (52). The models were
solvated in a periodic water box with a solvent distance of 10 Å and parameterized in tLeap (53).
141
Models were then energy-minimized and equilibrated in NAMD (44) by slowly removing
restraints from the initial atom positions over 1 ns with 2-fs steps. For simulations of SrtA, the
exchange umbrella sampling calculations (54). For the first dimension (the x coordinate in Fig. 6,
b–d), a vector was defined based on the difference in positions between residues in the SrtA
structure 2KID and the SrtB structure presented here for the heavy atoms in the backbones of
residues P1, P2, the catalytic cysteine, and the three residues upstream of it in the sortase
molecule, along with the heavy atoms in the backbone of residues P1 and the catalytic cysteine.
The second coordinate (the y axis in Fig. 6, b–d) was defined as the radius of gyration for the C
atoms in residues 98–100 and 142–144 in SrtA (or residues 171–173 and 235–237 in SrtB) along
with the heavy atoms of residues P4 and P3. Restraints for umbrella sampling were evenly
spaced every 10 Å from −100 to 100 Å in the first coordinate and every 0.5 Å from 5 to 14 Å in
the second coordinate. This created a total of 399 simulation “windows,” each of which were
simulated for 10 ns. Positions were exchanged between adjacent windows every 1 ps based upon
a Metropolis criteria with a temperature of 300 K. The weighted histogram analysis method was
used for computing the potential of mean force based upon the umbrella sampling calculations
(55). Analysis of subsamples from these simulations indicate that the overall free energy profiles
require on the order of 5 ns to equilibrate; thus the first 5 ns of each window is discarded in the
142
4.6 Figures
(a) a comparison of the chemical bonds that join the peptide substrate to the SrtB enzyme in the
SrtB-NPQT* complex (left panel) and the SrtB-NPQT thioacyl catalytic intermediate (right
panel). Atoms from SrtB are colored red. (b) ribbon diagram of the SrtB-NPQT* complex.
Green, β2/ β3 loop; orange, β6/β7 loop; red, β7/β8 loop. Active site residues and substrate
analog are shown as sticks. (c) surface representation of SrtB in the complex utilizing the same
color scheme as in (b). Active site residue Arg233 is highlighted in blue, and Cys223 is in
yellow.
143
Table 4.1 Crystallographic data collection and refinement statistics
Parameters SrtB-NPQT*
Data Collection
Space Group P21
Cell Dimensions
a, b, c (Å) 102.6, 59.12, 72.49
Resolution (Å) 72.49-2.49 (2.58-
2.49)
Rmerge 26.6 (212.4)
I/σ(I) 3.8 (0.7)
Completeness (%) 92.2 (94.2)
Multiplicity 5.1 (5.0)
CC1/2 98.6 (50.4)a
CC* 99.6 (81.8) a
Refinement
Resolution (Å) 72.49-2.49
No. Reflections 28147
Rwork/Rfree 21.5/27.1
CCwork 94.4 (76.4) a
CCfree 87.7 (75.5) a
No. atoms
Protein 7248
Ligand/ion 176
Water 27
B factors
Mean 72.5
Wilson 56.2
Protein 72.3
Ligand/ion 87.6
Water 29.9
RMSD
Bond lengths (Å) 0.010
Bond angles (˚) 1.26
Values in parentheses are for highest-resolution shell.
a
Values reported as a percentage (%)
144
Figure 4.2 Structure and interactions of the NPQT* modifier in complex with SrtB.
(a) Fo − Fc map contoured at 3 σ. Electron density (gray mesh) was generated by removing the
NPQT* peptide from the final model and repeating refinement. The map shown is an average of
145
the density from all four chains in the asymmetric unit. The peptide extends from the active site
cysteine. (b) stereo view of the NPQT* peptide (gray sticks) and interacting residues (blue
sticks). Hydrogen bonds are indicated by dashed green lines. (c) diagram of the interactions
between SrtB (blue) and the NPQT* peptide (gray). Hydrogen bonds are indicated by dashed
green lines. SrtB residues that make only hydrophobic contacts are depicted as blue circles
146
Figure 4.3 Expanded view of the active site in the energy-minimized model of the SrtB-
Interactions between SrtB and the threonine residue in the sorting signal are shown. The thioacyl
carbonyl oxygen atom is positioned to accept hydrogen bonds from the η-nitrogen atom of
Arg233 and the backbone nitrogen atom of Glu224. The side chain of Arg233 is held in position
by a hydrogen bond between its ϵ-nitrogen atom and the hydroxyl group on the Thr residue
147
Figure 4.4 Transpeptidation activity of wild-type and mutant SrtB.
(a) representative HPLC chromatograms showing the reaction products that are produced when
SrtB was incubated with SNKDKVENPQTNAGT (sorting signal in bold type) and GGGGG
peptides that mimic its sorting signal and secondary lipid II substrates, respectively. Reactions
performed in the presence (left panel) and absence (right panel) of SrtB are shown. Only when
SrtB is present (left panel) is the appropriate transpeptidation peptide product produced
mutant was incubated with 200 μM peptide containing an NPQTN sorting signal and
pentaglycine and monitored by HPLC as described above. The dark shaded bars indicate the
amount of full-length peptide remaining after reaction with the enzyme. The lightly shaded bar
indicates the amount of transpeptidation product that was formed. An asterisk indicates that no
transpeptidation product could be detected after 24 h. The error bars represent the standard
148
deviation of three reactions. (c) transpeptidation activity of sorting signal amino acid mutants.
Sorting signal peptides containing select alanine substitutions were assayed for their ability to be
utilized by SrtB as a substrate and monitored by HPLC as described above. The amount of
transpeptidation product formed for each mutant peptide is expressed as a percentage of the
amount formed from reaction of SrtB with the native NPQTN sorting signal.
149
Figure 4.5 Alignment of SrtB-NPQT* and SrtA-LPAT* (24).
SrtB is shown as blue ribbons, and SrtA is shown as green ribbons.
150
Figure 4.6 Structures and free energy profiles of the SrtA and SrtB thioacyl complexes.
(a) selected structures obtained from MD simulations of the SrtA and SrtB thioacyl complexes.
The left and middle panels show structures of the SrtA thioacyl complex in which the threonine
side chain in the sorting signal either interacts with the active site arginine (left panel, SrtA-
LPAT Thr-“In”) or projects away from the active site (middle panel, SrtA-LPAT Thr-“Out”).
The right panel shows the structure of the SrtB-NPQT thioacyl complex in which the threonine
side chain interacts with the arginine residue (Thr-in). The structures are displayed with the SrtA
and SrtB surfaces colored green and blue, respectively. Surface representations of labeled active
site residues were calculated independently of the remaining protein surface, which allows for
the visualization of SrtB His130 (cyan) behind Leu96 (dark gray) in the right panel, even though
this residue would not normally be considered solvent-accessible in this conformation. (b–d) free
151
energies were calculated from positions sampled during a Hamiltonian replica exchange
simulation for the SrtA-LPAT (b), SrtB-NPQT (c), and SrtB-NPAT (d) thioacyl complexes. The
x axis records the position of P1 and P2 residues in the bound sorting signal relative to the active
site cysteine residue. The y axis describes the positioning of residues P3 and P4 in the sorting
152
Figure 4.7 The SrtB transpeptidation mechanism showing how the sorting signal may
(a) the SrtB active site with His130 in its imidazolium form and Cys223 in its thiolate form. (b)
the NPQTN substrate binds with its P1 Thr residue in the in position within the active site. The
Cys223 thiolate performs a nucleophilic attack on the P1 Thr carbonyl carbon. (c) the first
tetrahedral intermediate is formed. The P1 Thr residue interacts with Arg233 to construct a
substrate-stabilized oxyanion hole in which the side chain of Arg233 and the backbone amide of
Glu224 hydrogen bond to the oxyanion. The His130 imidazolium group donates a proton to the
leaving group to complete breakage of scissile bond. (d) the P1 Thr residue maintains hydrogen
bonds with Arg233 to stabilize its interaction with the thioacyl intermediate. (e) the incoming
GGGGG peptide from lipid II acts as a nucleophile that attacks the carbonyl carbon of the
thioacyl bond. (f) the second tetrahedral intermediate is formed and is again stabilized by the
NPQTGGGGG transpeptidation product and returning the enzyme to its active form (a).
153
4.7 References
1. Spirig, T., Weiner, E. M., and Clubb, R. T. (2011) Sortase enzymes in Gram-positive
insights into mechanism, substrate specificity, and inhibition. Biopolymers 94, 385–396
and Mechanisms of Their Targeting to the Cell Wall Envelope. Microbiol. Mol. Biol.
5. Popp, M. W.-L., and Ploegh, H. L. (2011) Making and Breaking Peptide Bonds: Protein
6. Tsukiji, S., and Nagamune, T. (2009) Sortase-Mediated Ligation: A Gift from Gram-
7. Mazmanian, S. K., Liu, G., Jensen, E. R., Lenoy, E., and Schneewind, O. (2000)
Staphylococcus aureus sortase mutants defective in the display of surface proteins and in
the pathogenesis of animal infections. Proc. Natl. Acad. Sci. U. S. A. 97, 5510–5515
of surface proteins to the cell wall of Staphylococcus aureus. Mol. Microbiol. 40, 1049–
1057
154
9. Weiss, W. J., Lenoy, E., Murphy, T., Tardio, L., Burgio, P., Projan, S. J., Schneewind,
O., and Alksne, L. (2004) Effect of srtA and srtB gene expression on the virulence of
480–486
10. Maresso, A. W., and Schneewind, O. (2006) Iron acquisition and transport in
Staphylococcus aureus. Biometals Int. J. Role Met. Ions Biol. Biochem. Med. 19, 193–203
11. Maresso, A. W., Chapa, T. J., and Schneewind, O. (2006) Surface protein IsdC and
sortase B are required for heme-iron scavenging of Bacillus anthracis. J. Bacteriol. 188,
8145–8152
12. Mazmanian, S. K., Ton-That, H., Su, K., and Schneewind, O. (2002) An Iron-Regulated
13. Schneewind, O., Model, P., and Fischetti, V. A. (1992) Sorting of protein A to the
14. Frankel BA, Kruger RG, Robinson DE, Kelleher NL, McCafferty DG (2005)
Staphylococcus aureus sortase transpeptidase SrtA: insight into the kinetic mechanism
11200.
15. Huang, X., Aulabaugh, A., Ding, W., Kapoor, B., Alksne, L., Tabei, K., and Ellestad, G.
11307–11315
155
16. Perry, A. M., Ton-That, H., Mazmanian, S. K., and Schneewind, O. (2002) Anchoring of
surface proteins to the cell wall of Staphylococcus aureus. III. Lipid II is an in vivo
277, 16241–16248
17. Ruzin, A., Severin, A., Ritacco, F., Tabei, K., Singh, G., Bradford, P. A., Siegel, M. M.,
Projan, S. J., and Shlaes, D. M. (2002) Further evidence that a cell wall precursor [C(55)-
2141–2147
18. Schneewind, O., Fowler, A., and Faull, K. F. (1995) Structure of the cell wall anchor of
20. Mazmanian, S. K., Skaar, E. P., Gaspar, A. H., Humayun, M., Gornicki, P., Jelenska, J.,
21. Ton-That, H., Mazmanian, S. K., Alksne, L., and Schneewind, O. (2002) Anchoring of
surface proteins to the cell wall of Staphylococcus aureus. Cysteine 184 and histidine 120
of sortase form a thiolate-imidazolium ion pair for catalysis. J. Biol. Chem. 277, 7447–
7452
156
22. Frankel BA, Tong Y, Bentley ML, Fitzgerald MC, McCafferty DG (2007) Mutational
Biochemistry 46:7269–7278.
23. Bentley, M. L., Lamb, E. C., and McCafferty, D. G. (2008) Mutagenesis Studies of
24. Suree, N., Liew, C. K., Villareal, V. A., Thieu, W., Fadeev, E. A., Clemens, J. J., Jung,
M. E., and Clubb, R. T. (2009) The Structure of the Staphylococcus aureus Sortase-
Substrate Complex Reveals How the Universally Conserved LPXTG Sorting Signal Is
25. Tian, B.-X., and Eriksson, L. A. (2011) Catalytic Mechanism and Roles of Arg197 and
Thr183 in the Staphylococcus aureus Sortase A Enzyme. J. Phys. Chem. B 115, 13003–
13011
26. Liew, C. K., Smith, B. T., Pilpa, R., Suree, N., Ilangovan, U., Connolly, K. M., Jung, M.
E., and Clubb, R. T. (2004) Localization and mutagenesis of the sorting signal binding
27. Marraffini, L. A., Ton-That, H., Zong, Y., Narayana, S. V. L., and Schneewind, O.
conserved arginine residue is required for efficient catalysis of sortase A. J. Biol. Chem.
279, 37763–37770
157
28. Zhang, R., Wu, R., Joachimiak, G., Mazmanian, S. K., Missiakas, D. M., Gornicki, P.,
aureus and Bacillus anthracis Reveal Catalytic Amino Acid Triad in the Active Site.
29. Weiner, E. M., Robson, S., Marohn, M., and Clubb, R. T. (2010) The Sortase A enzyme
that attaches proteins to the cell wall of Bacillus anthracis contains an unusual active site
30. Race, P. R., Bentley, M. L., Melvin, J. A., Crow, A., Hughes, R. K., Smith, W. D.,
Sessions, R. B., Kehoe, M. A., McCafferty, D. G., and Banfield, M. J. (2009) Crystal
31. Ilangovan, U., Ton-That, H., Iwahara, J., Schneewind, O., and Clubb, R. T. (2001)
Structure of sortase, the transpeptidase that anchors proteins to the cell wall of
32. Khare, B., Krishnan, V., Rajashankar, K. R., I-Hsiu, H., Xin, M., Ton-That, H., and
Housekeeping and Pilus-Specific Sortases: SrtA and SrtC1. PLoS ONE 6, e22995
33. Zong, Y., Mazmanian, S. K., Schneewind, O., and Narayana, S. V. L. (2004) The
158
34. Jung, M. E., Clemens, J. J., Suree, N., Liew, C. K., Pilpa, R., Campbell, D. O., and
analogue for covalent inhibition of sortases. Bioorg. Med. Chem. Lett. 15, 5076–5079
35. Kabsch, W. (2010) XDS. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132
36. Karplus, P. A., and Diederichs, K. (2012) Linking Crystallographic Model and Data
37. Diederichs, K., and Karplus, P. A. (2013) Better models by discarding data? Acta
38. McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C., and
39. Emsley, P., Lohkamp, B., Scott, W. G., and Cowtan, K. (2010) Features and development
40. Emsley, P., and Cowtan, K. (2004) Coot: model-building tools for molecular graphics.
41. Smart, O. S., Womack, T. O., Flensburg, C., Keller, P., Paciorek, W., Sharff, A.,
42. Bricogne, G. (1993) Direct phase determination by entropy maximization and likelihood
ranking: status report and perspectives. Acta Crystallogr. D Biol. Crystallogr. 49, 37–60
159
43. Kruger, R. G., Dostal, P., and McCafferty, D. G. (2004) Development of a high-
performance liquid chromatography assay and revision of kinetic parameters for the
44. Phillips, J. C., Braun, R., Wang, W., Gumbart, J., Tajkhorshid, E., Villa, E., Chipot, C.,
Skeel, R. D., Kalé, L., and Schulten, K. (2005) Scalable molecular dynamics with
45. Lindorff-Larsen, K., Piana, S., Palmo, K., Maragakis, P., Klepeis, J. L., Dror, R. O., and
Shaw, D. E. (2010) Improved side-chain torsion potentials for the Amber ff99SB protein
46. Kräutler, V., van Gunsteren, W. F., and Hünenberger, P. H. (2001) A fast SHAKE
47. Darden, T., York, D., and Pedersen, L. (1993) Particle mesh Ewald: An N⋅log(N) method
48. Wang, J., Wang, W., Kollman, P. A., and Case, D. A. (2006) Automatic atom type and
bond type perception in molecular mechanical calculations. J. Mol. Graph. Model. 25,
247–260
49. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A., and Case, D. A. (2004)
Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–
1174
160
50. Feller, S. E., Zhang, Y., Pastor, R. W., and Brooks, B. R. (1995) Constant pressure
molecular dynamics simulation: The Langevin piston method. J. Chem. Phys. 103, 4613–
4621
51. Martyna, G. J., Tobias, D. J., and Klein, M. L. (1994) Constant pressure molecular
52. Schrödinger, L. (2010) The PyMOL Molecular Graphics System, Version 1.3r1.
53. D.A. Case, T.A. Darden, T.E. Cheatham, III, C.L. Simmerling, J. Wang, R.E. Duke, R.
Luo, R.C. Walker, W. Zhang, K.M. Merz, B. Roberts, S. Hayik, A. Roitberg, G. Seabra,
J. Swails, A.W. Goetz, I. Kolossváry, K.F. Wong, F. Paesani, J. Vanicek, R.M. Wolf, J.
Liu, X. Wu, S.R. Brozell, T. Steinbrecher, H. Gohlke, Q. Cai, X. Ye, J. Wang, M.-J.
Hsieh, G. Cui, D.R. Roe, D.H. Mathews, M.G. Seetin, R. Salomon-Ferrer, C. Sagui, V.
Babin, T. Luchko, S. Gusarov, A. Kovalenko, and P.A. Kollman (2012) AMBER 12,
54. Jiang, W., Luo, Y., Maragliano, L., and Roux, B. (2012) Calculation of Free Energy
55. Souaille, M., and Roux, B. (2001) Extension to the weighted histogram analysis method:
combining umbrella sampling with free energy calculations. Comput. Phys. Commun.
135, 40–57
56. Schechter, I., and Berger, A. (1967) On the size of the active site in proteases. I. Papain.
161
57. Bentley, M. L., Gaweska, H., Kielec, J. M., and McCafferty, D. G. (2007) Engineering
the substrate specificity of Staphylococcus aureus Sortase A. The beta6/beta7 loop from
58. Piotukh, K., Geltinger, B., Heinrich, N., Gerth, F., Beyermann, M., Freund, C., and
59. Mariscotti, J. F., García-del Portillo, F., and Pucciarelli, M. G. (2009) The Listeria
60. Comfort, D., and Clubb, R. T. (2004) A Comparative Genome Analysis Identifies
61. Kruger RG et al. (2004) Analysis of the Substrate Specificity of the Staphylococcus
62. Nicola, G., Peddi, S., Stefanova, M., Nicholas, R. A., Gutheil, W. G., and Davies, C.
Tripeptide Boronic Acid Inhibitor: A Role for Ser-110 in Deacylation. Biochemistry 44,
8207–8217
63. Naik, M. T., Suree, N., Ilangovan, U., Liew, C. K., Thieu, W., Campbell, D. O., Clemens,
transpeptidase. Calcium promotes sorting signal binding by altering the mobility and
162
64. Connolly, K. M., Smith, B. T., Pilpa, R., Ilangovan, U., Jung, M. E., and Clubb, R. T.
163
Chapter 5
164
The work described in this chapter is a version of a manuscript to be submitted for publication.
Jacobitz A. W., Yi, S. W., McConnell, S., Peterson, R., Jung, M. E., and Clubb, R. T. 2015.
I was the primary author of this work. I designed experiments, performed biochemical assays and
165
5.1 Overview
Sortase enzymes perform a transpeptidation reaction that covalently links their two
substrates via a new peptide bond. While most sortase enzymes are known for catalyzing this
transpeptidation between their primary protein substrate and the cell wall peptidoglycan, a subset
of these enzymes instead covalently links proteins to each other via isopeptide bonds to produce
elongated protein polymers known as pili. These protein-polymerizing class C sortases are
structurally distinguished by the presence of an N-terminal extension that occludes the active
site. Numerous publications have hypothesized that the lid is mobile in solution as this would be
necessary for the binding and catalysis of the enzyme’s substrate. Here we show using NMR
dynamics measurements and in vitro assays that the lid of wild-type Streptococcus pneumoniae
SrtC1 is rigid in solution and does not experience conformational dynamics on a mechanistically
relevant timescale. Additionally, we show that point mutations to the lid induce dynamic
behavior that correlates with an increase in both the hydrolysis of the primary substrate and
general access to the active site cysteine residue as evidenced by increased oxidation. These
results suggest that the lid of the S. pneumoniae SrtC1 enzyme performs a regulatory function,
and they imply that interaction with other regions of the full length protein substrate, or other as
yet unnamed surface factors, are likely required to activate the enzyme in vivo.
166
5.2 Introduction
Gram-positive bacteria commonly utilize their thick cell walls as a scaffold for anchoring
proteins necessary for interacting with their environment. Many of these proteins are covalently
attached to the cell wall through a reaction catalyzed by sortase enzymes. A particular subtype of
sortases, the class C sortases, are not only implicated in the attachment of surface proteins to the
cell wall of Gram-positive organisms but have also been shown to polymerize pili: long
filamentous protein fibers that are used by some bacteria to make initial, long-range contacts to
host cells1–3. These pili are thought to provide a preliminary point of attachment to the host cell
which acts like a grappling hook, allowing the bacterium to swing closer to the cell, thereby
facilitating the formation of additional close-range adhesive interactions via auxiliary surface
anchored adhesive proteins3. Pili have subsequently been shown to increase the virulence of
certain bacterial species3–5, participate in immune system evasion6, are implicated in biofilm
formation7,8 and have also been heavily researched for their potential as vaccine components due
Sortase enzymes are ubiquitous in Gram-positive bacteria and can be classified into 5
distinct classes based on their sequence and structural similarities12,13. Classes A, B, D, E, and F
are generally considered to be cell wall anchoring sortases, while the class C sortases stand alone
in their ability to polymerize their protein substrates (it should be noted that there have recently
been two sortases which have been structurally defined as class B that are in fact pilus associated
sortases)14,15. The cell wall anchoring sortases all recognize lipid II, a precursor to cell wall
synthesis, as their secondary substrate. Class A enzymes are often referred to as “housekeeping”
enzymes; they are constitutively expressed, and generally anchor a variety of proteins to the cell
wall. Class B, D, E, and F enzymes on the other hand are generally utilized for attaching a
167
particular subclass of proteins to the cell wall and are typically expressed along with their
substrate(s) only under specific conditions. The class C enzymes are similar to the class B, D, E
and F enzymes in that they are also sequestered in specific genomic islands along with their pilus
protein substrates, but they are unique in that most class C enzymes are known to recognize two
distinct positions on a single protein to polymerize it into a pilus instead of utilizing lipid II as
their secondary substrate. When fully polymerized, these pili can contain hundreds of individual
A large number of sortase structures have been solved and they all share a common 8-β-
barrel sortase fold and a His-Cys-Arg catalytic triad13,16. Each of these enzymes recognizes a C-
terminal cell wall sorting signal (CWSS) composed of a 5 residue sorting signal motif, followed
by a transmembrane helix and positively charged cytoplasmic anchoring domain. The 5 residue
sorting signal is generally of the form LPXTG, where the Pro and Thr residues are most highly
conserved, X can be any amino acid, and the first and last residue often confer specificity to the
individual sortase12,13,17. The reaction begins when the sortase binds the sorting signal and the
active site cysteine performs a nucleophilic attack on the threonine’s carbonyl carbon. This
active site arginine17, and the active site histidine is thought to act as an acid to protonate the
terminal amine on the glycine of the leaving group18,19. The tetrahedral intermediate then quickly
collapses to form a semi-stable thioacyl intermediate between the active site cysteine and the
protein substrate. For the more heavily studied cell wall anchoring sortases, the next step of the
reaction requires the binding of cell wall precursor lipid II. The cross-bridge peptide of the lipid
II molecule varies from organism to organism, but will invariably contain a terminal amine
group which is deprotonated by the active site histidine before performing a nucleophilic attack
168
on the Thr carbonyl carbon, leading to a second tetrahedral intermediate which collapses to form
a new peptide bond between the sorting signal threonine and the cross-bridge peptide13. Pilin
polymerizing sortases follow the same basic mechanism but utilize the amino-nucleophile of a
lysine sidechain within their pilin protein substrate to resolve the thioacyl intermediate and
generate an isopeptide bond between the sorting signal of one monomer and the lysine of
another2,3,10. That is, the first of these two recognition motifs is the same CWSS at the C-
terminus of the protein substrate, but the other is typically a 4 residue lysine-containing pilin
Several species of bacteria are known to harbor pilus gene clusters typically containing
genes for 2 or 3 pilin proteins and 1 to 3 sortases. Interestingly, these class C sortases are the
only known class of sortase to harbor an N-terminal extension known as the “lid” that
completely occludes the active site13,20–22(Fig. 1). The lid region has been found in all class C
sortases studied to date and invariably maintains a conserved DP(F/W/Y) anchor motif wherein
the aspartate residue forms a hydrogen bond with the active site arginine, and the aromatic at the
end of the motif is wedged into the active site where it forms a sulfur-aromatic interaction with
the active site cysteine22. Based on the fact that the lid completely blocks access to the sorting
signal binding site seen in all sortase-substrate complex structures reported to date17,23,24 along
with the observations that several class C sortase crystal structures are missing electron density
for regions flanking the lid anchor residues25–28 and many display elevated B-factors throughout
this region20–22, it has been proposed that the lid region of these sortases must be highly
flexible13,20–22,27.
169
Streptococcus pneumoniae is one of the most common causative agents of pneumonia,
and is also known to cause meningitis, and septicemia, and the ability of S. pneumoniae to cause
invasive disease is enhanced by the presence of pili4. S. pneumoniae can express two distinct
types of pili carried on separate genomic islands, known as pilus island 1 (PI-1) or pilus island 2
(PI-2)29. PI-1 is present in ~30% of S. pneumoniae strains and these pili have been more
thoroughly characterized. They consist of three pilus proteins, RrgA, RrgB, and RrgC, which are
polymerized by three sortases, SpnSrtC1, SpnSrtC2, and SpnSrtC3. Structures for each of these
sortases have been solved and they all display nearly identical structures, each with the
conserved lid occluding the active site20,25. Mutations to the anchor residues in the lid of
SpnSrtC1 have been shown to decrease the thermal stability of the enzyme in vitro21. These same
mutations were shown to alter the pattern of, but not abrogate, the production of polymerized
RrgB species in vitro20, but similar mutations to either Streptococcus agalactiae sortase C-1
(SagSrtC1) or Actinomyces oris sortase C-1 (AoSrtC1) had no visible effect on pilus
polymerization in vivo30,31.
Here we show that despite a number of papers previously referring to the lid region of
SrtC enzymes as flexible, the lid of SpnSrtC1 is not in fact dynamic in solution, but instead has
two dynamic regions flanking the lid which may function as hinges, as has been proposed
previously by Manzano and coworkers20. Mutations to the anchor residues of the lid make the
region highly dynamic and dramatically increase hydrolysis of a sorting signal peptide in vitro.
We also show that the lid protects the active site cysteine from oxidation, and mutations to the lid
abolish this protective effect. Based on these findings, we propose that the lid performs a
regulatory function to prevent access to the active site in the absence of additional factors, likely
170
found on the cell surface, which prepare the protein for activity by mobilizing or altering the
5.3 Results
Soluble SpnSrtC1, residues 17-228 was uniformly labeled with 15N and 13C and purified
from E. coli. 94% of backbone resonances could be definitively assigned using standard triple
resonance assignment methods, leaving a few 1-2 residue unassigned segments, and a single 7
residue stretch of amino acids (residues Q167-E172) which could not be assigned. Interestingly,
this 7 residue stretch of amino acids forms the majority of the interface seen between the two
SpnSrtC1 monomers in the asymmetric unit of the 2008 crystal structure solved by Manzano and
timescale (i.e. binding and dissociation occurs roughly on the order of 1 ms) are frequently
broadened beyond detection, we hypothesized that these resonances could be missing due to the
determine whether this crystallographic dimer was indeed present in solution, we performed a
series of dilutions from 2 mM to 100 μM and recorded 1H-15N HSQC spectra. As the
concentration of the protein was lowered, significant changes in the spectrum were seen in
isolated resonances, along with the appearance of several additional peaks which could only be
detected in the lowest concentration sample (Fig. 2). Although new resonances which appear in
the 100 μM spectrum could not be assigned, as this concentration is too low for the use of
significantly less sensitive triple resonance experiments that are required for backbone
assignments, the existing assigned resonances which either experienced an increase in intensity
171
on dilution, or a change in chemical shift, could be identified and plotted onto the crystal
structure (Fig 2B). These residues are largely clustered in the binding interface observed in the
tumbling time for the protein, τc, which is defined as the amount of time it takes for the protein to
rotate through 1 radian in space. τc values of known globular proteins consistently yield values of
~0.6 ns times the molecular weight (MW) of the protein in kilodaltons. The MW of soluble
SpnSrtC1 is 23.813 kDa yielding a theoretical τc of 14.3 ns. Based on measurements of the T1
concentration of 1 mM. This τc is significantly higher than would be expected for a monomer of
this size. If the protein formed a stable dimer, the τc would be expected to be around 28.6 ns, and
a protein of this size would experience extremely fast T2 relaxation preventing the detection of
good quality triple resonance spectra, an effect we did not observe. The slightly elevated
dimerization for this protein that had been suggested by the broadening of interfacial residues
discussed above. This would indicate that at any given time, only a small percentage of the
available proteins in solution would be dimeric and tumble at the slower ~28.6 ns rate, but that
these dimers are still detectable as an increase in the average τc of all the proteins in the sample,
5.3.2 NMR dynamics data indicate that the SpnSrtC1 lid is not dynamic in vitro.
172
technique for dissecting the motions of individual backbone amides across a wide range of
biologically relevant timescales. First, T1, T2, and HetNOE experiments were performed to detect
the presence of fast-motions on the ps-ns timescale. The results of T1 and T2 experiments were
used to determine the rotational correlation time, τe, for each backbone amide residue in the
protein (Fig. 3). τe represents the amount of time it takes for a single amide bond vector to rotate
through 1 radian of space. By comparing the τe to the overall rotational correlation time for the
entire protein, τc, (Fig.3A, red line) residues whose bond vectors are moving more rapidly than
the body of the protein can be detected. Any residue that rapidly samples disordered states in
solution would be expected to have a τe that is much lower than the protein’s τc, indicating that
the residue can move through space more quickly than the body of the protein. This is
comparable to a flag rapidly flapping in the wind (the mobile residues) while affixed to the bow
of a slowly rocking ship (the body of the protein). The flag moves on the same slow timescale of
the ship, but will experience faster overall motion because of additional dynamic behavior.
Unexpectedly, lid residues contacting the active site (D58 and W60) have similar τe values as the
entire structured core domain of the protein, 16.4 and 16.9 ns respectively, compared to a τc of
17.8 ± 1.1 ns for the structured portion of the protein. For comparison, the unstructured C-
terminus of the protein has τe values that drop as low as 10 ns. This is confirmed by HetNOE
values above 0.6 which indicates that these residues are structured in solution (Fig. 3B), as
residues are typically only considered disordered if their HetNOE is below 0.6 (Fig. 3B, red
line). Notably, residues 56-57 before and 65-70 after the lid show increased dynamics on this
timescale indicating that they are highly mobile, even though the lid residues from this N-
terminal extension that are actually engaged with the active site are not. This supports the
hypothesis that the N-terminal extension contains “hinges,” originally proposed by Manzano and
173
coworkers to be present in SpnSrtC3 based on elevated B-factors for this protein being highest in
two discrete regions flanking the lid20. This hypothesis is further supported by the significant
number of SrtC structures solved to date wherein these regions either showcase the highest B-
factors in the structure, or lack density altogether20–22,25–28. Together these NMR experiments
indicate that the lid residues that contact the active site do not experience motions over the ps-ns
timescale, indicating that they are predominantly in a rigid, ordered conformation, likely akin to
that visualized in the crystal structure, whereas regions of sequence adjacent to the lid are in fact
highly mobile, and likely represent hinges for the opening of the lid and subsequent unmasking
additional experiments to determine if the lid was instead mobile over longer μs-ms timescales.
To measure this phenomenon, we first conducted CPMG experiments32 to directly detect the Rex
component of relaxation on a per residue basis (Fig. 3C). Rex values are generally considered to
be an indicator of this slower timescale motions if they are greater than ~30% of the measured T2
for an individual residue which typically means values above 10 s-1 are considered significant.
Residues in and around the lid region do not show significantly elevated Rex values, indicating
that they are in fact not mobile on the μs-ms timescale. Additional experiments to sample within
this timescale (CEST), and out to the ms-s timescale (NZ-exchange) were also conducted, and
showed no spectral perturbations which would be indicative of longer timescale motions (data
not shown). Together, these experiments indicate that the lid of SpnSrtC1 is not dynamic in
solution as has been suggested by previous reports and is instead largely structured in the
174
5.3.3 Mutations to the “lid” of SpnSrtC1 make the region highly dynamic in solution.
Given that the “lid” of wild-type SpnSrtC1 is not dynamic in vitro, we set out to
determine the effects of point mutations in the lid on its mobility in solution. Several previous
studies have shown that point mutations to the conserved Asp and Trp residues in the
DP(F/W/Y) motif of SpnSrtC1 lowered its thermostability21. Similar mutations were also shown
to increase the rate of hydrolysis and were even necessary to induce in vitro pilin polymerization
by SagSrtC130,33. To determine if these mutations modified the mobility of the lid, we generated
SpnSrtC1-D58A and W60A point mutants and characterized their effects on the protein’s
dynamics using the same fast-timescale T1, T2, and HetNOE experiments discussed above for the
WT enzyme. SpnSrtC1-W60A shows significantly increased mobility in the lid, and throughout
the entire region between the two hinges identified for the WT enzyme (residues 56-70). T1 and
T2 experiment derived τe values indicate that the region tumbles at 10.7 ns, 5.2 ns faster than the
average for the structured portion of this enzyme (Fig. 4). The HetNOE experiment additionally
indicates that these residues are largely disordered, with every residue that could be assigned and
characterized between the two hinges, residues 58 to 64, having reduced HetNOE values
compared to WT, with many of these being less than 0.6 indicating that they are disordered.
SpnSrtC1-D58A also shows significantly increased mobility in the lid, and especially in flanking
regions based on an analysis of the τe and HetNOE measurements (Fig. 4). Interestingly, these
experiments show decreased dynamics in a small region immediately after the conserved lid
residue, W60. To further investigate whether this less dynamic region indicates that the lid
tryptophan might be sampling a closed state similar to the WT enzyme, we performed additional
NOESY experiments and compared the spectra for the Trp indole proton. As the NOESY
experiment is able to detect the through-space coupling of nuclei within 5 Å of each other, the
175
pattern of numerous peaks seen for the WT enzyme indicates that the side chain of the residue is
folded and in close proximity to several other residues (Fig. 4E). The W60 indole in the D58A
mutant on the other hand is split into 3 peaks in the HSQC, each of which shares only 1 or 2
NOESY peaks with the WT spectrum that may represent intraresidue NOEs to adjacent protons
in the sidechain. This large difference in NOE pattern suggests that the tryptophan sidechain in
the D58A mutant no longer participates in native contacts to the enzyme active site, and has
likely become dislodged from the position it maintains in the WT enzyme. This suggests that
even though several of the residues following W60 do not appear to be dynamic based on
HetNOE and τe values, the conformation of this entire region has been altered significantly, and
Knowing that the SpnSrtC1 lid is rigid in solution and becomes mobile upon mutation,
we set out to determine whether this increased mobility would affect the ability of the enzyme to
function in vitro. To test this, we performed in vitro hydrolysis experiments by incubating the
sortase with a 15 amino acid peptide derived from RrgB and containing the IPQTG sorting signal
sequence for 24 hrs at 37°C and then separating the mixture via HPLC. We then utilized MALDI
mass spectrometry to confirm the identities of the resultant peaks in the chromatogram. By
separating the substrate peptide from the cleaved product using HPLC and measuring the
absorbance at 215 nm we could monitor both the amount of cleaved product peptide formed, and
substrate peptide remaining, allowing for a thorough disambiguation of the reaction progress
which is not possible with the more commonly used FRET based assay which only reports on the
breakage of the peptide bond and not the subsequent release of the thioacyl intermediate from the
176
enzyme. We conducted the experiments with WT SpnSrtC1, SpnSrtC1-D58A and W60A point
mutants used in our NMR studies, as well as an SpnSrtC1-D58A-W60A double mutant and an
additional mutant wherein the entire lid containing region, residues 55-69 inclusive, was replaced
approximation of a completely open state of the lid, since a construct where we completely
deleted the N-terminal extension containing the lid proved to be unstable over the timecourse of
the reaction. Interestingly, though the WT enzyme has been shown to perform a transpeptidation
reaction in vitro with a full length RrgB protein which contains both the primary IPQTG sorting
signal substrate, and secondary YPKN pilin motif substrate20, the enzyme does not catalyze the
initial cleavage step of the reaction in vitro in the absence of its secondary substrate (Fig. 5). It is
important to note that not only is the WT enzyme incapable of hydrolyzing the substrate, as
evidenced by the lack of detectable product peptide, but it is also incapable of beginning its
reaction by performing a nucleophilic attack on the IPQTG motif’s Thr carbonyl to catalyze the
creation of the thioacyl intermediate. This latter reaction could be visualized in our assay as a
reduction in the amount of substrate peptide remaining at the end of the reaction without the
appearance of a new product peak, but this was not observed. Instead, the WT SpnSrtC1 reaction
is indistinguishable from peptide alone or from a reaction with the C193A active site mutant. The
lid mutants on the other hand, which we showed by NMR to have much more mobile lid regions,
all catalyze a hydrolysis reaction to a similar extent within the error of the measurement except
for the D58A mutant which was slightly slower, perhaps due to partial occlusion of the active
site by one or more of the multiple semi-stable positions we detected for its lid Trp residue based
on NOESY data. While even the mutant enzymes only process about 15% of their substrate in 24
hrs under the conditions tested, this rate of cleavage is similar to that seen for the WT SrtB
177
enzyme from Staphylococcus aureus (Jacobitz unpublished data) indicating that this could be a
5.3.5 A closed lid is important for protecting the active site cysteine from oxidation.
To further test the importance of the closed lid, we performed disulfide bonding
carbobenzyloxy protecting group and T* is a modified threonine residue with a thiol group in
disulfide bond with the active site cysteine. This type of modified substrate has been used
previously to successfully produce stable enzyme substrate complexes for 3 other sortases17,23,24,
indicating the utility of this method. When incubated with the IPQT* substrate at room
temperature for 24 hrs, the WT SpnSrtC1 showed no detectable disulfide formation by MALDI-
MS (Fig. 6A). When the same experiment was repeated with the SpnSrtC1-W60A mutant, a
mass shift of 595 Da (theoretical mass of Cbz-IPQT* = 593.4 Da) could be detected indicating
the formation of a disulfide-bonded complex (Fig. 6B). The difference in cysteine availability
highlighted by this experiment further indicates that the lid of WT SpnSrtC1 not only prevents
unwanted catalytic activity, but actually physically limits access to the active site cysteine, which
would be able to form a disulfide bond whether or not the enzyme was in a fully catalytically
active state.
5.4 Discussion
Class C sortases are unique among sortases in that they catalyze the covalent linkage of
two protein substrates instead of attaching proteins to the cell wall peptidoglycan13. This reaction
178
leads to the polymerization of long filamentous pili which are necessary for adhesion to host
cells and biofilm formation. Although the class C sortase enzymes utilize a lysine side chain
from their protein substrate as the nucleophile to conclude their reaction instead of the N-
terminus of a lipid II cross bridge peptide that other classes employ, the basic reaction is highly
similar, perhaps owing to the high degree of conservation within the core β-barrel of the sortase
motif. All sortases studied to date utilize the same conserved His-Cys-Arg catalytic triad, and are
thought to recognize their substrates using the same binding groove beginning at the β6/β7 loop
and moving towards the active site. Given the number of similarities shared between members of
this family it is perhaps unsurprising that we observed the dimerization of SpnSrtC1 in solution
(Fig. 2), as dimerization in solution has already been detected for several other members of the
family34–36. Interestingly, while the details of this dimerization event are not fully understood,
and each of the proposed binding interfaces is somewhat variable, all of these proteins have
utilized residues from the β4-β5 loop, on the “back” face of the enzyme, opposite the active site.
This dimerization was recently shown to alter the rate of catalysis for S. aureus SrtA in vivo36. It
is interesting to note that the vast majority of sequences deposited for SpnSrtC1 that can be
identified from a BLAST search with the 279 residue UniProt sequence, Q97SB9, referenced in
previous studies contain exactly 21 additional N-terminal residues, and analysis of this full 300
residue sequence with the TMHMM server37 predicts an additional N-terminal transmembrane
(TM) helix with >99.9% probability that could not have been predicted from the shortened
sequence referenced previously. This predicted N-terminal TM-helix has been shown to be
commonplace among other class C sortases and is known to be necessary for determining
appropriate membrane localization30. Surprisingly, both N and C-termini from both structures in
the proposed dimer visualized in the SpnSrtC1 crystal structure all localize to a single face of the
179
dimeric assembly which puts the two active sites on opposite ends, each facing solvent. It is
possible that this dimer is significant in vivo when faced with a full length RrgB substrate that
contains both the primary and secondary recognition motifs as these two motifs on the RrgB
protein are separated from each other by 95 Å which is significantly longer than the 35 Å that
separate the two sortase active sites in the dimer. This indicates that a single tri-domain RrgB
protein could feasibly interact with both active sites of the dimer simultaneously, potentially
allowing a single sortase dimer to polymerize an entire pilus by simply adding monomers
iteratively at one active site then the next without the need for sortases to “line up” to carry out a
In addition to their functional distinction, class C sortases are also structurally unique in
that they exhibit an N-terminal extension which occludes the active site and has thus been termed
the lid13,20–22. Numerous publications have referred to this lid as “flexible” based on elevated B-
factors or missing density in adjacent regions of crystal structures, or based on the fact that
removal of the lid would be necessary for substrates to access the active site13,20–22,27. NMR
dynamics data indicates to the contrary that in solution the key lid residues that actually contact
the active site are not dynamic, but are instead flanked by dynamic regions which potentially act
as hinges for the removal of the lid and unmasking of the active site under the appropriate
circumstances (Fig. 3). Additional analysis of the ability of the WT enzyme to perform the initial
cleavage step of its reaction in vitro with purified peptide containing its primary IPQTG sorting
signal substrate indicates that this closed conformation of the lid actually prevents catalysis. Only
when mutations to the lid are introduced which increase the flexibility of this region, as indicated
by reduction in both the τe and HetNOE values for these residues (Fig. 4), does the enzyme
catalyze the initial step of its reaction in vitro (Fig. 5). The same motion-inducing mutations are
180
also required to grant access to the cysteine even for the simple production of a disulfide bond,
the formation of which does not require access to the rest of the binding site or participation from
any of the other active site residues (Fig. 6). This indicates that not only does the rigidity of the
lid in the WT enzyme prevent the hydrolysis of the enzyme’s native substrate in solution, but
that it does so by completely preventing access to the active site cysteine instead of simply
Numerous publications have confirmed the enzyme’s ability to perform this reaction
under native conditions in vivo25,38,39, and it has even been shown to perform the native
transpeptidation reaction in vitro when presented with full length substrates containing both its
primary, IPQTG sorting signal, and secondary, YPKN pilin motif substrates20,21. All other
classes of sortase enzymes studied to date have been shown to catalyze the hydrolysis of their
primary substrate in vitro when deprived of a secondary substrate with which to perform their
native transpeptidation reaction28,30,35,40–43. Since SpnSrtC1 cannot perform the initial cleavage of
its primary binding motif in vitro in the absence of secondary substrate as numerous other
sortases have been shown to do (without the artificial mobilization of its lid through the
must be required to remove the lid and allow for catalysis. Interestingly, our in vitro data also
suggests that in the absence of a secondary substrate or other additional factors, WT SpnSrtC1
does not even form a thioacyl intermediate, as this would be detectable as a loss of substrate,
without an accompanying formation of hydrolysis product (Fig. 5). This indicates that the
thioacyl intermediate reported previously from in vitro reactions with full length protein
substrates can only form when the lid is opened by additional interactions with the substrate that
181
are outside of the sorting signal21. It is tempting to assume that YPKN pilin motif would fulfill
the role of this additional factor that unhinges the lid, but simply based on the fact that this
enzyme has been shown to catalyze the transpeptidation in vitro when provided with a full length
substrate, it is only possible to conclude that there is another region of the RrgB substrate protein
outside of the IPQTG sorting signal that activates the sortase, presumably by opening the lid.
Unfortunately even this conclusion is not infallible, as the data for the in vitro reaction is not
entirely convincing given that we, and others, have observed soluble pilin proteins to associate
into stable higher molecular weight species which closely mimic product of the sortase catalyzed
15
polymerization reaction (unpublished data, and ). Additionally, the WT S. agalactiae SrtC1
protein was shown to cleave its primary substrate in vitro, but still does not perform the
transpeptidation reaction in the presence of a second pilin motif containing peptide. Mutations to
the lid of this protein were shown to increase the rate of peptide hydrolysis in vitro by ~2.5 fold,
and had the additional activating effect of permitting transpeptidation in solution30,33. The fact
that both the S. pneumoniae system studied in this work and the S. agalactiae system studied
previously are not fully active in vitro with an intact lid reiterates the necessity for additional
factors, outside of the two well-known substrate motifs, to open the lid of the enzyme and allow
catalysis to proceed as it does in vivo. Sortase enzymes have been observed to colocalize with a
number of factors on the cell surface, including the SEC translocon, cell wall synthesis
machinery, and their substrates13,44–46, and we postulate that either additional regions within the
pilin proteins themselves, or one of these additional surface factors must be necessary for
From this evidence it is clear that the lid present in class C enzymes is competent in
preventing the reaction from taking place. But why would an entire class of enzymes have
182
evolved a conserved feature to prevent or retard their primary function? First and foremost, the
hydrolysis and subsequent loss of protein substrates to solvent, which are intended to instead be
polymerized and attached to the cell surface, is inherently wasteful for the cell. While accidental
cleavage and loss of a few proteins here and there (as may be common for cell wall anchoring
sortases) might be acceptable to the cell, the energetic cost of building a 100 protein, 1 μm long
pilus, and then subsequently cleaving its C-terminus and releasing it into solvent could be
crippling especially if this was allowed to happen at regular intervals. It’s also possible that
preventing the functionality of the enzyme is simply a byproduct of protecting the active site
from inactivation by oxidation of its cysteine. It has been shown previously that the cysteines of
several other sortase enzymes are susceptible to oxidation as evidenced by both an increase in
catalytic activity reported in the presence of reducing agents, and structures of Streptococcus
pyogenes SrtA solved with its active site cysteine in multiple oxidation states47,48. It has even
been proposed that certain host organisms may create an oxidizing environment surrounding a
site of infection to deactivate sortase enzymes at the cell surface and render those cells
avirulent48,49. The fact that we observed rapid formation of a disulfide linked complex between
an IPQT* sorting signal mimic and the W60A lid mutant of SpnSrtC1, but not between the same
mimic and the WT protein suggests that the lid indeed functions to prevent oxidation of the
active site cysteine (Fig. 6). When tested in vivo, SrtC1 lid mutants showed no discernable effect
on transpeptidation30,31. This is in agreement with our proposal that the lid is necessary largely to
avoid the energetic penalties associated with cleavage and loss of large numbers of pilus proteins
or oxidative inactivation of the sortase itself, as these energetic penalties would likely only
provide a fitness advantage in a host setting where slight improvements in efficiency could be
pivotal in determining the cell’s ability to maintain a successful infection. We thus propose that
183
the lid functions as a regulatory mechanism for the class C sortase that prevents the unwanted
cleavage of substrates, and protects the active site from oxidative inactivation.
DNA encoding codon optimized soluble SpnSrtC1, residues 17-228 based on uniprot
sequence Q97SB9, was generated by recursive PCR from overlapping primers50. Codon
optimization was performed with OPTIMIZER51. DNA encoding SpnSrtC1 was then cloned into
the pESUMO vector (LifeSensors) and transformed into Escherichia coli BL21 (DE3) cells
thiogalactopyranoside (IPTG) for 5 hrs at 37 °C. Protein was purified by affinity purification
using HisPur Co2+ resin (Thermo) as per the manufacturer’s instructions. The His6-SUMO tag
was cleaved by incubating the protein overnight at 4 °C with recombinant ULP1 protease and
subsequently removed by reapplying the protein mixture to HisPur Co2+ resin. D58A, W60A,
D58A-W60A double mutant, and GSlid mutants of SpnSrtC1 were generated by site directed
15 13
mutagenesis using standard procedures. N and C labeled variants were produced by
expressing the proteins in M9-minimal media supplemented with 15N NH4Cl and/or 13C glucose.
For these expressions, cells were grown to an OD600 of 0.5 at 37 °C, then transferred to an 18 °C
incubator and induced with 1 mM IPTG for 16 hrs. Labeled proteins were purified as specified
above.
184
5.5.2 NMR spectroscopy.
All protein samples used for NMR experiments were concentrated and dialyzed into
NMR buffer (50 mM sodium phosphate pH 6.5, 50 mM NaCl, 10% D2O). NMR spectra were
recorded on Bruker 500, 600, and 800 MHZ spectrometers equipped with triple resonance
cryogenic probes and processed using NMRPipe52. Chemical shift assignments were obtained
performed by collecting 15N-HSQC spectra beginning at 2 mM, and diluting the protein serially
with NMR buffer + 10% D2O to 1.5, 1, 1.5, 0.5, 0.25, 0.2, and 0.1 mM. The number of scans was
increased to account for a loss of signal from a reduction of concentration based on the following
relationship:
NS = NS0(C1/C2)2
Analysis of signal intensities and chemical shift changes was performed in SPARKY54, and
described above. Heteronuclear NOE, T1 and T2 experiments were conducted at 298 K and
analyzed in SPARKY54. The residue specific tumbling time, τe was calculated using a spherical
1
= 6 −7
4
where νN is the 15N resonance frequency in Hz. The molecular tumbling time, τc, was calculated
as the average of all τe values for the protein. CPMG experiments were conducted using a
185
modified approach based on the original method developed by Palmer et. al.56. Here, full T2
datasets were collected for 3 different CPMG pulse delays, 0.0023 s, 0.0048 s, and 0.000514 s.
R2 values (1/T2) from these measurements were used to calculate Rex by subtracting the
maximum calculated R2 (either the 0.0023 s or 0.0048 s delay, as R2 values calculated at these
longer delays are known to oscillate slightly), from the minimum R2 determined at the shortest
SpnSrtC1 and mutants were dialyzed into assay buffer (25 mM TRIS-HCl pH 7.5, 100
by LifeTein and used without further purification. Lyophilized peptide was dissolved in assay
buffer to 2 mM and final concentration was verified using a BCA assay kit (Thermo) per
manufacturer’s instructions. Reactions were conducted with 100 μM enzyme and 200 μM
peptide in a total volume of 100 μL at 37 °C. After 24 hrs, reactions were quenched by addition
of 50 μL 1 M HCl and 100 μL of the reaction mixture was injected onto a Waters XSelect HSS
C18 reversed phase HPLC column. Peptides were eluted with a gradient from 5-30% acetonitrile,
0.1% trifluoroacetic acid over 20 mins. Elution and quantification of peptides was conducted by
monitoring absorbance of the peptide bond at 215 nm. Fractions from peaks corresponding to
substrate and product were collected and their identities confirmed by matrix assisted laser
186
5.5.4 In vitro oxidation assay.
synthesized as in 57. The compound was mixed with SpnSrtC1 WT or SpnSrtC1-W60A mutant at
a 10:1 molar ratio of mimic to protein and allowed to mix on a rotisserie at room temperature for
24 hrs. Reactions were then applied to zip-tip C18 pipette tips (Millipore) and eluted as per
manufacturer’s instructions. 1 μL eluent was then spotted onto a MALDI plate in triplicate and
crystallized by the addition of 0.5 μL 2,5-dihydroxybenzoic acid (DHB) matrix dissolved in 50%
ethanol, 0.1% TFA. Formation of a disulfide bond with the mimic was detected via MALDI-MS
187
5.6 Figures
A B C
(A) Cartoon representation of SpnSrtC1 from PDB 2W1J20. Active site His, Cys, Arg, and lid Trp
residue shown as sticks. N-terminal extension preceding the sortase β-barrel core is colored red.
(B) Transparent surface representation of SpnSrtC1, showing the lid occluding the active site.
(C) Zoom in of SpnSrtC1 active site, showing conserved DP(F/W/Y) lid motif in active site as
sticks.
188
A
(A) HSQC of 300 μM SpnSrtC1 (green) overlayed over an HSQC of 100 μM SpnSrtC1 (red),
showing the appearance of several peaks along with the movement or increase in intensity of
several others. (B) The two monomers in the asymmetric unit of the crystal structure of
SpnSrtC1, PDB ID 2W1J20, are shown as a cartoon representation. Residues with increases in
peak height or changes in chemical shift of greater than 1 stdev from the mean are displayed on
the structure in red, unassigned residues are shown in beige, and both are displayed as sticks.
189
A 30
25
20
Te (ns)
15
10
5
0
B 1.5
1
0.5
HetNOE
0
-0.5
-1
-1.5
-2
C 50
40
Rex (s-1)
30
20
10
0
Residue
A diagram of secondary structure elements is shown at the top of the figure where helices are
represented as cylinders and β-strands as arrows. (A) Residue specific tumbling time (τe) for
each assigned residue in SpnSrtC1. Average for the structured region of the protein plotted as a
blue line. (B) Heteronuclear NOE for each residue in SpnSrtC1. A red line is shown at a HetNOE
value of 0.6. A HetNOE value below this line indicates that the residue can be considered
190
dynamic. (C) Rex component of relaxation as determined by direct CPMG detection, plotted for
each residue.
191
Figure 5.4 NMR shows SpnSrtC1 mutants have more dynamic lids.
A diagram of secondary structure elements is shown at the top where helices are represented as
cylinders and β-strands as arrows. (A) Residue specific tumbling time (τe) for each assigned
192
residue in SrtC1-W60A. (B) Heteronuclear NOE for each residue in SpnSrtC1-W60A. Red line
at 0.6 indicates the cutoff below which residues are considered disordered (C) Residue specific
tumbling time (τe) for each assigned residue in SrtC1-D58A. (D) Heteronuclear NOE for each
residue in SpnSrtC1-D58A. Red line at 0.6 indicates the cutoff below which residues are
considered disordered. (E) Comparison of NOESY spectra for the W60 indole proton for the WT
SpnSrtC1 protein, left, or three possible indole configurations seen in the D58A mutant,
193
A
B
IPQTG Remaining (µM)
250
200
150
100
50
0
(A) Product formed after 24 hr reactions with IPQTG containing peptide is shown for WT, and
various SpnSrtC1 mutant constructs. (B) The amount of substrate remaining after 24 hr
194
A 100
23689.60
B 100
23607.04
90
% Intensity
90
% Intensity
80
80 70
70 60
60 50
50 40
40 30
30 20
20 10
10 0
0 15000 19000 23000 27000 31000 35000
15000 19000 23000 27000 31000 35000
Mass (m/z) Mass (m/z)
23691.96 24202.25
100 100
90
% Intensity
90
% Intensity
80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
15000 19000 23000 27000 31000 35000 15000 19000 23000 27000 31000 35000
Mass (m/z) Mass (m/z)
Figure 5.6 The lid of SpnSrtC1 protects the active site cysteine from modification by a
sulfhydryl modifier.
(A) SrtC1 before (top) and after (bottom) 24 hr incubation with IPQT* disulfide bond forming
substrate mimic showing no significant formation of the disulfide complex. Theoretical mass of
SrtC1-WT = 23,813 Da, SrtC1-IPQT* complex = 24,405 Da (B) SrtC1-W60A mutant before
(top) and after (bottom) 24 hr incubation with IPQT* disulfide forming peptide showing
conversion to the disulfide bonded complex. Theoretical mass of SrtC1-W60A = 23,698 Da,
195
5.7 References
2008;16(1):33-40. doi:10.1016/j.tim.2007.10.010.
specific minor pilins to target human pharyngeal epithelial cells. Mol Microbiol.
2007;64(1):111-124. doi:10.1111/j.1365-2958.2007.05630.x.
2014;9(11):e113922. doi:10.1371/journal.pone.0113922.
7. Mishra A, Wu C, Yang J, Cisar JO, Das A, Ton-That H. The Actinomyces oris type 2
fimbrial shaft FimA mediates co-aggregation with oral streptococci, adherence to red
196
blood cells and biofilm development. Mol Microbiol. June 2010. doi:10.1111/j.1365-
2958.2010.07252.x.
8. Mishra A, Devarajan B, Reardon ME, et al. Two autonomous structural modules in the
fimbrial shaft adhesin FimA mediate Actinomyces interactions with streptococci and host
doi:10.1111/j.1365-2958.2011.07745.x.
9. Margarit I, Rinaudo CD, Galeotti CL, et al. Preventing bacterial infections with pilus-
doi:10.1086/595564.
10. Hendrickx APA, Budzik JM, Oh S-Y, Schneewind O. Architects at the bacterial surface -
sortases and the assembly of pili with isopeptide bonds. Nat Rev Microbiol.
2011;9(3):166-176. doi:10.1038/nrmicro2520.
11. Young PG, Moreland NJ, Loh JM, et al. Structural conservation, variability, and
12. Comfort D, Clubb RT. A Comparative Genome Analysis Identifies Distinct Sorting
doi:10.1128/IAI.72.5.2710-2722.2004.
13. Spirig T, Weiner EM, Clubb RT. Sortase enzymes in Gram-positive bacteria. Mol
197
14. Kang HJ, Coulibaly F, Proft T, Baker EN. Crystal Structure of Spy0129, a Streptococcus
doi:10.1371/journal.pone.0015969.
15. Shaik MM, Lombardi C, Maragno Trindade D, et al. A structural snapshot of type II pilus
doi:10.1074/jbc.M115.647834.
16. Clancy KW, Melvin JA, McCafferty DG. Sortase transpeptidases: insights into
doi:10.1002/bip.21472.
17. Jacobitz AW, Wereszczynski J, Yi SW, et al. Structural and computational studies of the
18. Frankel BA, Kruger RG, Robinson DE, Kelleher NL, McCafferty DG. Staphylococcus
aureus sortase transpeptidase SrtA: insight into the kinetic mechanism and evidence for a
doi:10.1021/bi050141j.
19. Frankel BA, Tong Y, Bentley ML, Fitzgerald MC, McCafferty DG. Mutational analysis of
2007;46(24):7269-7278. doi:10.1021/bi700448e.
198
20. Manzano C, Contreras-Martel C, El Mortaji L, et al. Sortase-mediated pilus fiber
1848. doi:10.1016/j.str.2008.10.007.
21. Manzano C, Izoré T, Job V, Di Guilmi AM, Dessen A. Sortase activity is controlled by a
2009;48(44):10549-10557. doi:10.1021/bi901261y.
22. Persson K. Structure of the sortase AcSrtC-1 from Actinomyces oris. Acta Crystallogr D
23. Suree N, Liew CK, Villareal VA, et al. The Structure of the Staphylococcus aureus
doi:10.1074/jbc.M109.022624.
24. Chan AH, Yi SW, Terwilliger AL, Maresso AW, Jung ME, Clubb RT. Structure of the
Bacillus anthracis Sortase A Enzyme Bound to Its Sorting Signal: a Flexible Amino-
25. Neiers F, Madhurantakam C, Fälker S, et al. Two crystal structures of pneumococcal pilus
sortase C provide novel insights into catalysis and substrate specificity. J Mol Biol.
2009;393(3):704-716. doi:10.1016/j.jmb.2009.08.058.
199
26. Khare B, Krishnan V, Rajashankar KR, et al. Structural differences between the
Streptococcus agalactiae housekeeping and pilus-specific sortases: SrtA and SrtC1. PLoS
27. Khare B, Fu Z-Q, Huang I-H, Ton-That H, Narayana SVL. The crystal structure analysis
of group B Streptococcus sortase C1: a model for the “lid” movement upon substrate
28. Cozzi R, Prigozhin D, Rosini R, et al. Structural Basis for Group B Streptococcus Pilus 1
doi:10.1371/journal.pone.0049048.
29. Danne C, Dramsi S. Pili of gram-positive bacteria: roles in host colonization. Res
30. Cozzi R, Malito E, Nuccitelli A, et al. Structure analysis and site-directed mutagenesis of
defined key residues and motives for pilus-related sortase C1 in group B Streptococcus.
SrtC2 required for membrane localization and assembly of type 2 fimbriae for
2539. doi:10.1128/JB.00093-12.
32. Mittermaier A, Kay LE. New Tools Provide New Insights in NMR Studies of Protein
200
33. Cozzi R, Zerbini F, Assfalg M, et al. Group B Streptococcus pilus sortase regulation: a
single mutation in the lid region induces pilin protein polymerization in vitro. FASEB J.
2013;27(8):3144-3154. doi:10.1096/fj.13-227793.
34. Lu C, Zhu J, Wang Y, et al. Staphylococcus aureus sortase A exists as a dimeric protein in
35. Robson SA, Jacobitz AW, Phillips ML, Clubb RT. Solution Structure of the Sortase
2012;51(40):7953-7963. doi:10.1021/bi300867t.
Staphylococcus aureus cell surface mediates its cell wall sorting activity. Exp Biol Med
37. Sonnhammer EL, von Heijne G, Krogh A. A hidden Markov model for predicting
transmembrane helices in protein sequences. Proc Int Conf Intell Syst Mol Biol.
38. Fälker S, Nelson AL, Morfeldt E, et al. Sortase-mediated assembly and surface topology
2958.2008.06396.x.
39. Shaik MM, Maccagni A, Tourcier G, Di Guilmi AM, Dessen A. Structural basis of pilus
2014;289(24):16988-16997. doi:10.1074/jbc.M114.555854.
201
40. Weiner EM, Robson S, Marohn M, Clubb RT. The Sortase A enzyme that attaches
proteins to the cell wall of Bacillus anthracis contains an unusual active site architecture. J
42. Maresso AW, Chapa TJ, Schneewind O. Surface protein IsdC and sortase B are required
doi:10.1128/JB.01011-06.
43. Ton-That, H., Mazmanian, S. K., Alksne, L., and Schneewind O. Anchoring of Surface
44. Hu P, Bian Z, Fan M, Huang M, Zhang P. Sec translocase and sortase A are colocalised in
2008;53(2):150-154. doi:10.1016/j.archoralbio.2007.08.008.
45. Kline KA, Kau AL, Chen SL, et al. Mechanism for sortase localization and the role of
2009;191(10):3237-3247. doi:10.1128/JB.01837-08.
202
46. Raz A, Fischetti VA. Sortase A localizes to distinct foci on the Streptococcus pyogenes
doi:10.1073/pnas.0808301105.
class of surface protein during Staphylococcus aureus pathogenesis. Proc Natl Acad Sci U
S A. 2002;99(4):2293-2298. doi:10.2307/3057958.
48. Race PR, Bentley ML, Melvin JA, et al. Crystal Structure of Streptococcus pyogenes
49. Melvin JA, Murphy CF, Dubois LG, Thompson JW, Moseley MA, McCafferty DG.
2011;50(35):7591-7599. doi:10.1021/bi200844h.
50. Prodromou C, Pearl LH. Recursive PCR: a novel technique for total gene synthesis.
optimizing the codon usage of DNA sequences. Nucleic Acids Res. 2007;35(Web Server
issue):W126-W131. doi:10.1093/nar/gkm219.
203
1995;6(3):277-293. http://www.ncbi.nlm.nih.gov/pubmed/8520220. Accessed February
13, 2015.
54. Goddard TD, Kneller DG. Sparky 3. Univ California, San Fr. 2006.
55. Schrödinger L. The PyMOL Molecular Graphics System, Version 1.3r1. August 2010.
56. Palmer AG, Kroenke CD, Loria JP. Nuclear magnetic resonance methods for quantifying
2015.
57. Jung ME, Clemens JJ, Suree N, et al. Synthesis of (2R,3S) 3-amino-4-mercapto-2-butanol,
a threonine analogue for covalent inhibition of sortases. Bioorg Med Chem Lett.
2005;15(22):5076-5079. doi:10.1016/j.bmcl.2005.07.073.
204