Download Two-step and one-step secretion mechanisms in Gram

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Cytokinesis wikipedia , lookup

Theories of general anaesthetic action wikipedia , lookup

SR protein wikipedia , lookup

Flagellum wikipedia , lookup

Protein wikipedia , lookup

Protein moonlighting wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Proteasome wikipedia , lookup

Signal transduction wikipedia , lookup

Intrinsically disordered proteins wikipedia , lookup

Cell membrane wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Magnesium transporter wikipedia , lookup

SNARE (protein) wikipedia , lookup

Thylakoid wikipedia , lookup

JADE1 wikipedia , lookup

Apoptosome wikipedia , lookup

Endomembrane system wikipedia , lookup

P-type ATPase wikipedia , lookup

List of types of proteins wikipedia , lookup

Western blot wikipedia , lookup

Type three secretion system wikipedia , lookup

Trimeric autotransporter adhesin wikipedia , lookup

Transcript
Biochem. J. (2010) 425, 475–488 (Printed in Great Britain)
475
doi:10.1042/BJ20091518
REVIEW ARTICLE
Two-step and one-step secretion mechanisms in Gram-negative bacteria:
contrasting the type IV secretion system and the chaperone-usher pathway
of pilus biogenesis
Ana Toste RÊGO1 , Vidya CHANDRAN1 and Gabriel WAKSMAN1,2
Institute of Structural and Molecular Biology, Birkbeck and University College London, Malet Street, London, WC1E 7HX, U.K.
Gram-negative bacteria have evolved diverse secretion
systems/machineries to translocate substrates across the cell
envelope. These various machineries fulfil a wide variety of
functions but are also essential for pathogenic bacteria to infect
human or plant cells. Secretion systems, of which there are
seven, utilize one of two secretion mechanisms: (i) the onestep mechanism, whereby substrates are translocated directly
from the bacterial cytoplasm to the extracellular medium or
into the eukaryotic target cell; (ii) the two-step mechanism,
whereby substrates are first translocated across the bacterial inner
membrane; once in the periplasm, substrates are targeted to one
of the secretion systems that mediate transport across the outer
membrane and released outside the bacterial cell. The present
review provides an example for each of these two classes of
secretion systems and contrasts the various solutions evolved to
secrete substrates.
INTRODUCTION
Two main mechanisms for transport operate in the various
secretion systems described above: one-step or Sec-independent
compared with two-step or Sec-dependent transport mechanisms.
The T2SS, T5SS and the CU systems lack an inner membrane
transporter that can transfer substrates across the inner membrane
in response to ATP. For the first step of transfer across the inner
membrane, these systems require either the general secretory
pathway, also called Sec translocon, or the Tat pathway in the case
of some T2SS substrates. Specific secretion system machineries
are then used for transport of substrates from the periplasm across
the outer membrane. These systems thus utilize a two-step or
Sec-dependent mechanism for translocation. On the other hand,
T1SS, T3SS, T4SS and T6SS all utilize a simultaneous onestep translocation of substrates across both membranes. These
are thus Sec-independent secretion pathways. The CU pathway
is a very simple two-step translocation pathway requiring just
two specialized proteins: the periplasmic chaperone and an outer
membrane usher [10,11]. The T4SS, on the other hand, is
a complex assembly of 12 or more proteins spanning across
both the membranes with cytoplasmic ATPases and inner and
outer membrane proteins assisting substrate secretion and pilus
biogenesis [12]. The CU pathway and T4SS are thus examples
of two types of secretion nanomachines, one utilizing a two-step
mechanism for secretion (the CU pathway) and the other a onestep one (T4SS). In the present review, we use these two systems to
compare and contrast the architectures and secretion mechanisms
of these two classes of secretion systems. Other secretion systems
could have been chosen for the same purpose, but describing them
all would have made for a large and unwieldy review, the length of
which was beyond the limits imposed for reviews published in this
journal. Instead, the authors have opted for a review comparing
Gram-negative bacteria have evolved several simple to complex
machineries for transport of substrates across their cell membrane
in response to various environmental cues. These machineries,
known as secretion systems, are essential for bacterial virulence
and are used by bacteria to acquire nutrients, transport various
proteins, nucleic acids or toxins or assemble cell-surface
organelles such as pili or fimbriae. Transport has to occur across
both the membranes with the periplasmic space enclosed within.
To achieve this, bacteria have evolved at least seven different
secretion systems classified as type I (T1SS) to type VI (T6SS)
secretion systems and the CU (chaperone-usher) system [1,2].
T1SS is a tripartite system made up of an inner membrane ABC
(ATP-binding cassette) transporter, an adaptor protein bridging
the two membrane components and an outer membrane pore [3].
The T2SS (type II secretion system) spans both the membranes but
require the Sec or Tat pathway for transporting substrates across
the inner membrane [4]. The T3SS (type III secretion system)
or injectisome forms large supramolecular structures spanning
both membranes and is structurally and evolutionarily related
to bacterial flagella [5]. The T4SS (type IV secretion system)
also forms large assemblies spanning both membranes and is
evolutionarily related to the conjugation machinery [6,7]. The
T5SS (type V secretion systems) are autotransporters or twopartner secretion systems assembled only in the outer membrane
[8]. The T6SS is a recently discovered multi-protein system
still largely uncharacterized [2]. Finally, the CU system is a
simple system assembled in the outer membrane and responsible
for assembly and secretion of virulence surface organelles
[9,10].
Key words: chaperone-usher, protein secretion, structural biology,
type IV secretion, virulence factor.
Abbreviations used: AAD, all-α-helical domain; CU, chaperone-usher; DSC, donor-strand complementation; DSE, donor-strand exchange; EM, electron
microscopy; I-layer, inner layer; NBD, nucleotide-binding domain; Nte, N-terminal extension; O-layer, outer layer; T1SS, T2SS etc., type I secretion system,
type II secretion system etc; TM, transmembrane.
1
These authors contributed equally to the present review.
2
To whom correspondence should be addressed (email [email protected]).
c The Authors Journal compilation c 2010 Biochemical Society
476
A.T. Rêgo, V. Chandran and G. Waksman
the one- and two-step secretion systems they study, i.e. the CU
pathway and the T4SS.
CU SECRETION SYSTEM
The CU system is responsible for the assembly and secretion
of pili (also called fimbria), multisubunit cell surface appendages
present in a wide range of pathogenic Gram-negative bacteria (e.g.
Escherichia coli, Salmonella enterica, Yersinia species, Klebsiella
pneumoniae, Pseudomonas aeruginosa etc.) [13]. CU pili are
virulence factors employed by bacteria for host recognition and
attachment, invasion and biofilm formation.
CU pili are polymers of protein subunits. Two specialized
proteins are required for CU pilus assembly: a periplasmic
chaperone that stabilizes pilus subunits and primes them for
subsequent polymerization into a pilus [14], and an outer
membrane usher that works as the assembly platform where
the chaperone subunit complexes are recruited, the subunits are
polymerized and translocated for presentation at the cell surface
[10,11].
CU pilus morphology varies from linear thick rods to more
flexible thin structures. Phylogenetic analysis of usher sequences
classified them into six major clades: α, β, γ (subdivided into γ 1 ,
γ 2 , γ 3 and γ 4 ), κ, π and σ fimbriae [15]. Rod-like fimbria, also
called ‘typical’ fimbria (found in α, γ and π clades) are one of the
best studied morphology types of the CU family, mainly through
extensive studies of P and Type 1 pili (π and γ 1 fimbrial clades
respectively) in UPEC (uropathogenic E. coli).
Architecture of P and Type 1 CU pili
CU fimbriae are typically encoded in individual gene clusters,
where all subunits in addition to the chaperone and usher are
encoded. The number of subunits varies from system to system.
For instance, the Type 1 pili cluster codes for four known subunits
(Fim A, F, G and H), whereas the P pili cluster codes for six known
subunits (Pap A, H, K, E, F and G). In both systems, these subunits
are arranged into two sub-assemblies: the pilus ‘rod’ and the ‘tip
fibrillum’ (Figure 1). The rod is a long thick rigid structure with
a ∼ 6.8 nm diameter made up of repeating subunits of PapA or
FimA (>1000 copies), in P and Type 1 pili respectively. These
repeating subunits form a right-handed helical structure with 3.3
subunits per turn [16,17]. The tip fibrillum of the P pilus is a thin
and flexible structure with a ∼ 2 nm diameter, made of repeating
PapE subunits (5–10 subunits) arranged in an open helical form
and connected to the rod by a copy of the adaptor subunit PapK
(Figure 1). A second single copy adaptor subunit PapF bridges
the distal end specialized adhesin subunit PapG with the PapE
polymer [18]. PapG has been shown to specifically recognize
the globo-series of glycolipids present in the human kidney, and
PapG has been shown to be a major virulence factor in kidney
infections [19,20]. In Type 1 pili, the linear tip fibrillum is simply
formed by one copy of each of the subunits FimF, FimG and
the adhesin FimH [21,22] (Figure 1B). For that reason, Type 1
pili form shorter tip fibres. FimH recognizes mannose residues
present in the human bladder [23] and Type 1 pili have been
shown to be responsible for bladder infections (cystisis), as well
as being implicated in attachment to abiotic surfaces in a model
biofilm system [24,25]. The P pilus rod is terminated at the base
by the PapH subunit that anchors the pilus structure to the cell
wall [26,27] (Figure 1A). No similar subunit has yet been found
in the Type 1 pili.
c The Authors Journal compilation c 2010 Biochemical Society
Figure 1
General diagram of the CU mechanism
(A) Diagram of the P pilus representing all the subunits involved at the outer membrane. Each
subunit is represented by oval shapes with a distinct colour and a letter; for example PapG
is labelled G, Pap F is labelled F, and so forth. (B) Diagram of Type 1 pili with the same
colour scheme as in (A). The pilus biogenesis is initiated by the export of the unfolded pilus
subunits to the periplasm via the Sec translocation machinery, schematically represented in the
inner membrane [155]. The periplasmic chaperone then captures subunits at the exit of the Sec
translocon and assists in their folding and targeting to the outer membrane dimeric usher where
one of the pores is activated. Once at the activated pore, subunits polymerize in an ordered
sequence until pili completion. The extracellular, outer membrane, periplasmic space, inner
membrane and cytoplasm are labelled E, OM, P, IM and C respectively.
Assembly of CU pilus
The initial step of the pilus biogenesis is the export of the pilus
subunits to the periplasm via the Sec translocation machinery
(Figure 1) (Sec-mediated transport through the inner membrane
is not within the scope of the present review. For excellent reviews
on this aspect, we refer the reader to [28,29]). At the exit of the Sec
machinery, each subunit interacts with a periplasmic chaperone
(PapD and FimC for P and Type 1 pilus subunits respectively)
for proper folding and guidance to the outer membrane usher.
In the absence of the chaperone, subunits aggregate and DegP
protease transcription is activated, promoting subunit degradation
[14,30,31].
Two-step versus one-step secretion systems
Figure 2
477
DSC and DSE
(A) DSC. Left panel: topology diagram of PapK (yellow) showing the PapD G1 (red) in the gap between strand A1 and F. Middle panel: structure of the PapD–PapK complex (PDB: 1PDK). PapD and
PapK are in ribbon and surface representations respectively, colour-coded red and yellow respectively. The location of the P1 to P4 pockets in the subunit’s groove as well as that of the conserved
chaperone residues co-ordinating the subunit’s C-terminus (Arg8 and Lys112 ) area are indicated. Right panel: close-up view of the P5 pocket of PapK. (B) DSE. Left panel: topology diagram of PapF
(in green) bound to PapD (in red). Middle panel: topology diagram of the same subunit (in green) bound to Nte of PapK (in yellow). Right panel: surface representation of PapF bound to the Nte of
PapK (PDB: 1N12). The location of the P2 to P5 pockets is shown. (C) The PapD–PapH structure (PDB: 2J2Z). Left panel: PapD and PapH are in surface and ribbon representations respectively,
colour-coded in red and grey respectively. The location of the P2 to P5 pockets is shown. Right panel: a close-up view illustrating the fact that PapH does not have a P5 pocket.
Mechanism of chaperone function: donor strand complementation
CU chaperones have a conserved boomerang-like structure,
composed of two Ig-like domains [32,33]. They possess two
conserved residues essential for forming complexes with the
subunits (Figure 2A). These residues are located between the two
domains of the chaperone. In the PapD chaperone, these residues
are Arg8 and Lys112 , and other chaperones have homologous
residues in the same position [33,34]. Mutations in these
residues abolishes pilus biogenesis [33]. Formation of chaperone–
subunit complexes is also dependent upon a series of alternating
hydrophobic surface residues (P1 to P4 residues) located in the
last β-strand of the chaperone’s N-terminal domain (strand G1 )
(Figure 2A).
Subunits share an incomplete Ig-like structure [35,36] with
only six of the seven characteristic β-strands (βA to βF) present
in the fold (Figures 2A and 2B). The C-terminal β-strand (βG)
is entirely missing. As a result, a hydrophobic groove in the
structure is created where the missing secondary structure should
be. Because of the missing strand, subunits are unstable and
unable to fold correctly on their own. Folding and stabilization
of subunits is afforded by the insertion of the chaperone’s G1
strand into the subunit’s hydrophobic groove and interaction of
the conserved arginine and lysine residues with the C-terminal
c The Authors Journal compilation c 2010 Biochemical Society
478
A.T. Rêgo, V. Chandran and G. Waksman
carboxylic group of the subunit’s strand F (Figure 2A). Insertion
of the G1 strand results in its P1 to P4 residues occupying
the P1 to P4 sites/pockets of the subunit’s groove (Figure 2A).
Thus the chaperone ‘donates’ in trans the missing secondary
structure: this mechanism of chaperone-mediated subunit folding
and stabilization is termed DSC (donor-strand complementation)
[14,30,35–37]. The resulting complemented fold, however, is not
a canonical Ig-fold, since the chaperone G1 strand runs parallel to
the subunits F strand instead of anti-parallel [38].
Mechanism of subunit polymerization: donor-strand exchange
Every pilus subunit possesses a 10–20 residue long disordered
Nte (N-terminal extension), which, similar to the G1 β-strand of
the chaperone, contains alternating hydrophobic residues. This
Nte was shown not to be part of the Ig-like fold [36]. During
the course of polymerization, the Nte of the incoming subunit
displaces the G1 donor β-strand of the chaperone, complementing
the Ig-like fold of the previous subunit assembled in the pili
[39] (Figure 2B). This mechanism of subunit polymerization
is called ‘donor-strand exchange’ (DSE) [35,36,39,40]. In the
subunit groove, the chaperone G1 strand hydrophobic residues
occupy the P1 to P4 sites/pockets, as mentioned above, whereas
after DSE the incoming Nte hydrophobic residues (termed P2
to P5 residues) occupy sites/pockets P2 to P5 of the groove
(Figures 2A and 2B). The subunit fold is now a canonical Igfold, since the Nte strand runs anti-parallel to the subunit’s F
strand, thus in a more energetically favourable conformation.
The reaction by which the chaperone G1 strand is substituted
by the incoming subunit Nte in the groove proceeds via a zipin-zip-out mechanism, whereby the process of DSE is initiated
by the insertion of the P5 residue of the incoming subunit’s Nte
into the P5 pocket of the previously assembled subunit groove
[41,42] (see Figure 2A, right panel for the P5 pocket). DSE
progresses with the new Nte gradually displacing the chaperone’s
G1 strand step by step from P5 to P1 [42] (Figure 2B). The
receiving subunit structure then transitions from a chaperonestabilized semi-unfolded state of high-energy to a folded state of
lower energy when in interaction with an other subunit [43]: it
is believed that it is the folding energy released upon DSE that
drives subunit polymerization [43,44]. No other external source
of energy is required to drive pilus polymerization. Termination
of pilus biogenesis happens upon incorporation of the subunit
PapH in the growing pilus. This happens because PapH is unable
to undergo DSE since it lacks a P5 pocket [27] (Figure 2C). No
mechanism for type 1 pilus termination has been found yet as a
PapH-like termination subunit is unknown in that pilus system.
The usher assembly platform
In vivo DSE occurs at the outer membrane usher [44]. DSE is
observed in vitro and in the absence of the usher, but proceeds
only at very slow rates [45]. Rapid fibre formation, even in vitro,
requires the presence of the usher. Thus the usher acts as a catalyst
of DSE [46].
Outer membrane ushers are ∼ 90 kDa proteins with four distinct
functional domains: an N-terminal periplasmic domain with
∼ 125 residues, a large translocation pore domain with ∼ 500
residues that is interrupted by a soluble plug domain of ∼ 110
residues and a C-terminal periplasmic domain with ∼ 170 residues
[47,48] (Figure 3A).
The N-terminal domain binds chaperone–subunit complexes
[49,50] with binding affinities for each of the complexes
decreasing in an order reflecting the natural order of subunits
c The Authors Journal compilation c 2010 Biochemical Society
in the pilus, a process that is believed to be responsible, in
part, for subunit ordering. Structural and biophysical studies
of the N-terminal domain of the FimD usher bound to both
FimC–FimH and FimC–FimF showed that a flexible tail at the
N-terminus (residues 1–24) was responsible for 40%–50 % of
the binding interface and that tail was directly interacting with
the subunit (FimH or FimF) (Figure 3B). Deletion mutants of
residues 1–24 abolished chaperone–subunit interaction with the
usher [49,51]. Similarly, for the PapC N-terminal domain, deletion
of residues 1–11 abolished targeting of the usher to chaperone–
subunit complexes [50]. Thus the tail of the N-terminal domain
seems to be responsible for subunit-specific interactions. In vitro,
the N-terminal domain doesn’t seem to be fundamental for pilus
biogenesis, which may suggest that its role in vivo may be to
capture chaperone–subunit complexes and increase their local
concentration to accelerate pilus assembly [53].
The usher translocation pore forms an outer membrane
β-barrel containing 24 TM (transmembrane) β-strands [52,53]
(Figure 3C). The pore has an inner diameter of 45 Å×25 Å
(1 Å=0.1 nm), a size compatible with the passage of folded
subunits. Between strands 6 and 7 lays the plug domain, a βsandwich fold that, in an inactive usher, completely occludes the
pore, preventing the passage of the subunits. Deletion of this
domain completely abolishes the assembly of pili in vitro [53],
thus the plug domain participates actively in pilus biogenesis.
The C-terminal domain role is still not completely understood,
but C-terminal truncation and deletion mutants of the P pilus
usher PapC demonstrated that the C-terminus is fundamental
for proper binding of chaperone–subunit complexes and pilus
assembly [53,54]. Differences in protease susceptibility in the
Type 1 pilus usher FimD show that following targeting to the usher
N-terminus, the N-terminal domain of FimH (from the
FimC–FimH complex) forms stable interactions with the FimD
C-terminus, inducing a conformational change in FimD that may
be fundamental in the activation step of pilus biogenesis [55–
57]. Such conformational change must include the movement of
the plug either inwards within the translocation pore or outwards
in the periplasm [52]. Binding of the chaperone-adhesin (PapDPapG) with the usher C-terminus was also demonstrated in vitro.
Complementation studies on PapC C-terminal truncates have
also shown that, in the absence of the C-terminal domain, the
translocation pore is unable to become activated. These results
suggest an important role of the C-terminal domain in the first
step of pilus biogenesis, the activation of the usher [54].
The P pilus usher PapC assembles into ring-shaped homodimeric functional units in vivo. This was demonstrated
by cryo-EM (where EM is electron microscopy) images of
two-dimensional crystals where PapC was reconstituted in
E. coli lipids and in complementation studies of PapC lossof-function mutants [54,58]. Dimer formation doesn’t seem
to be necessary for fibre formation in vitro [53]. Cryo-EM
reconstruction of FimD–FimH–FimG–FimF–FimC shows
that only one of the ushers is activated and able to secrete
subunits (Figure 3D) [52]. In fact, the requirement for a
second usher in vivo is not completely understood. Remaut
et al. [52] proposed a model where the two usher N-terminal
domains alternate in the recruitment of new chaperone–subunit
complexes. The structure of the C-terminal domain and of
usher–chaperone–subunit complexes will provide answers to the
mechanism of usher activation and assembly mechanism.
TYPE IV SECRETION SYSTEM
T4SSs are specialized assemblies utilized by both Grampositive and Gram-negative bacteria for one-step transport of
Two-step versus one-step secretion systems
Figure 3
479
The usher PapC structure
(A) Domain organization of PapC. The translocation domain and the N-terminal domains (Ntd) are represented in dark blue, the plug domain in magenta, and the C-terminal domain (Ctd) in white.
Residue boundaries are indicated. (B) Structure of the FimD N-terminal domain (dark blue) bound to the FimC–FimHp (red and light green respectively) complex (PDB: 1ZE3). FimHp is the pilin
domain of FimH. (C) Structure of the translocation domain of the PapC usher (PDB: 2VQI). The domain is in ribbon representation colour-coded in dark blue for the barrel, yellow for the single helix
present in the structure, orange for the trigger hairpin, and magenta for the plug domain. Upper panel: side views. Lower panels: top views. (D) Type 1 pili tip captured by cryo-EM. X-ray structures
from FimH (light green), FimD (dark blue), FimG (orange), FimC (red) and FimF (yellow) were docked [52]. X-ray crystallography structures used were FimH (PDB: 1QUN), FimD (PDB: 2VQI),
FimG (PDB: 3BFQ), FimC–FimF–FimDN (PDB: 3BWU) and the cryo-EM map of the FimD2:C:F:G:H type 1 pilus tip assembly intermediate is from emd_5009 (EM Data Bank). A three-dimensional
interactive Figure of this structure is available at http://www.BiochemJ.org/bj/425/0475/bj4250475add.htm.
protein, nucleic acid or nucleoprotein substrates across cell
membranes [6,7,59,60]. They are evolutionarily related to
bacterial conjugation systems. T4SSs are implicated in the
transport of virulent proteins or DNAs in various plant, animal or
human pathogens [61,62], in bacterial conjugation and in DNA
uptake or release into the extracellular space [63,64]. T4SSs
facilitate the exchange of genetic material, promoting genetic
plasticity and adaptive response of bacteria to the environment.
Conjugative transfer of antibiotic-resistance genes via T4SSs is a
leading cause for emergence of multidrug resistance [65].
T4SSs are formed by at least 12 proteins, termed VirB1–
VirB11 and VirD4 according to the prototypical T4SS from
Agrobacterium tumefaciens, although some T4SSs are composed
of just a subset of these proteins [60,66]. The VirB/VirD4
proteins can be grouped based on their function and location
into energizing proteins, core translocation pore complex, pilusassociated proteins and other T4SS proteins. Energizing proteins
include the ATPases, VirB4, VirB11 and VirD4. VirB6 to VirB10
proteins form the core translocation complex. VirB6, VirB8
and VirB10 are inner membrane proteins, whereas lipoproteins
VirB7 and VirB9 were thought to form the outer membrane
pore. Most T4SSs form an extracellular pilus formed by the
major VirB2 and minor VirB5 proteins. The other T4SS proteins
include the putative lytic transglycosylase VirB1 and membrane
protein VirB3, the location and function of which are still either
ambiguous (location) or unknown (function).
Structure and role of the various subunits
The VirB4, VirB11 and VirD4 ATPases
The three ATPases VirD4, VirB4 and VirB11 power T4SSs. They
are located in the cytoplasm, sometimes inserted in the inner
membrane, and form a network of interactions among themselves
and the rest of the assembly [67].
c The Authors Journal compilation c 2010 Biochemical Society
480
Figure 4
A.T. Rêgo, V. Chandran and G. Waksman
Atomic structures of known T4SS components
Top panels, the crystal structures in cartoon representation for the E. coli VirB5 homologue TraC (PDB: 1R8I), B. suis VirB8 (PDB: 2BHM), H. pylori VirB10 homologue ComB10 (PDB: 2BHV) and the
NMR structure of E. coli VirB9/TraO C-terminal domain in complex with VirB7/TraN (PDB: 2OFQ). The lower panels show in cartoon representation, the oligomeric and monomeric crystal structures
of the H. pylori VirB11 homologue HP0525 (PDB: 1NLZ) and the E. coli VirD4 homologue TrwB from plasmid R388 (PDB: 1E9S). CTD, C-terminal domain; NTD, N-terminal domain.
VirD4
VirD4 proteins are the cytoplasmic gatekeepers that recruit T4SS
substrates to the T4SS channel. VirD4s are inner membrane
proteins with an N-terminal TM domain and a large cytoplasmic
domain [68]. TrwB, encoded by the IncW conjugative plasmid
R388, has been the most characterized VirD4 homologue. The
TM domain of TrwB plays a role in stability, substrate selectivity
and oligomerization of the protein [69,70]. The crystal structure
of the cytoplasmic soluble domain of TrwB [71] revealed a large
110 Å × 90 Å globular hexameric assembly with a 20 Å channel
(Figure 4). This channel narrows down to an ∼ 8 Å constriction
at the cytoplasmic end. Each subunit is shaped like an orange
slice and made up of two distinct domains, an NBD (nucleotidebinding domain) and an AAD (all-α-helical domain). The NBD
domain is a mixed α/β RecA-like domain with a central highly
twisted nine-stranded β-sheet flanked by α-helices on both sides
and orientated towards the inner membrane. Both the Walker A
and B motifs for NTP-binding are essential for secretion [72]. A
DNA-dependent ATPase activity has been demonstrated for TrwB
[73]. No helicase activity has been observed so far for VirD4s.
The AAD domain, consisting of seven α-helices, is orientated
towards the cytoplasm. A DNA-binding role for this domain has
been suggested, based on structural homology.
VirB11
VirB11 belongs to a large family of traffic ATPases involved
in various secretion systems such as type II, III, IV and VI
[66,74,75]. These NTPases are important for their ATP-bindingdependent substrate selection and translocation, and for pilus
production [76]. VirB11-like proteins exhibit distinct substrate
c The Authors Journal compilation c 2010 Biochemical Society
specificity and their NTPase activity is enhanced by the presence
of detergent/phospholipids in vitro [77,78]. VirB11-like proteins
have indeed been shown to be peripheral membrane proteins associated with the inner membrane [79]. Crystal structures have been
solved for the Helicobacter pylori VirB11 homologue HP0525
in its apo and substrate-bound forms, and for the Brucella suis
VirB11 protein [80–82] (Figure 4). Each HP0525 monomer is
made up of two domains: an N-terminal domain with a novel fold
and a C-terminal domain with the typical RecA fold. The two
domains form hexameric rings mounted on each other shaped like
a six-clawed grapple. The N-terminal domain consists of a central
six-stranded β-sheet packed against two α-helices on one side,
whereas the C-terminal domain is made up of a seven-stranded
β-sheet sandwiched between three and four α-helices on either
side. The interface between the two domains forms the NTPbinding site. VirB11 from B. suis shows a very similar double
hexameric ring structure with the individual domains in each
monomer structurally similar to those of HP0525 [82]. However,
a large domain swap is noticed that leaves the overall hexamer
intact but alters both the substrate-binding site and inter-subunit
interfaces.
The apo and substrate-bound structures of HP0525 have
given valuable structural insight into the function of VirB11like ATPases [80,81,83]. The apo-form of HP0525 shows an
asymmetric hexamer due to rigid body movements in the NTD
ring of the assembly. The conformational flexibility has been
suggested to help further interactions with the rest of the T4SS
assembly. ATP binding takes place initially at three catalytic sites,
giving rise to a more rigid conformation to the entire assembly.
ATP hydrolysis at these sites followed by ATP binding to the
remaining three sites locks the entire assembly into a perfectly
Two-step versus one-step secretion systems
symmetrical hexamer. Further hydrolysis and release of ADP from
all sites returns the molecule to its initial asymmetric apo-form.
VirB4
VirB4-like proteins are the largest ATPases in T4SSs and are
evolutionarily conserved in both Gram-negative and Grampositive bacteria [6,84]. Bioinformatic studies suggest four conserved motifs, including the Walker A and B motifs and a
conserved domain for the VirB4 family of proteins [85]. The NTPbinding motifs are essential for the functionality of T4SSs [86].
However, unambiguous demonstration of the ATPase activity for
VirB4 homologues has been difficult, with two positive reports
for A. tumefaciens VirB4 and more recently TrwK from plasmid
R388 [87,88].
No structural detail is available for VirB4-like proteins and
there is ambiguity regarding the exact cellular location and oligomerization state. Topological studies on A. tumefaciens VirB4 map
it as a polytopic cytoplasmic membrane protein with probably two
periplasmic and three cytoplasmic domains [89]. TrwK, the VirB4
homologue from plasmid R388, however, is a soluble protein [88].
A hexameric model of the C-terminal domain of VirB4 has been
proposed based on its relationship to the cytoplasmic domain
of VirD4 homologue TrwB by divergent evolution [90]. This
initial model suggested a cytoplasmic location for VirB4, although
subsequent work places the C-terminal domain predominantly in
the periplasm [91].
Core channel components
VirB6
VirB6 is a highly hydrophobic inner membrane channel subunit
essential for T4SS. No structural information is available to
date for this T4SS component. Topological studies on both
A. tumefaciens VirB6 and H. pylori ComB6 have mapped an
N-terminal periplasmic domain, five TM segments and a cytoplasmic C-terminal domain [92,93]. A large periplasmic loop (P2
loop) in the centre of the protein between TM2 and TM3 presumably contributes to the inner walls of the transport channel. The
C-terminal TM segments orchestrate substrate transfer to the other
inner membrane component VirB8, whereas both terminal regions
co-ordinate with the outer membrane and pilin subunit VirB9 and
VirB2. The physiological oligomeric state of VirB6 itself is not
known.
VirB8
VirB8 is a bitopic inner membrane protein with a short
cytoplasmic tail, a single TM helix and a large periplasmic
domain [68]. VirB8 is thought to serve as the nucleation centre
for the assembly of T4SSs [60,94]. Immunofluorescence and
immunoelectron microscopic methods showed unique clustering
of VirB8 gold particles independent of other VirB proteins
and also the influence of VirB8 on VirB9–VirB10 localization
[94]. VirB8 participates in a rich network of protein–protein
interactions in the T4SS with VirB1, VirB4, VirB5, VirB9, VirB10
and VirB11, and is in direct contact with the substrate during
transport [94–100].
High-resolution structural information is available for the
periplasmic domains of two homologues, A. tumefaciens and
B. suis VirB8 (Figure 4) [101,102]. Both crystal structures reveal
a dimer of VirB8 with an extended four-stranded antiparallel
β-sheet juxtaposed by five α-helices. The periplasmic domains
have a fold similar to nuclear transport factor NTF2. The crystal
481
structures reveal a deep groove lined by highly conserved residues,
a structural feature absent in NTF2 and other similar fold proteins
and is probably a hot spot for protein–protein interactions.
Mutational analysis of VirB8 identified a β-sheet region adjacent
to this groove to be important for interactions with VirB4 and
VirB10 [103]. The physiological oligomeric state of this protein
in the T4SS assembly is still unknown.
VirB10
VirB10 is also a bitopic inner membrane protein with a similar
topology as VirB8 with a short N-terminal cytoplasmic tail,
a single predicted TM helix and a large periplasmic domain
consisting of a proline-rich region and a β-barrel domain [68,104].
VirB10, although inserted in the inner membrane, has also been
observed associated with the outer membrane [94]. Interactions
of VirB10 with VirB9 [100] and the energizing proteins VirB4,
VirB11 and VirD4 [67,105–108] substantiated by mutational
analysis have established VirB10 as the structural scaffold of
T4SSs bridging the two membranes together [104].
A high-resolution crystal structure is available for part of the
periplasmic domain of H. pylori ComB10 protein [102] (Figure 4).
This C-terminal periplasmic domain consists of an atypical
β-barrel flanked by an α-helix on one side and an α-helical
antennae-like projection at the top. The altered β-barrel is splayed
open on one side. A head to tail dimer is observed in the crystal
structure, which is probably a crystal-packing artefact. Mutational
studies have highlighted the role of the side helix and α-helical
antennae in uncoupling pilus biogenesis and substrate transfer
functions of T4SSs [104]. See Note added in proof.
VirB10 shares features with the inner membrane energy sensor
protein TonB: both are bitopic membrane proteins with a prolinerich region and conserved hydrophobic or coiled-coil regions
important for protein–protein interaction and function. Similar to
TonB, VirB10 probably couples inner membrane energy status
to the outer membrane gating required for transport. Indeed,
VirB10 responds to the cellular ATP levels with different protease
susceptibility, suggesting different structural conformations. The
structural transition of VirB10 in response to ATP levels also
affects the VirB7–VirB9 heterodimer and VirB7–VirB9–VirB10
complex formation [109].
VirB7 and VirB9
VirB7 is a monotopic lipoprotein [110]. The modified N-terminus
is anchored in the outer membrane and the C-terminus is
periplasmic. In A. tumefaciens, VirB7 stabilizes VirB9 by forming
a disulfide link between the two proteins [111]. VirB9 is mainly
a periplasmic protein that has been suggested to form the outer
membrane pore, although no TM segments have been predicted.
Mutational studies on A. tumefaciens VirB9 has mapped the
N-terminal domain to be important for substrate selection and
the C-terminal domain for stabilizing the VirB10–VirB9–VirB2
proteins that form the distal end of the channel and are essential
for secretion [112]. The NMR structure has been solved for
the interacting domains, for example the C-terminal domain
of the VirB9 homologue encoded by the E. coli plasmid pKM101,
TraO (TraO-CT), bound to the VirB7 homologue TraN encoded
by the same plasmid (Figure 4) [113]. The structure reveals a
β-sandwich fold for TraO-CT, around which TraN wraps. Two
edge strands in TraO that diverge away from the β-sandwich
fold form an appendage which has been shown to be surface
exposed in A. tumefaciens, suggesting a role in outer membrane
pore formation. See Note added in proof.
c The Authors Journal compilation c 2010 Biochemical Society
482
A.T. Rêgo, V. Chandran and G. Waksman
Pilus components
The function of the type IV secretion pilus (not to be mistaken
for the type IV pilus, the assembly of which is a T2SSrelated process) is still unclear. It might be used as conduit
for substrates [13,114] or adhesion [115,116]. The T pilus of
A. tumefaciens consists of a processed form of VirB2, the major
component of pilin, and VirB5, the minor component [117,118].
Pilus biogenesis requires other T4SS components and several
uncoupling mutations have been mapped that allow secretion but
not pilus assembly [76,104,112]. VirB2 is highly processed: the
N-terminal leader signal peptide that targets the pilus subunit
across the inner membrane is cleaved by specific peptidases and
further processing might occur, sometimes including cyclization
[119–122]. No structural information for VirB2-like proteins
alone is available. However, a crystal structure is available for
the VirB5 homologue in pKM101, TraC (Figure 4C) [123].
The structure reveals a three-helix bundle capped by a globular
appendage. Structure-based mutagenesis studies suggested a role
of VirB5 in adhesion and host recognition [123]. Recently,
the F pili structure has been determined using cryo-EM and
single-particle reconstruction [124]. Two pilin subunits pack with
two different symmetries that co-exist; a C4 symmetry forming
stacked rings axially separated by ∼ 12.8 Å and a one-start helical
symmetry with an axial rise of 3.5 Å per subunit and a pitch of
∼ 12.2 Å. The central lumen seems large enough for transferring
single-stranded DNA but not folded proteins.
Other T4SS-associated proteins
VirB1
VirB1, although non-essential for secretion, is essential for
pilus biogenesis [125–127]. Deletions of the virB1 gene lead to
attenuated virulence with respect to T4SS functions [99]. An
exception in this case is that of H. pylori, where the VirB1
homologue, HP0523, is essential to bacterial virulence [128].
VirB1 is a periplasmic protein with a conserved lysozyme-like
transglycosylase domain [129]. The catalytic activity of the transglycosylase domain to lyse the peptidoglycan cell layer probably
helps to accommodate the large T4SS assembly in the periplasmic
space. VirB1 is processed [130] to release a C-terminal third of
the protein VirB1* which is required for pilus biogenesis and is
secreted to the extracellular space [127]. Two hybrid experiments
and later biochemical experiments have shown interaction of
VirB1 with other T4SS components VirB9, VirB8 and VirB11
suggesting more (as yet undefined) roles for VirB1 in T4SS
assembly and function [99].
VirB3
VirB3 is one of the most ill understood proteins in the T4SS
assembly both in terms of localization and function. It is
a membrane protein with TM segments predicted near the
N-terminus. Localization of this protein is very uncertain, with
reports suggesting either inner or outer membrane localization
[131,132]. Chimaeric VirB3/VirB4 proteins previously identified
in certain T4SS assemblies such as Campylobacter jejuni [133]
suggest an inner membrane localization to be more probable due
to association with VirB4, which is associated with the inner
membrane. However, no functional role has been identified so far
for this protein.
Assembly of T4SSs
The biogenesis of the T4SS assembly has been proposed to start
with the formation of a core complex (VirB6–VirB10) (Figure 5),
c The Authors Journal compilation c 2010 Biochemical Society
followed by recruitment of the ATPases (VirD4, VirB4 and
VirB11) and/or pilus biogenesis (VirB2 and VirB5). Secretion
and pilus biogenesis might require two different complexes [6,7],
and thus, after assembly of the core complex, there might be
a bifurcation in the assembly process leading to formation of
two different complexes, one used for secretion, the other used
for pilus biogenesis. Indeed, mutations in some VirB proteins
decouple secretion from pilus biogenesis [76,92,104,112]. Also,
VirD4 is only used in secretion, whereas VirB11 is only used
in pilus biogenesis [127,134]. However, at this stage, the exact
composition of these complexes is not known.
A recent breakthrough in the field has been the 15 Å cryoEM structure of the core translocation channel formed by the
VirB7, VirB9 and VirB10 homologues of the E. coli pKM101
plasmid (TraN–TraO–TraF) (Figure 5) [135]. This provided the
first snapshot of a T4SS assembly. The core complex is a 1.05 MDa
structure with 14-fold symmetry and traverses across both inner
and outer membranes. The nucleation factor VirB8/TraE, although
co-expressed with the other proteins, is not required for forming
the core complex. The lipidation of lipoprotein VirB7/TraN is
also not essential for core complex formation, but required for
targeting the complex to the outer membrane.
The structure shows multilayered rings consisting of 14 copies
each of TraN/VirB7, TraO/VirB9 and TraF/VirB10, with a height
and diameter of 185 Å. Two distinct layers are seen, termed the I
(inner)- and O (outer)-layers with respect to location in the membrane (Figure 5). The two layers are partly or fully double walled.
The O-layer consists of a cap and a main body. The cap, which
probably traverses the outer membrane, encloses a 20 Å pore that
narrows to 10 Å at the junction with the main body. The main
body has a larger internal chamber that is 110 Å wide narrowing
down to 55 Å at the base. The I-layer, that resembles a cup, has
a partial double wall that merges down at the base and encloses a
chamber similar to the O-layer. A total of 14 finger-like densities
extend below the I-layer. Using various biochemical studies, the
C-terminal domains of both VirB9 and VirB10 have been located
in the O-layer and the N-terminal domains of both proteins in the
I-layer. The finger-like protrusions probably are the TM segments
of TraF inserted in the inner membrane.
T4SS mechanism
One of the most valuable contributions to the T4SS field has
been the TrIP (transfer DNA immunoprecipitation) study on the
A. tumefaciens T4SS, delineating the substrate transfer route
[126]. In A. tumefaciens, the substrate is the T-DNA (transferred
DNA). It successively contacts VirD4, VirB11, VirB6, VirB8,
VirB2 and VirB9 during transfer. Other T4SS components, VirB3,
VirB4, VirB5, VirB7 and VirB10, assist the transfer indirectly,
modulating/regulating differentially each of the transfer steps. The
coupling protein VirD4 contacts the substrate first, independently
of ATP hydrolysis [67]. The substrate is then passed on to VirB11,
independently of any energy requirements. The lipoprotein VirB7
is important for this initial step of transfer, highlighting its
importance for mediating the formation of the stable core
assembly. The substrate then comes in to contact with the inner
membrane proteins VirB6 and VirB8. This early transfer depends
on the energetic components VirD4, VirB4 and VirB11, and ATP
hydrolysis is required at this stage. This seems to suggest that
VirB6 and VirB8 function together and probably line the inner
surface of the VirB7–VirB9–VirB10 core complex to contact the
substrate directly. The substrate is then transferred to the outermembrane-associated proteins VirB9 and the major pilin subunit
VirB2. VirB3, VirB5 and VirB10 are important at this stage of
transfer. The importance of VirB10 is evident from the fact that it
Two-step versus one-step secretion systems
Figure 5
483
A model for Type IV secretion systems
A model for the assembly of T4SS components VirB1–VirB11/VirD4 is shown. A cut-away view of the cryo-EM structure of the core translocation assembly formed by VirB7/VirB9/VirB10 homologues
in E. coli pKM101 is shown in grey crossing both the inner and outer membranes. A surface representation of the atomic structures of T4SS components (VirB11 homologue, PDB entry 1NLZ; VirD4
homologue, PDB entry 1E9S; VirB5 homologue, 1R8I; and VirB8, PDB entry 2BHM) that are available are shown. The T4SS components for which no known structural information is available are
shown in distinct shapes and colours. The three ATPases powering the T4SS on the cytoplasmic side of the membrane are shown. Inner membrane components VirB6 and VirB8 are shown lining the
inner surface of the core complex. Pilus-associated proteins VirB2 and VirB5 and other proteins VirB1 and VirB3 are shown. All components are labelled according to their A. tumefaciens VirB/VirD4
nomenclature.
bridges the two membrane components together (Figure 5) [104]
and plays the role of an energy sensor for bringing about the
required conformational changes important for substrate transfer
[109]. VirB5 is a minor pilin subunit probably assisting the
passage of substrate via the pilus. Thus the entire pathway for
substrate translocation has been defined and, except for VirB1, all
proteins play an important role for secretion.
CONTRASTING THE TWO SYSTEMS
Simplicity compared with complexity in the translocation of
substrates
Secretion systems range from relatively simple structures, such as
the CU secretion system composed of two subunits (the chaperone
and the usher) specialized in assembling and driving the secretion
of pilus subunits, to more complex machineries such as T4SS,
which is composed of at least 12 subunits able to mediate secretion
of a wide variety of substrates. T4SS proteins span both the inner
and outer membranes, forming MDa multicomplex machineries,
whereas the CU system proteins only form complexes with the
chaperone when they are recruited, and then with the usher and
each other to form the pilus structure. The translocation pore
in the CU system is composed simply of a ∼ 180 kDa outer
membrane dimeric usher [52] that works as an assembly platform.
The pore diameter of one of the PapC monomers (45 Å × 25 Å)
is large enough to allow folded subunits translocation (ranging
from 27 Å × 22 Å to 32 Å × 23 Å in diameter for PapA and PapK
respectively) upon movement of the plug domain. How the plug
is let go is not known, but simple mechanical processes might
operate such a release, probably mediated by the interaction of
the first chaperone–subunit to be presented to the usher (the chaperone–adhesin complex) with the C-terminal domain of the
usher. In contrast, the translocation pore complex in T4SSs, in
its minimal form, involves at least three proteins (VirB7, VirB9
and VirB10) with 14 copies of each assembling into a 1.05 MDa
channel complex [135]. At least on one side (the O-layer), the
channel is closed and, thus, it must open to let substrates through.
This necessitates complex multi-proteins sensing and mechanical
devices to operate the conformational changes required to drive
secretion of substrates, probably mediated by VirB10, but also
other VirB proteins not contacting the substrate.
ATP dependency/independency
Protein secretion across the outer membrane is dependent on
an energy source. However, the most commonly used energy
store in living organisms, ATP, is not present in the bacterial
c The Authors Journal compilation c 2010 Biochemical Society
484
A.T. Rêgo, V. Chandran and G. Waksman
periplasm. In the T4SS several cytoplasmic components are
assembled in a large cytoplasmic/inner membrane ATPase
complex (VirD4, VirB4 and VirB11) that utilizes ATP to operate
the conformational changes required to drive secretion across
both the inner and outer membranes. In the CU system no
cytoplasmic or inner membrane components are encoded and no
ATP is used. In fact, no external source of energy is required for
DSE and consecutive pilus formation [44]. Instead, the chaperone
stabilizes a high-energy conformation of the bound subunit, which
upon DSE is released, resulting in an exceptionally stable and
closed subunit conformation. The free energy released during the
process is believed to be sufficient to drive pilus formation [43].
It should not be inferred from the comparison above that,
in all two-step secretion systems, transport through the outer
membrane is ATP-independent. For example, in type II secretion,
although substrates are dependent on Sec or Tat (and thus ATP) for
transport through the inner membrane, transport through the outer
membrane also requires ATP. Although this is not completely
understood, it has been proposed that passage through the outer
membrane pore (termed the ‘secretin’) is powered by pilus growth
and retraction, a process dependent on inner membrane ATPases
and consumption of ATP. Thus, in T2SS, an active ATP-dependent
mechanism utilizing the pilus as a piston might mechanically push
substrates through the outer membrane secretin [4]. Conversely,
not all one-step secretion systems are dependent on ATP. For
example, although one-step transport by some T1SS are powered
by ATP, other T1SSs might use the inner membrane proton
gradient as a source of energy [136,137]. The present review does
not have the space to expand the description of these systems, but it
must be realized that, although common features are clearly shared
among secretion systems, each also exhibit unique characteristics
[138].
Chaperone presence and dependency
Molecular chaperones have been characterized in all living
cells as accessory proteins in various cellular functions. These
include de novo protein folding, stabilization of proteins under
stress conditions and maintenance of a loosely folded state for
translocation across membranes. Cytoplasmic chaperones are
involved in different types of protein secretion. They have been
best characterized in T3SSs where they are thought to maintain
the stability and secretion competence of effector proteins, but
have also been found in the T4SS. For example, the VirE1 protein
of A. tumefaciens is necessary for the translocation of the effector
protein VirE2 into plant cells [139–143]. Other ‘type IV adaptor
proteins’ required for translocation of a subset of substrates, such
as IcmS, IcmW and LvgA proteins of the L. pneumophila Icm/Dot
T4SS [144–146] may have a similar function. CagF is another
candidate for secretion chaperone-like molecule (present in both
soluble and membrane-associated milieu) of the CagA protein
[147].
In the periplasm, chaperone function is distinct from the
cytoplasmic chaperones and independent of any known source
of cellular energy. PapD-like periplasmic chaperones are small
monomeric proteins that transiently contribute steric information
to pilus subunits to facilitate their folding, maintain them in a
high-energy state, and prime them for future assembly. In the CU
secretion system, the role and mechanism of the chaperone is
very well understood and a single chaperone seems to be enough
to drive all of the different subunits to the assembly platform.
On the other hand, in T4SSs and other systems, chaperones are
still not very well characterized and their function is still poorly
understood.
c The Authors Journal compilation c 2010 Biochemical Society
Type IV secretion pilus compared with CU pilus
T4SS pili are varied in shape and size. The conjugation pili
in T4SS systems are classified based on morphology into two
classes: the long flexible F-type pili and the short rigid P-type
pili [148]. The F-type T4SS pili measure 2–20 μm in length
and 8–9 nm in diameter with a 2 nm central lumen [149,150].
The P-type pili are shorter than 1 μm in length and 8–12 nm
in diameter [149,150]. The overall structure of a F-type pilus
has been determined (see above) and the subunits appear to
stack in a superhelical arrangement. On the other hand, the H.
pylori T4SS system encoded by the Cag pathogenicity island
produces distinctive, large sheathed pili, ∼ 100–200 nm long and
∼ 70 nm wide with a ∼ 45 nm central needle made up of a
VirB10 homologue (HP0527/CagY) [115,128,151]. It is believed
that T4SS pili essentially play an adhesion role. For example,
CagL, possibly the VirB5 homologue of the Cag-encoded T4SS,
is targeted to the pilus surface and appears to play a role in host
recognition [115]. However, the role of T4SS pili as adhesive
devices has not been as clearly demonstrated as in the CU pili
systems, notably the P and Type I CU pili systems. Thus it remains
unclear whether T4SS pili are used for adhesion or as conduit for
substrate secretion or both.
CU systems also encode a variety of pilus types ranging from
the rigid rod-like pili, the detailed description of which is provided
above (P and Type 1 CU pili), but also very thin fimbriae or
filaments forming mingled masses at the surface of the bacteria
that harbour them [40,42,152,153]. In P and Type 1 pili, only one
protein is responsible for adhesion and host recognition (PapG
and FimH respectively), and this protein locates at the pilus tip
due to the unique properties of the chaperone–adhesin complex
to activate the usher. However, in some other CU pili systems
where only one or two pilus subunits have been identified, these
are likely to combine the functions of structural and adhesive
subunits, i.e. they assemble to form a pilus or fimbria but also act
as adhesins. In that case, the adhesive property of the pilus would
no longer be confined to the tip but thought to be spread out over
the entire length of the pilus/fimbriae [154].
PROSPECTS
We have used two different secretion systems, the CU pilus
system and the T4SS, to contrast two classes of secretion systems,
those using two distinct machineries to release substrates first
through the inner membrane and next through the outer membrane
(CU pili), and those that elaborate a complex double membranespanning machinery to cross both membranes (T4SSs) in a single
step. In the two-step systems, the Sec translocon is used to
cross the inner membrane. No details on the Sec machinery
is provided in the present review as it has been excellently
documented elsewhere [28,29]. Subsequent transport through the
outer membrane involves two specialized proteins, a chaperone
and an usher. Both proteins are used to drive a process of
assembly and secretion, which is made energetically favourable
by the exquisite modulation of protein folding by the chaperone.
By maintaining the subunits in a semi-unfolded conformation,
the chaperone stores up energy which, when released upon DSE,
drives the pilus biogenesis process forward. Processes involving
folding energy to drive secretion are also at play in Type V
secretion, a secretion process through the outer membrane, which
requires a single transporter polypeptide forming a β-barrel in
the outer membrane. Type V secretion substrates, often (but not
always) are part of the same polypeptide that forms the β-barrel,
are driven through the barrel in an unfolded state, and folding
Two-step versus one-step secretion systems
post-transport is thought to drive the extrusion of the unfolded
peptide.
Because T4SSs are double membrane spanning, they have
access to the store of ATP in the cytoplasm and thus the
T4SS ATPases power secretion probably through complex,
undiscovered macromolecular relays regulating pore opening and
closure at both the inner and outer membranes. It will be one of
the challenges in the next few years to elucidate the molecular
basis of such relays. In this respect, the challenge extends to all
secretion systems, whether in bacteria or eukaryotes. It is likely
that within one class of secretion systems, common features will
be discovered, as well as distinct ones, but expanding knowledge
in this field is bound to provide inspiration for designing more
efficacious secretion systems inhibitors, which, at least in the
case of bacterial pathogens, are urgently needed.
Note added in proof (received 3 December 2009)
The structure of the outer membrane complex of the T4SS
encoded by the pKM101 plasmid has just been published [156].
This complex is made of the assembly of 14 copies of each of the
TraN/VirB7 protein, the C-terminal domain of the TraO/VirB9
protein and the C-terminal domain of the TraF/VirB10 protein.
This structure demonstrates unambiguously that (i) TraO/VirB9
does not form the outer membrane channel – instead, together
with TraN/VirB7, it forms a ring surrounding the TraF/VirB10
channel and buttressing the entire system against the inner leaflet
of the outer membrane, and (ii) the antennae of TraF/VirB10
forms the outer membrane channel and projects out across the
outer membrane.
FUNDING
Work in the authors’ laboratory was funded by the Wellcome Trust [grant number 082227
(to G.W.)]. A.T.R. was supported by the Fundação para a Ciência e Tecnologia [grant
number SFRH/BD/22254/2005].
REFERENCES
1 Saier, Jr, M. H. (2006) Protein secretion and membrane insertion systems in
Gram-negative bacteria. J. Membr. Biol. 214, 75–90
2 Filloux, A., Hachani, A. and Bleves, S. (2008) The bacterial type VI secretion machine: yet
another player for protein transport across membranes. Microbiology 154, 1570–1583
3 Koronakis, V., Hughes, C. and Koronakis, E. (1991) Energetically distinct early and late
stages of HlyB/HlyD-dependent secretion across both Escherichia coli membranes.
EMBO J. 10, 3263–3272
4 Filloux, A. (2004) The underlying mechanisms of type II protein secretion. Biochim.
Biophys. Acta 1694, 163–179
5 Cornelis, G. R. and Van Gijsegem, F. (2000) Assembly and function of type III secretory
systems. Annu. Rev. Microbiol. 54, 735–774
6 Christie, P. J., Atmakuri, K., Krishnamoorthy, V., Jakubowski, S. and Cascales, E. (2005)
Biogenesis, architecture, and function of bacterial type IV secretion systems. Annu. Rev.
Microbiol. 59, 451–485
7 Fronzes, R., Christie, P. J. and Waksman, G. (2009) The structural biology of type IV
secretion systems. Nat. Rev. Microbiol. 7, 703–714
8 Henderson, I. R., Navarro-Garcia, F., Desvaux, M., Fernandez, R. C. and Ala’Aldeen, D.
(2004) Type V protein secretion pathway: the autotransporter story. Microbiol. Mol. Biol.
Rev. 68, 692–744
9 Waksman, G. and Hultgren, S. J. (2009) Structural biology of the chaperone-usher
pathway of pilus biogenesis. Nat. Rev. Microbiol. 7, 765–774
10 Thanassi, D. G., Saulino, E. T. and Hultgren, S. J. (1998) The chaperone/usher pathway:
a major terminal branch of the general secretory pathway. Curr. Opin. Microbiol. 1,
223–231
11 Dodson, K. W., Jacob-Dubuisson, F., Striker, R. T. and Hultgren, S. J. (1993)
Outer-membrane PapC molecular usher discriminately recognizes periplasmic
chaperone-pilus subunit complexes. Proc. Natl. Acad. Sci. U.S.A. 90, 3670–3674
485
12 Christie, P. J. and Cascales, E. (2005) Structural and dynamic properties of bacterial
type IV secretion systems (review). Mol. Membr. Biol. 22, 51–61
13 Fronzes, R., Remaut, H. and Waksman, G. (2008) Architectures and biogenesis of
non-flagellar protein appendages in Gram-negative bacteria. EMBO J. 27, 2271–2280
14 Barnhart, M. M., Pinkner, J. S., Soto, G. E., Sauer, F. G., Langermann, S., Waksman, G.,
Frieden, C. and Hultgren, S. J. (2000) PapD-like chaperones provide the missing
information for folding of pilin proteins. Proc. Natl. Acad. Sci. U.S.A. 97, 7709–7714
15 Nuccio, S. P. and Baumler, A. J. (2007) Evolution of the chaperone/usher assembly
pathway: fimbrial classification goes Greek. Microbiol. Mol. Biol. Rev. 71, 551–575
16 Gong, M. and Makowski, L. (1992) Helical structure of P pili from Escherichia coli .
Evidence from X-ray fiber diffraction and scanning transmission electron microscopy.
J. Mol. Biol. 228, 735–742
17 Bullitt, E. and Makowski, L. (1995) Structural polymorphism of bacterial adhesion pili.
Nature 373, 164–167
18 Kuehn, M. J., Heuser, J., Normark, S. and Hultgren, S. J. (1992) P pili in uropathogenic
E. coli are composite fibres with distinct fibrillar adhesive tips. Nature 356,
252–255
19 Roberts, J. A., Marklund, B. I., Ilver, D., Haslam, D., Kaack, M. B., Baskin, G., Louis, M.,
Mollby, R., Winberg, J. and Normark, S. (1994) The Gal(α1–4)Gal-specific tip adhesin
of Escherichia coli P-fimbriae is needed for pyelonephritis to occur in the normal urinary
tract. Proc. Natl. Acad. Sci. U.S.A. 91, 11889–11893
20 Lund, B., Lindberg, F., Marklund, B. I. and Normark, S. (1987) The PapG protein is the
α-D-galactopyranosyl-(1→4)-β-D-galactopyranose-binding adhesin of uropathogenic
Escherichia coli . Proc. Natl. Acad. Sci. U.S.A. 84, 5898–5902
21 Jones, C. H., Pinkner, J. S., Roth, R., Heuser, J., Nicholes, A. V., Abraham, S. N. and
Hultgren, S. J. (1995) FimH adhesin of type 1 pili is assembled into a fibrillar tip
structure in the Enterobacteriaceae . Proc. Natl. Acad. Sci. U.S.A. 92, 2081–2085
22 Hahn, E., Wild, P., Hermanns, U., Sebbel, P., Glockshuber, R., Haner, M., Taschner, N.,
Burkhard, P., Aebi, U. and Muller, S. A. (2002) Exploring the 3D molecular architecture
of Escherichia coli type 1 pili. J. Mol. Biol. 323, 845–857
23 Hull, R. A., Gill, R. E., Hsu, P., Minshew, B. H. and Falkow, S. (1981) Construction and
expression of recombinant plasmids encoding type 1 or D-mannose-resistant pili from a
urinary tract infection Escherichia coli isolate. Infect. Immun. 33, 933–938
24 Pratt, L. A. and Kolter, R. (1998) Genetic analysis of Escherichia coli biofilm formation:
roles of flagella, motility, chemotaxis and type I pili. Mol. Microbiol. 30, 285–293
25 Mulvey, M. A., Lopez-Boado, Y. S., Wilson, C. L., Roth, R., Parks, W. C., Heuser, J. and
Hultgren, S. J. (1998) Induction and evasion of host defenses by type 1-piliated
uropathogenic Escherichia coli . Science 282, 1494–1497
26 Baga, M., Norgren, M. and Normark, S. (1987) Biogenesis of E. coli Pap pili: papH, a
minor pilin subunit involved in cell anchoring and length modulation. Cell 49, 241–251
27 Verger, D., Miller, E., Remaut, H., Waksman, G. and Hultgren, S. (2006) Molecular
mechanism of P pilus termination in uropathogenic Escherichia coli . EMBO Rep. 7,
1228–1232
28 Rusch, S. L. and Kendall, D. A. (2007) Interactions that drive Sec-dependent bacterial
protein transport. Biochemistry 46, 9665–9673
29 Mori, H. and Ito, K. (2001) The Sec protein-translocation pathway. Trends Microbiol. 9,
494–500
30 Vetsch, M., Puorger, C., Spirig, T., Grauschopf, U., Weber-Ban, E. U. and Glockshuber,
R. (2004) Pilus chaperones represent a new type of protein-folding catalyst. Nature 431,
329–333
31 Jones, C. H., Danese, P. N., Pinkner, J. S., Silhavy, T. J. and Hultgren, S. J. (1997) The
chaperone-assisted membrane release and folding pathway is sensed by two signal
transduction systems. EMBO J. 16, 6394–6406
32 Holmgren, A. and Branden, C. I. (1989) Crystal structure of chaperone protein PapD
reveals an immunoglobulin fold. Nature 342, 248–251
33 Kuehn, M. J., Ogg, D. J., Kihlberg, J., Slonim, L. N., Flemmer, K., Bergfors, T. and
Hultgren, S. J. (1993) Structural basis of pilus subunit recognition by the PapD
chaperone. Science 262, 1234–1241
34 Slonim, L. N., Pinkner, J. S., Branden, C. I. and Hultgren, S. J. (1992) Interactive surface
in the PapD chaperone cleft is conserved in pilus chaperone superfamily and essential in
subunit recognition and assembly. EMBO J. 11, 4747–4756
35 Choudhury, D., Thompson, A., Stojanoff, V., Langermann, S., Pinkner, J., Hultgren, S. J.
and Knight, S. D. (1999) X-ray structure of the FimC-FimH chaperone-adhesin complex
from uropathogenic Escherichia coli . Science 285, 1061–1066
36 Sauer, F. G., Futterer, K., Pinkner, J. S., Dodson, K. W., Hultgren, S. J. and Waksman, G.
(1999) Structural basis of chaperone function and pilus biogenesis. Science 285,
1058–1061
37 Bann, J. G., Pinkner, J. S., Frieden, C. and Hultgren, S. J. (2004) Catalysis of protein
folding by chaperones in pathogenic bacteria. Proc. Natl. Acad. Sci. U.S.A. 101,
17389–17393
38 Remaut, H. and Waksman, G. (2006) Protein-protein interaction through β-strand
addition. Trends Biochem. Sci. 31, 436–444
c The Authors Journal compilation c 2010 Biochemical Society
486
A.T. Rêgo, V. Chandran and G. Waksman
39 Sauer, F. G., Pinkner, J. S., Waksman, G. and Hultgren, S. J. (2002) Chaperone priming
of pilus subunits facilitates a topological transition that drives fiber formation. Cell 111,
543–551
40 Zavialov, A., Berglund, J. and Knight, S. D. (2003) Overexpression, purification,
crystallization and preliminary X-ray diffraction analysis of the F1 antigen Caf1M-Caf1
chaperone-subunit pre-assembly complex from Yersinia pestis . Acta Crystallogr. D Biol.
Crystallogr. 59, 359–362
41 Vetsch, M., Erilov, D., Moliere, N., Nishiyama, M., Ignatov, O. and Glockshuber, R.
(2006) Mechanism of fibre assembly through the chaperone-usher pathway. EMBO Rep.
7, 734–738
42 Remaut, H., Rose, R. J., Hannan, T. J., Hultgren, S. J., Radford, S. E., Ashcroft, A. E. and
Waksman, G. (2006) Donor-strand exchange in chaperone-assisted pilus assembly
proceeds through a concerted β strand displacement mechanism. Mol. Cell 22,
831–842
43 Zavialov, A. V., Tischenko, V. M., Fooks, L. J., Brandsdal, B. O., Aqvist, J., Zav’yalov,
V. P., Macintyre, S. and Knight, S. D. (2005) Resolving the energy paradox of
chaperone/usher-mediated fibre assembly. Biochem. J. 389, 685–694
44 Jacob-Dubuisson, F., Striker, R. and Hultgren, S. J. (1994) Chaperone-assisted
self-assembly of pili independent of cellular energy. J. Biol. Chem. 269,
12447–12455
45 Rose, R. J., Verger, D., Daviter, T., Remaut, H., Paci, E., Waksman, G., Ashcroft, A. E. and
Radford, S. E. (2008) Unraveling the molecular basis of subunit specificity in P pilus
assembly by mass spectrometry. Proc. Natl. Acad. Sci. U.S.A. 105, 12873–12878
46 Nishiyama, M., Ishikawa, T., Rechsteiner, H. and Glockshuber, R. (2008) Reconstitution
of pilus assembly reveals a bacterial outer membrane catalyst. Science 320, 376–379
47 Thanassi, D. G., Stathopoulos, C., Dodson, K., Geiger, D. and Hultgren, S. J. (2002)
Bacterial outer membrane ushers contain distinct targeting and assembly domains for
pilus biogenesis. J. Bacteriol. 184, 6260–6269
48 Nishiyama, M., Vetsch, M., Puorger, C., Jelesarov, I. and Glockshuber, R. (2003)
Identification and characterization of the chaperone-subunit complex-binding domain
from the type 1 pilus assembly platform FimD. J. Mol. Biol. 330, 513–525
49 Nishiyama, M., Horst, R., Eidam, O., Herrmann, T., Ignatov, O., Vetsch, M., Bettendorff,
P., Jelesarov, I., Grutter, M. G., Wuthrich, K. et al. (2005) Structural basis of
chaperone-subunit complex recognition by the type 1 pilus assembly platform FimD.
EMBO J. 24, 2075–2086
50 Ng, T. W., Akman, L., Osisami, M. and Thanassi, D. G. (2004) The usher N terminus is
the initial targeting site for chaperone-subunit complexes and participates in subsequent
pilus biogenesis events. J. Bacteriol. 186, 5321–5331
51 Eidam, O., Dworkowski, F. S., Glockshuber, R., Grutter, M. G. and Capitani, G. (2008)
Crystal structure of the ternary FimC-FimF(t)-FimD(N) complex indicates conserved
pilus chaperone-subunit complex recognition by the usher FimD. FEBS Lett. 582,
651–655
52 Remaut, H., Tang, C., Henderson, N. S., Pinkner, J. S., Wang, T., Hultgren, S. J.,
Thanassi, D. G., Waksman, G. and Li, H. (2008) Fiber formation across the bacterial
outer membrane by the chaperone/usher pathway. Cell 133, 640–652
53 Huang, Y., Smith, B. S., Chen, L. X., Baxter, R. H. and Deisenhofer, J. (2009) Insights
into pilus assembly and secretion from the structure and functional characterization of
usher PapC. Proc. Natl. Acad. Sci. U.S.A. 106, 7403–7407
54 So, S. S. and Thanassi, D. G. (2006) Analysis of the requirements for pilus biogenesis at
the outer membrane usher and the function of the usher C-terminus. Mol. Microbiol. 60,
364–375
55 Saulino, E. T., Thanassi, D. G., Pinkner, J. S. and Hultgren, S. J. (1998) Ramifications of
kinetic partitioning on usher-mediated pilus biogenesis. EMBO J. 17, 2177–2185
56 Munera, D., Palomino, C. and Fernandez, L. A. (2008) Specific residues in the
N-terminal domain of FimH stimulate type 1 fimbriae assembly in Escherichia coli
following the initial binding of the adhesin to FimD usher. Mol. Microbiol. 69, 911–925
57 Munera, D., Hultgren, S. and Fernandez, L. A. (2007) Recognition of the N-terminal
lectin domain of FimH adhesin by the usher FimD is required for type 1 pilus
biogenesis. Mol. Microbiol. 64, 333–346
58 Li, H., Qian, L., Chen, Z., Thibault, D., Liu, G., Liu, T. and Thanassi, D. G. (2004) The
outer membrane usher forms a twin-pore secretion complex. J. Mol. Biol. 344,
1397–1407
59 Cascales, E. and Christie, P. J. (2003) The versatile bacterial type IV secretion systems.
Nat. Rev. Microbiol. 1, 137–149
60 Schroder, G. and Lanka, E. (2005) The mating pair formation system of conjugative
plasmids - a versatile secretion machinery for transfer of proteins and DNA. Plasmid
54, 1–25
61 Backert, S. and Selbach, M. (2008) Role of type IV secretion in Helicobacter pylori
pathogenesis. Cell. Microbiol. 10, 1573–1581
62 Ninio, S. and Roy, C. R. (2007) Effector proteins translocated by Legionella
pneumophila : strength in numbers. Trends Microbiol. 15, 372–380
63 Smeets, L. C. and Kusters, J. G. (2002) Natural transformation in Helicobacter pylori :
DNA transport in an unexpected way. Trends Microbiol. 10, 159–162
c The Authors Journal compilation c 2010 Biochemical Society
64 McCullen, C. A. and Binns, A. N. (2006) Agrobacterium tumefaciens and plant cell
interactions and activities required for interkingdom macromolecular transfer. Annu. Rev.
Cell Dev. Biol. 22, 101–127
65 Thomas, C. M. and Nielsen, K. M. (2005) Mechanisms of, and barriers to, horizontal
gene transfer between bacteria. Nat. Rev. Microbiol. 3, 711–721
66 Cao, T. B. and Saier, Jr, M. H. (2001) Conjugal type IV macromolecular transfer systems
of Gram-negative bacteria: organismal distribution, structural constraints and
evolutionary conclusions. Microbiology 147, 3201–3214
67 Atmakuri, K., Cascales, E. and Christie, P. J. (2004) Energetic components VirD4, VirB11
and VirB4 mediate early DNA transfer reactions required for bacterial type IV secretion.
Mol. Microbiol. 54, 1199–1211
68 Das, A. and Xie, Y. H. (1998) Construction of transposon Tn3phoA: its application in
defining the membrane topology of the Agrobacterium tumefaciens DNA transfer
proteins. Mol. Microbiol. 27, 405–414
69 Hormaeche, I., Segura, R. L., Vecino, A. J., Goni, F. M., de la Cruz, F. and Alkorta, I.
(2006) The transmembrane domain provides nucleotide binding specificity to the
bacterial conjugation protein TrwB. FEBS Lett. 580, 3075–3082
70 Hormaeche, I., Iloro, I., Arrondo, J. L., Goni, F. M., de la Cruz, F. and Alkorta, I. (2004)
Role of the transmembrane domain in the stability of TrwB, an integral protein involved
in bacterial conjugation. J. Biol. Chem. 279, 10955–10961
71 Gomis-Ruth, F. X., Moncalian, G., de la Cruz, F. and Coll, M. (2002) Conjugative plasmid
protein TrwB, an integral membrane type IV secretion system coupling protein. Detailed
structural features and mapping of the active site cleft. J. Biol. Chem. 277, 7556–7566
72 Moncalian, G., Cabezon, E., Alkorta, I., Valle, M., Moro, F., Valpuesta, J. M., Goni, F. M.
and de La Cruz, F. (1999) Characterization of ATP and DNA binding activities of TrwB,
the coupling protein essential in plasmid R388 conjugation. J. Biol. Chem. 274,
36117–36124
73 Tato, I., Zunzunegui, S., de la Cruz, F. and Cabezon, E. (2005) TrwB, the coupling protein
involved in DNA transport during bacterial conjugation, is a DNA-dependent ATPase.
Proc. Natl. Acad. Sci. U.S.A. 102, 8156–8161
74 Planet, P. J., Kachlany, S. C., DeSalle, R. and Figurski, D. H. (2001) Phylogeny of genes
for secretion NTPases: identification of the widespread tadA subfamily and development
of a diagnostic key for gene classification. Proc. Natl. Acad. Sci. U.S.A. 98, 2503–2508
75 Filloux, A. (2009) The type VI secretion system: a tubular story. EMBO J. 28, 309–310
76 Sagulenko, E., Sagulenko, V., Chen, J. and Christie, P. J. (2001) Role of Agrobacterium
VirB11 ATPase in T-pilus assembly and substrate selection. J. Bacteriol. 183,
5813–5825
77 Krause, S., Barcena, M., Pansegrau, W., Lurz, R., Carazo, J. M. and Lanka, E. (2000)
Sequence-related protein export NTPases encoded by the conjugative transfer region of
RP4 and by the cag pathogenicity island of Helicobacter pylori share similar hexameric
ring structures. Proc. Natl. Acad. Sci. U.S.A. 97, 3067–3072
78 Rivas, S., Bolland, S., Cabezon, E., Goni, F. M. and de la Cruz, F. (1997) TrwD, a protein
encoded by the IncW plasmid R388, displays an ATP hydrolase activity essential for
bacterial conjugation. J. Biol. Chem. 272, 25583–25590
79 Machon, C., Rivas, S., Albert, A., Goni, F. M. and de la Cruz, F. (2002) TrwD, the
hexameric traffic ATPase encoded by plasmid R388, induces membrane destabilization
and hemifusion of lipid vesicles. J. Bacteriol. 184, 1661–1668
80 Yeo, H. J., Savvides, S. N., Herr, A. B., Lanka, E. and Waksman, G. (2000) Crystal
structure of the hexameric traffic ATPase of the Helicobacter pylori type IV secretion
system. Mol. Cell 6, 1461–1472
81 Savvides, S. N., Yeo, H. J., Beck, M. R., Blaesing, F., Lurz, R., Lanka, E., Buhrdorf, R.,
Fischer, W., Haas, R. and Waksman, G. (2003) VirB11 ATPases are dynamic hexameric
assemblies: new insights into bacterial type IV secretion. EMBO J. 22, 1969–1980
82 Hare, S., Bayliss, R., Baron, C. and Waksman, G. (2006) A large domain swap in the
VirB11 ATPase of Brucella suis leaves the hexameric assembly intact. J. Mol. Biol.
360, 56–66
83 Yeo, H. J. and Waksman, G. (2004) Unveiling molecular scaffolds of the type IV
secretion system. J. Bacteriol. 186, 1919–1926
84 Grohmann, E., Muth, G. and Espinosa, M. (2003) Conjugative plasmid transfer in
Gram-positive bacteria. Microbiol. Mol. Biol. Rev. 67, 277–301
85 Rabel, C., Grahn, A. M., Lurz, R. and Lanka, E. (2003) The VirB4 family of proposed
traffic nucleoside triphosphatases: common motifs in plasmid RP4 TrbE are essential for
conjugation and phage adsorption. J. Bacteriol. 185, 1045–1058
86 Fullner, K. J., Stephens, K. M. and Nester, E. W. (1994) An essential virulence protein of
Agrobacterium tumefaciens , VirB4, requires an intact mononucleotide binding domain
to function in transfer of T-DNA. Mol. Gen. Genet. 245, 704–715
87 Shirasu, K., Koukolikova-Nicola, Z., Hohn, B. and Kado, C. I. (1994) An
inner-membrane-associated virulence protein essential for T-DNA transfer from
Agrobacterium tumefaciens to plants exhibits ATPase activity and similarities to
conjugative transfer genes. Mol. Microbiol. 11, 581–588
88 Arechaga, I., Pena, A., Zunzunegui, S., del Carmen Fernandez-Alonso, M., Rivas, G. and
de la Cruz, F. (2008) ATPase activity and oligomeric state of TrwK, the VirB4 homologue
of the plasmid R388 type IV secretion system. J. Bacteriol. 190, 5472–5479
Two-step versus one-step secretion systems
89 Dang, T. A. and Christie, P. J. (1997) The VirB4 ATPase of Agrobacterium tumefaciens is
a cytoplasmic membrane protein exposed at the periplasmic surface. J. Bacteriol. 179,
453–462
90 Middleton, R., Sjolander, K., Krishnamurthy, N., Foley, J. and Zambryski, P. (2005)
Predicted hexameric structure of the Agrobacterium VirB4 C terminus suggests VirB4
acts as a docking site during type IV secretion. Proc. Natl. Acad. Sci. U.S.A. 102,
1685–1690
91 Draper, O., Middleton, R., Doucleff, M. and Zambryski, P. C. (2006) Topology of the
VirB4 C terminus in the Agrobacterium tumefaciens VirB/D4 type IV secretion system.
J. Biol. Chem. 281, 37628–37635
92 Jakubowski, S. J., Krishnamoorthy, V., Cascales, E. and Christie, P. J. (2004)
Agrobacterium tumefaciens VirB6 domains direct the ordered export of a DNA substrate
through a type IV secretion system. J. Mol. Biol. 341, 961–977
93 Karnholz, A., Hoefler, C., Odenbreit, S., Fischer, W., Hofreuter, D. and Haas, R. (2006)
Functional and topological characterization of novel components of the comB DNA
transformation competence system in Helicobacter pylori . J. Bacteriol. 188, 882–893
94 Kumar, R. B., Xie, Y. H. and Das, A. (2000) Subcellular localization of the Agrobacterium
tumefaciens T-DNA transport pore proteins: VirB8 is essential for the assembly of the
transport pore. Mol. Microbiol. 36, 608–617
95 Kumar, R. B. and Das, A. (2001) Functional analysis of the Agrobacterium tumefaciens
T-DNA transport pore protein VirB8. J. Bacteriol. 183, 3636–3641
96 Krall, L., Wiedemann, U., Unsin, G., Weiss, S., Domke, N. and Baron, C. (2002)
Detergent extraction identifies different VirB protein subassemblies of the type IV
secretion machinery in the membranes of Agrobacterium tumefaciens . Proc. Natl. Acad.
Sci. U.S.A. 99, 11405–11410
97 Ward, D. V., Draper, O., Zupan, J. R. and Zambryski, P. C. (2002) Peptide linkage
mapping of the Agrobacterium tumefaciens vir-encoded type IV secretion system reveals
protein subassemblies. Proc. Natl. Acad. Sci. U.S.A. 99, 11493–11500
98 Yuan, Q., Carle, A., Gao, C., Sivanesan, D., Aly, K. A., Hoppner, C., Krall, L., Domke, N.
and Baron, C. (2005) Identification of the VirB4-VirB8-VirB5-VirB2 pilus assembly
sequence of type IV secretion systems. J. Biol. Chem. 280, 26349–26359
99 Hoppner, C., Carle, A., Sivanesan, D., Hoeppner, S. and Baron, C. (2005) The putative
lytic transglycosylase VirB1 from Brucella suis interacts with the type IV secretion
system core components VirB8, VirB9 and VirB11. Microbiology 151, 3469–3482
100 Das, A. and Xie, Y. H. (2000) The Agrobacterium T-DNA transport pore proteins VirB8,
VirB9, and VirB10 interact with one another. J. Bacteriol. 182, 758–763
101 Bailey, S., Ward, D., Middleton, R., Grossmann, J. G. and Zambryski, P. C. (2006)
Agrobacterium tumefaciens VirB8 structure reveals potential protein-protein interaction
sites. Proc. Natl. Acad. Sci. U.S.A. 103, 2582–2587
102 Terradot, L., Bayliss, R., Oomen, C., Leonard, G. A., Baron, C. and Waksman, G. (2005)
Structures of two core subunits of the bacterial type IV secretion system, VirB8 from
Brucella suis and ComB10 from Helicobacter pylori . Proc. Natl. Acad. Sci. U.S.A. 102,
4596–4601
103 Paschos, A., Patey, G., Sivanesan, D., Gao, C., Bayliss, R., Waksman, G., O’Callaghan,
D. and Baron, C. (2006) Dimerization and interactions of Brucella suis VirB8 with VirB4
and VirB10 are required for its biological activity. Proc. Natl. Acad. Sci. U.S.A. 103,
7252–7257
104 Jakubowski, S. J., Kerr, J. E., Garza, I., Krishnamoorthy, V., Bayliss, R., Waksman, G. and
Christie, P. J. (2009) Agrobacterium VirB10 domain requirements for type IV secretion
and T pilus biogenesis. Mol. Microbiol. 71, 779–794
105 Gilmour, M. W., Gunton, J. E., Lawley, T. D. and Taylor, D. E. (2003) Interaction between
the IncHI1 plasmid R27 coupling protein and type IV secretion system: TraG associates
with the coiled-coil mating pair formation protein TrhB. Mol. Microbiol. 49, 105–116
106 Terradot, L., Durnell, N., Li, M., Ory, J., Labigne, A., Legrain, P., Colland, F. and
Waksman, G. (2004) Biochemical characterization of protein complexes from the
Helicobacter pylori protein interaction map: strategies for complex formation and
evidence for novel interactions within type IV secretion systems. Mol. Cell. Proteomics
3, 809–819
107 Llosa, M., Zunzunegui, S. and de la Cruz, F. (2003) Conjugative coupling proteins
interact with cognate and heterologous VirB10-like proteins while exhibiting specificity
for cognate relaxosomes. Proc. Natl. Acad. Sci. U.S.A. 100, 10465–10470
108 Finberg, K. E., Muth, T. R., Young, S. P., Maken, J. B., Heitritter, S. M., Binns, A. N. and
Banta, L. M. (1995) Interactions of VirB9, -10, and -11 with the membrane fraction of
Agrobacterium tumefaciens : solubility studies provide evidence for tight associations.
J. Bacteriol. 177, 4881–4889
109 Cascales, E. and Christie, P. J. (2004) Agrobacterium VirB10, an ATP energy sensor
required for type IV secretion. Proc. Natl. Acad. Sci. U.S.A. 101, 17228–17233
110 Fernandez, D., Dang, T. A., Spudich, G. M., Zhou, X. R., Berger, B. R. and Christie, P. J.
(1996) The Agrobacterium tumefaciens virB7 gene product, a proposed component of
the T-complex transport apparatus, is a membrane-associated lipoprotein exposed at the
periplasmic surface. J. Bacteriol. 178, 3156–3167
487
111 Baron, C., Thorstenson, Y. R. and Zambryski, P. C. (1997) The lipoprotein VirB7 interacts
with VirB9 in the membranes of Agrobacterium tumefaciens . J. Bacteriol. 179,
1211–1218
112 Jakubowski, S. J., Cascales, E., Krishnamoorthy, V. and Christie, P. J. (2005)
Agrobacterium tumefaciens VirB9, an outer-membrane-associated component of a
type IV secretion system, regulates substrate selection and T-pilus biogenesis. J.
Bacteriol. 187, 3486–3495
113 Bayliss, R., Harris, R., Coutte, L., Monier, A., Fronzes, R., Christie, P. J., Driscoll, P. C.
and Waksman, G. (2007) NMR structure of a complex between the VirB9/VirB7
interaction domains of the pKM101 type IV secretion system. Proc. Natl. Acad. Sci.
U.S.A. 104, 1673–1678
114 Babic, A., Lindner, A. B., Vulic, M., Stewart, E. J. and Radman, M. (2008) Direct
visualization of horizontal gene transfer. Science 319, 1533–1536
115 Kwok, T., Zabler, D., Urman, S., Rohde, M., Hartig, R., Wessler, S., Misselwitz, R., Berger,
J., Sewald, N., Konig, W. and Backert, S. (2007) Helicobacter exploits integrin for type IV
secretion and kinase activation. Nature 449, 862–866
116 Backert, S., Fronzes, R. and Waksman, G. (2008) VirB2 and VirB5 proteins: specialized
adhesins in bacterial type-IV secretion systems? Trends Microbiol. 16, 409–413
117 Lai, E. M. and Kado, C. I. (1998) Processed VirB2 is the major subunit of the
promiscuous pilus of Agrobacterium tumefaciens . J. Bacteriol. 180, 2711–2717
118 Schmidt-Eisenlohr, H., Domke, N., Angerer, C., Wanner, G., Zambryski, P. C. and Baron,
C. (1999) Vir proteins stabilize VirB5 and mediate its association with the T pilus of
Agrobacterium tumefaciens . J. Bacteriol. 181, 7485–7492
119 Majdalani, N. and Ippen-Ihler, K. (1996) Membrane insertion of the F-pilin subunit is
Sec independent but requires leader peptidase B and the proton motive force. J.
Bacteriol. 178, 3742–3747
120 Majdalani, N., Moore, D., Maneewannakul, S. and Ippen-Ihler, K. (1996) Role of the
propilin leader peptide in the maturation of F pilin. J. Bacteriol. 178, 3748–3754
121 Lai, E. M., Eisenbrandt, R., Kalkum, M., Lanka, E. and Kado, C. I. (2002) Biogenesis of T
pili in Agrobacterium tumefaciens requires precise VirB2 propilin cleavage and
cyclization. J. Bacteriol. 184, 327–330
122 Jones, A. L., Lai, E. M., Shirasu, K. and Kado, C. I. (1996) VirB2 is a processed pilin-like
protein encoded by the Agrobacterium tumefaciens Ti plasmid. J. Bacteriol. 178,
5706–5711
123 Yeo, H. J., Yuan, Q., Beck, M. R., Baron, C. and Waksman, G. (2003) Structural and
functional characterization of the VirB5 protein from the type IV secretion system
encoded by the conjugative plasmid pKM101. Proc. Natl. Acad. Sci. U.S.A. 100,
15947–15952
124 Wang, Y. A., Yu, X., Silverman, P. M., Harris, R. L. and Egelman, E. H. (2009) The
structure of F-pili. J. Mol. Biol. 385, 22–29
125 Berger, B. R. and Christie, P. J. (1994) Genetic complementation analysis of the
Agrobacterium tumefaciens virB operon: virB2 through virB11 are essential virulence
genes. J. Bacteriol. 176, 3646–3660
126 Cascales, E. and Christie, P. J. (2004) Definition of a bacterial type IV secretion pathway
for a DNA substrate. Science 304, 1170–1173
127 Zupan, J., Hackworth, C. A., Aguilar, J., Ward, D. and Zambryski, P. (2007) VirB1*
promotes T-pilus formation in the vir-Type IV secretion system of Agrobacterium
tumefaciens . J. Bacteriol. 189, 6551–6563
128 Rohde, M., Puls, J., Buhrdorf, R., Fischer, W. and Haas, R. (2003) A novel sheathed
surface organelle of the Helicobacter pylori cag type IV secretion system. Mol.
Microbiol. 49, 219–234
129 Bayer, M., Eferl, R., Zellnig, G., Teferle, K., Dijkstra, A., Koraimann, G. and Hogenauer, G.
(1995) Gene 19 of plasmid R1 is required for both efficient conjugative DNA transfer and
bacteriophage R17 infection. J. Bacteriol. 177, 4279–4288
130 Baron, C., Llosa, M., Zhou, S. and Zambryski, P. C. (1997) VirB1, a component of the
T-complex transfer machinery of Agrobacterium tumefaciens , is processed to a
C-terminal secreted product, VirB1. J. Bacteriol. 179, 1203–1210
131 Jones, A. L., Shirasu, K. and Kado, C. I. (1994) The product of the virB4 gene of
Agrobacterium tumefaciens promotes accumulation of VirB3 protein. J. Bacteriol. 176,
5255–5261
132 Beijersbergen, A., Smith, S. J. and Hooykaas, P. J. (1994) Localization and topology of
VirB proteins of Agrobacterium tumefaciens . Plasmid 32, 212–218
133 Batchelor, R. A., Pearson, B. M., Friis, L. M., Guerry, P. and Wells, J. M. (2004)
Nucleotide sequences and comparison of two large conjugative plasmids from different
Campylobacter species. Microbiology 150, 3507–3517
134 Lai, E. M., Chesnokova, O., Banta, L. M. and Kado, C. I. (2000) Genetic and
environmental factors affecting T-pilin export and T-pilus biogenesis in relation to
flagellation of Agrobacterium tumefaciens . J. Bacteriol. 182, 3705–3716
135 Fronzes, R., Schafer, E., Wang, L., Saibil, H. R., Orlova, E. V. and Waksman, G. (2009)
Structure of a type IV secretion system core complex. Science 323, 266–268
c The Authors Journal compilation c 2010 Biochemical Society
488
A.T. Rêgo, V. Chandran and G. Waksman
136 Nikaido, H. and Takatsuka, Y. (2009) Mechanisms of RND multidrug efflux pumps.
Biochim. Biophys. Acta 1794, 769–781
137 Pos, K. M. (2009) Drug transport mechanism of the AcrB efflux pump. Biochim.
Biophys. Acta 1794, 782–793
138 Durand, E., Verger, D., Rego, A. T., Chandran, V., Meng, G., Fronzes, R. and Waksman,
G. (2009) Structural biology of bacterial secretion systems in Gram-negative pathogens
– potential for new drug targets. Infect. Disord. Drug Targets 9, 518–547
139 Zhao, Z., Sagulenko, E., Ding, Z. and Christie, P. J. (2001) Activities of virE1 and the
VirE1 secretion chaperone in export of the multifunctional VirE2 effector via an
Agrobacterium type IV secretion pathway. J. Bacteriol. 183, 3855–3865
140 Deng, W., Chen, L., Peng, W. T., Liang, X., Sekiguchi, S., Gordon, M. P., Comai, L. and
Nester, E. W. (1999) VirE1 is a specific molecular chaperone for the exported
single-stranded-DNA-binding protein VirE2 in Agrobacterium . Mol. Microbiol. 31,
1795–1807
141 Zhou, X. R. and Christie, P. J. (1999) Mutagenesis of the Agrobacterium VirE2
single-stranded DNA-binding protein identifies regions required for self-association and
interaction with VirE1 and a permissive site for hybrid protein construction. J. Bacteriol.
181, 4342–4352
142 Sundberg, C., Meek, L., Carroll, K., Das, A. and Ream, W. (1996) VirE1 protein mediates
export of the single-stranded DNA-binding protein VirE2 from Agrobacterium
tumefaciens into plant cells. J. Bacteriol. 178, 1207–1212
143 Sundberg, C. D. and Ream, W. (1999) The Agrobacterium tumefaciens chaperone-like
protein, VirE1, interacts with VirE2 at domains required for single-stranded DNA binding
and cooperative interaction. J. Bacteriol. 181, 6850–6855
144 Ninio, S., Zuckman-Cholon, D. M., Cambronne, E. D. and Roy, C. R. (2005) The
Legionella IcmS-IcmW protein complex is important for Dot/Icm-mediated protein
translocation. Mol. Microbiol. 55, 912–926
145 Bardill, J. P., Miller, J. L. and Vogel, J. P. (2005) IcmS-dependent translocation of SdeA
into macrophages by the Legionella pneumophila type IV secretion system. Mol.
Microbiol. 56, 90–103
Received 6 October 2009/16 November 2009; accepted 18 November 2009
Published on the Internet 15 January 2010, doi:10.1042/BJ20091518
c The Authors Journal compilation c 2010 Biochemical Society
146 Vincent, C. D. and Vogel, J. P. (2006) The Legionella pneumophila IcmS-LvgA protein
complex is important for Dot/Icm-dependent intracellular growth. Mol. Microbiol. 61,
596–613
147 Pattis, I., Weiss, E., Laugks, R., Haas, R. and Fischer, W. (2007) The Helicobacter pylori
CagF protein is a type IV secretion chaperone-like molecule that binds close to the
C-terminal secretion signal of the CagA effector protein. Microbiology 153, 2896–2909
148 Bradley, D. E., Taylor, D. E. and Cohen, D. R. (1980) Specification of surface mating
systems among conjugative drug resistance plasmids in Escherichia coli K-12. J.
Bacteriol. 143, 1466–1470
149 Lawley, T. D., Klimke, W. A., Gubbins, M. J. and Frost, L. S. (2003) F factor conjugation
is a true type IV secretion system. FEMS Microbiol. Lett. 224, 1–15
150 Eisenbrandt, R., Kalkum, M., Lai, E. M., Lurz, R., Kado, C. I. and Lanka, E. (1999)
Conjugative pili of IncP plasmids, and the Ti plasmid T pilus are composed of cyclic
subunits. J. Biol. Chem. 274, 22548–22555
151 Tanaka, J., Suzuki, T., Mimuro, H. and Sasakawa, C. (2003) Structural definition on the
surface of Helicobacter pylori type IV secretion apparatus. Cell. Microbiol. 5, 395–404
152 Soto, G. E. and Hultgren, S. J. (1999) Bacterial adhesins: common themes and
variations in architecture and assembly. J. Bacteriol. 181, 1059–1071
153 Hung, D. L., Knight, S. D., Woods, R. M., Pinkner, J. S. and Hultgren, S. J. (1996)
Molecular basis of two subfamilies of immunoglobulin-like chaperones. EMBO J. 15,
3792–3805
154 Anderson, K. L., Billington, J., Pettigrew, D., Cota, E., Simpson, P., Roversi, P., Chen,
H. A., Urvil, P., du Merle, L., Barlow, P. N. et al. (2004) An atomic resolution model for
assembly, architecture, and function of the Dr adhesins. Mol. Cell 15, 647–657
155 Gerlach, R. G. and Hensel, M. (2007) Protein secretion systems and adhesins: the
molecular armory of Gram-negative pathogens. Int. J. Med. Microbiol. 297,
401–415
156 Chandran, V., Fronzes, R., Duquerroy, S., Cronin, N., Navaza, J. and Waksman, G.
(2009) Structure of the outer membrane complex of a type IV secretion system. Nature
462, 1011–1015