ArticlePDF AvailableLiterature Review

Papanikou, E., Karamanou, S. & Economou, A. Bacterial protein secretion through the translocase nanomachine. Nature Rev. Microbiol. 5, 839-851

Authors:

Abstract

All cells must traffic proteins across their membranes. This essential process is responsible for the biogenesis of membranes and cell walls, motility and nutrient scavenging and uptake, and is also involved in pathogenesis and symbiosis. The translocase is an impressively dynamic nanomachine that is the central component which catalyses transmembrane crossing. This complex, multi-stage reaction involves a cascade of inter- and intramolecular interactions that select, sort and target polypeptides to the membrane, and use energy to promote the movement of these polypeptides across--or their lateral escape and integration into--the phospholipid bilayer, with high fidelity and efficiency. Here, we review the most recent data on the structure and function of the translocase nanomachine.
In cells from all three domains of life, more than one-
third of the proteome is secreted across, or inserted
into, biological membranes. Secretory proteins include
hydrolytic enzymes, periplasmic lipoproteins, toxins
and surface appendages such as pilli and flagella.
Integral membrane proteins mediate selective trans-
port, energy conversion, cell division, extracellular sig-
nal sensing, and membrane and cell-wall biogenesis.
In many cases, the plasma membrane is the main
and in Gram-positive bacteria and archaea the
only — membrane that is involved. However, proteins
are also targeted to several other membranes, such as
those of eukaryotic sub-cellular organelles, the outer
membrane of Gram-negative bacteria and the inter-
nal-membrane enclaves of photosynthetic bacteria.
Moreover, in the case of bacterial pathogens that infect
eukaryotic hosts, bacterial exotoxins must negotiate
multiple self and host membranes before they reach
their target locations.
To successfully localize polypeptides extracytoplas-
mically, the cell must tackle five daunting tasks: discrimi-
nate the cytoplasmic-resident proteins from those that
are destined for export; deal with the inherent tendency
of polypeptides to fold rapidly; target exported proteins
to the membrane with specificity and fidelity; achieve
transmembrane crossing of these elongated, heteropoly-
meric substrates, which are several times as long as the
membrane is wide; and finally, manage a second sorting
event that releases membrane proteins into the lipid
bilayer and secretory proteins to the trans side of the
membrane.
Evolution has produced a remarkable array of
mechanisms to export proteins. Sixteen such systems,
which handle protein secretion, sorting and membrane
integration, are present in the Bacteria alone
1,2
(BOX 1).
Of these, only the Sec pathway is ubiquitous and essen-
tial for viability in all three domains of life. In addition,
the Sec pathway acts as the entry point for many of the
other protein export and sorting pathways (for Bacteria
see BOX 1). Over the past 30 years, all the genes that are
involved in the Sec pathway have been identified using
genetic and biochemical approaches. Export-specific
chaperones, pilot-like factors and signals in exported
proteins overcome the sorting and folding problem
and facilitate membrane targeting
3,4
. A dynamic mem-
brane-embedded nanomachine called the pre-protein
translocase or translocon (FIG. 1a) recognizes exported
proteins at the membrane and catalyses their export.
High-
5,6
and medium to low-
7–13
resolution structures
of the protein-conducting channel SecYEG, of both
eukaryotic and prokaryotic origin, and high-resolution
structures of the bacterial SecA motor
14–19
, are now avail-
able. Sequential recognition events and the expenditure
of energy allow the polymeric polypeptide substrate to
move through the translocase, effectively being threaded
through the membrane, and allow hydrophobic regions
to escape laterally into the lipid bilayer.
Here, we will review the latest advances in bacterial
protein secretion through the translocase, focusing
on the journey that secretory proteins take across the
bacterial plasma membrane. Membrane-protein inte-
gration through the translocase will not be discussed in
*Institute of Molecular
Biology and Biotechnology,
Foundation of Research
and Technology-Hellas,
PO Box 1385, Heraklion
GR-71110, Crete, Greece.
Department of Biology,
University of Crete, Iraklio
GR-71110, Crete, Greece.
e-mail:
aeconomo@imbb.
forth.gr
doi:10.1038/nrmicro1771
Bacterial protein secretion through
the translocase nanomachine
Effrosyni Papanikou*, Spyridoula Karamanou* and Anastassios Economou*
Abstract | All cells must traffic proteins across their membranes. This essential process is
responsible for the biogenesis of membranes and cell walls, motility and nutrient
scavenging and uptake, and is also involved in pathogenesis and symbiosis. The translocase
is an impressively dynamic nanomachine that is the central component which catalyses
transmembrane crossing. This complex, multi-stage reaction involves a cascade of
inter- and intramolecular interactions that select, sort and target polypeptides to the
membrane, and use energy to promote the movement of these polypeptides across — or
their lateral escape and integration into — the phospholipid bilayer, with high fidelity and
efficiency. Here, we review the most recent data on the structure and function of the
translocase nanomachine.
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
839
REVIEWS
Box 1 | The bacterial protein-export and secretion-systems zoo
The figure shows a simplified
schematic summary of the
major protein-export, protein-
secretion and protein-
membrane-integration
systems in Bacteria. Peptide-
processing peptidases,
periplasmic chaperones and
folding factors are not shown.
For a comprehensive
presentation, the reader is
directed to REFS 1, 125. The
nomenclature of many
systems follows the type-‘n’-
secretion or type-‘n’-pilus
convention. The type IV
secretion system (T4SS) shares
ancestry with conjugation
systems and can also transport
DNA. The type VI secretion
system (T6SS) has only recently been identified and how exported proteins cross the cell envelope in this system is not
yet clear. The membrane protein YidC has homologues (known as Alb3 or Oxa1) that function in mitochondria and
chloroplasts but not in the archaeal or endoplasmic reticulum membrane
95
. The arrows indicate the path that is taken
by the exported protein. Arrows that initiate in the periplasm indicate that Sec (or rarely TAT)-dependent
translocation across the plasma membrane is a necessary first step for these systems.
Bam, beta-barrel assembly machine; CU, chaperone–usher pathway; Esx, specialized secretion system that is found
in Gram-positive bacteria (for example, mycobacteria); Fla, flagellum; HM, host cell membrane; LOL, lipoprotein
outer-membrane localization; OM, outer membrane; Omp85, also known as YaeT; Per, periplasm; PM, plasma
membrane; Sort, sortase; TPS, two-partner secretion; T2S, type II secretion; T3S, type III secretion; T4P, type IV pili;
T4S, type IV secretion; T5S, type V secretion (autotransportation), T6S, type VI secretion.
detail (BOX 2). These findings are mainly the result of the
focused dissection of the secretion pathway in a single
model bacterium, Escherichia coli. Nevertheless, genome
sequence data and some biochemical data suggest that
most of these Sec-pathway features are universal in the
Bacteria.
Overview of the Sec pathway
Secretory-protein export through the Sec pathway
is a multi-stage reaction that occurs mainly post-
translationally. This continuous process can be con-
ceptually and biochemically divided into three distinct
stages (FIG. 1b).
Stage one: protein sorting and targeting. Secretory proteins
are usually referred to as pre-proteins because they carry
removable amino (N)-terminal signal peptides. These
are 20–30-residue extensions, which contain 1 (or more)
positive charge at the N terminus and a short hydrophobic
core of 8–12 residues. One role of these ‘address tags’ is
to sort secretory proteins from cytoplasmic proteins.
Nascent pre-proteins are recognized directly by piloting
factors, such as the ribonucleoprotein signal-recognition
particle (SRP)
3
or the SecB chaperone
4,20
. Despite the fact
that in bacteria the SRP is predominantly involved in the
targeting of inner-membrane proteins (BOX 2), there are
examples of long, strongly hydrophobic signal peptides
of nascent secretory proteins, which are probably pref-
erentially bound by the SRP
21,22
. Other signal peptides
delay pre-protein folding and thus allow SecB to bind
to the mature region of the pre-protein
4
(FIG. 1b, part 1).
This process can occur while the polypeptide chain is still
being synthesized, but it is not mechanistically coupled to
elongation
23
. In both cases, the resulting SRP–pre-protein
and SecB–pre-protein complexes are targeted to the
translocase at the membrane (FIG. 1b, part 1). For SRP, this
is achieved by docking to its membrane receptor FtsY
3
,
and for SecB, by docking to the SecA subunit of the trans-
locase
24
. Although SecA does not contribute to the SRP-
targeting route
25
, when long, hydrophilic segments are
encountered, SecA is recruited to catalyse their export
26
.
Stage two: translocation. Irrespective of the targeting
route, all pre-proteins eventually reach the translocase at
the membrane (FIG. 1b, part 2). The translocase comprises
a membrane-embedded protein-conducting channel that
is built of the SecY, SecE and SecG polypeptides (corre-
sponding to the Sec61α, Sec61γ and Sec61β polypeptides,
respectively, in eukaryotes), through which pre-proteins
cross the membrane, and a flexible molecular motor, the
SecA ATPase, which drives translocation at the expense
of metabolic energy in the form of ATP and the proton-
motive force (PMF). Other cytoplasmic and membrane
subunits (FIG. 1a) optimize the translocation reaction.
Stage three: release and maturation. At the last stage,
pre-proteins are converted into mature proteins (FIG.1b,
part 3). After signal peptidase cleaves the signal peptides
27
,
correct folding of the polypeptide chain is initiated at the
trans side of the membrane
28–30
.
Nature Reviews | Microbiology
OM
PM
Per
HM
T6S T4S Fla T3S CU T4P T2S LOL T5S TPS TAT Sec YidC Sort Esx Omp
85
+ + + + + + + + + + + + + +
+ + + + + + +
Gram-negative
Gram-positive
REVIE W S
840
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
The structure of the translocase and its parts
The translocase consists of the membrane protein
SecYEG, the SecA ATPase and a subset of partner pro-
teins (FIG. 1a). Each component will be discussed in turn
in the following section.
The SecYEG complex. The translocase core consists of
the integral membrane SecYEG heterotrimer. SecYEG
retains limited but detectable sequence homology from
bacteria to humans, and is remarkable in its ability to
translocate proteins not only through, but also laterally
into, the membrane (BOX 2), keeping the membrane bar-
rier intact throughout these processes. The purification
of E. coli SecYEG (ecSecYEG) and its functional reconsti-
tution into proteoliposomes
31
was a major breakthrough
that facilitated subsequent biochemical, biophysical and
structural analyses.
In a landmark study, the Rapoport group presented
a 3.2 Å structure of a monomer of SecYEG from the
archaeon Methanococcus jannaschii (mjSecYEG)
5
(FIG. 2a,b). This protein shares approximately 20%
sequence identity with, and contains 3 helices less than,
ecSecYEG. Nevertheless, the architecture of both chan-
nels is thought to be similar
32
and the missing helices
are not essential for ecSecYEG function
33
. M. jannaschii
SecY (mjSecY) is divided into 2 independent halves that
are connected with a loop an N-terminal region
that contains transmembrane helices 1–5 (TM1–TM5)
and a carboxy (C)-terminal region that contains TM6–
TM10. The two subdomains share an inverted pseudo-
symmetry that resembles a clam shell (FIG. 2b, right), with a
narrow hourglass-shaped pore between the subdomains.
TM2 and TM7 are juxtaposed and are thought to form
the lateral gate of the channel (FIG. 2b, right).
The pore in a single SecYEG trimer is blocked from
its periplasmic side by a short loop–helix–loop substruc-
ture that is called the plug (shown in green in FIG. 2b). In
the crystal structure, the plug is in the ‘closed’ state and
SecYEG
YidC
SecDF
YajC
Nature Reviews | Microbiology
FtsY
SecA
Ribosome
ATP ADP
SecB
∆µH
+
+
Periplasm
Cytoplasm
a
b
SPase I
SecYEG
SecA
SRP
SRP-mediated targeting
(co-translational)
1 2 3
SecB-mediated targeting
(co- and post-translational)
Figure 1 | Bacterial protein secretion. a | A schematic representation of the bacterial pre-protein translocase subunits.
The translocase consists of the SecYEG pre-protein-conducting channel (yellow) and the ATPase motor SecA (red).
SecYEG can associate with the auxiliary proteins SecDFYajC and YidC (green)
95,98
. SecA, SecY and SecDFYajC are
known to form higher-order oligomers, but their actual oligomeric state in the functional holoenzyme has not been
conclusively determined. b | A general scheme for the secretion process. Secretory pre-proteins (thick orange line) are
synthesized with amino-terminal signal peptides (orange rectangle) and are targeted to the translocase either by the
ribosome-bound signal-recognition particle (SRP; blue) as soon as they emerge from the ribosome exit tunnel (this is
co-translational translocation, which occurs mainly with membrane proteins and proteins with extremely hydrophobic
signal peptides) or by the tetrameric SecB chaperone (pink) after translation has largely been completed (this is co- and
post-translational translocation, which occurs mainly with secretory proteins) (stage 1). Both targeting routes merge at
the membrane at SecYEG or SecYEG complexed with SecA. Some pre-proteins are targeted by SRP to SecYEG, but if
they have substantial hydrophilic stretches SecA is also required
26
. FtsY (pink) and SecA act as receptors for SRP and
SecB, respectively. For simplicity, SecYEG is presented without any auxiliary subunits. Pre-proteins are translocated
through SecYEG (stage 2) and they fold at the trans side of the membrane or integrate into the lipid bilayer after signal-
peptide cleavage by the signal peptidase
27
(SPase I; magenta) (stage 3). Signal peptides of lipoproteins require cleavage
by SPase II (not shown). Energy is provided by ATP binding and hydrolysis by the SecA ATPase and the proton-motive
force (∆µH
+
). The components are not drawn to scale.
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
841
sits on top of a ring of hydrophobic residues (FIG. 2b) that
forms the narrowest constriction of the hourglass pore
(approximately 5–8 Å). Both the non-essential SecG
(shown in orange in FIG. 2a) and the essential SecE (vio-
let) bind peripherally to SecY, and SecE acts as an external
clamp. This form of the channel appears to be a ‘resting
state, and there is no apparent ion conductance
34
.
The pore within SecY could be the export conduit
for polypeptides
5
. Strong evidence for this comes from
the observation that pre-proteins that stall during trans-
location can be crosslinked to residues in the pore
35
.
Moreover, several prl (protein localization) mutations,
which were isolated in genetic screens and allow the
secretion of polypeptides that contain defective, non-
functional signal sequences
36,37
, map on residues that
are located either on the helices that line the pore or are
on the plug
5
. This distribution led to the proposal that
prl mutations might loosenthe closed state or stabilize a
more open state of the export channel
5
. Signal sequences
could bind at a hydrophobic patch between TM2 and
TM7 (REF. 38), and this could trigger channel opening
6
.
It was proposed that the plug must leave the pore
during transmembrane crossing of the pre-protein
5
.
This was demonstrated directly by biochemical experi-
ments that showed that the plug can flip out towards
the periplasm
39
. Surprisingly, although the plug seems
to be important, and polypeptide translocation is
largely inhibited when the plug is crosslinked in the
closed state
40
, it is not essential for protein transloca-
tion or viability
40,41
. One possible explanation for this is
that in the mutated versions of SecY that lack the plug,
the residues that border the site of the deletion form a
functional pseudoplug
6
. Another concern is that the
~5–8 Å pore diameter is too small to accommodate a
pre-protein chain — having a minimal width of ~12 Å
in an extended conformation or a disulphide-bonded
peptide loop (13 residues long) in a pre-protein
42
or
pre-proteins with covalent attachments
43
. Therefore,
for the single-SecY pore model, it is a prerequisite that
the pore dilates to become functional. This possibility
is supported by computer-based simulations
44
and elec-
trophysiological measurements
34
. The ‘diaphragm-like
dilation of SecY could be facilitated by its several tilted
helices, the topological inversion of SecG
45
, the specific
binding of lipid molecules
46
, the partial detachment of
SecE, and by SecA, through ATP-driven conformational
effects (discussed below).
A single SecYEG trimer is necessary and, by reference
to Occams razor, might be sufficient for translocation.
Interestingly, however, fluorescence resonance energy
transfer (FRET)-based
47
and electron-microscopy (EM)
studies suggest the presence of higher-order oligomers.
By using cryo-electron microscopy
12
,
two-dimensional
8 Å crystals of
ecSecYEG in the lipid bilayer were viewed
and the contours of the 15 predicted TM helices (10 from
SecY, 3 from SecE and 2 from SecG) in the complex
were revealed. In this medium-resolution model,
ecSecYEG is arranged as a dimer of back-to-back’
trimers (FIG. 2c, left). In another cryo-EM study, two
ecSecYEG trimers were shown to be bound to the exit
Box 2 | The sticky problem of membrane proteins
Integral membrane proteins (IMPs) account for approximately 20% of the total proteome of a bacterium such as
Escherichia coli. The key property that allows a protein to be selected for integration into the plasma membrane is the
presence of one or more transmembrane (TM) stretches of approximately 20 hydrophobic residues that form an α-helix
that is long enough (~3 nm) to span the hydrophobic core of the lipid bilayer
126
. Our current understanding of inner-
membrane biogenesis is based on the analysis of only a few model proteins. Membrane integration makes use of the
signal-recognition particle (SRP) and the protein-conducting channel SecYEG, with or without the membrane protein
YidC. In addition, YidC might act both on its own and/or together with the heterotrimeric complex SecDFYajC.
The ribosome-bound SRP identifies TM anchor sequences and long, highly hydrophobic signal peptides. It also targets
the ribosome–nascent chain complex to SecYEG at the membrane mostly co-translationally
3
where it binds to its
membrane-receptor protein (FtsY in bacteria; FIG. 1b). TM domains of IMPs are threaded through SecY during translation
and must eventually leave SecYEG and escape laterally into the lipid bilayer, possibly through the lateral gate of SecY, and
their orientation is partly regulated by phospholipids and the charged amino acids that lie at their termini and the
connecting loops
126–128
. Larger SecYEG assemblies could also facilitate the integration of TM bundles.
YidC can act as an auxiliary component of SecYEG
95–97,129
. Site-specific crosslinking places YidC and lipids in the vicinity
of the TM-anchor sequences of two model IMPs, whereas the hydrophilic regions of these proteins are located near SecA.
Interestingly, a bitopic membrane protein comes close to lipids and SecY before approaching YidC at a later stage of its
elongation
129
. Crosslinking data suggest that YidC is the core of assembly for multiple TMs of polytopic membrane
proteins such as MtlA and MalF, as it was found to be simultaneously attached to more than two consecutive TM segments
in these proteins
97
. YidC is thought to act as an independent insertase for small integral membrane substrates, such as the
Pf3 coat protein
96
and the c subunit of the F
1
F
0
ATP synthase
130
. The mechanism of this insertase function is unknown but it
is speculated that YidC provides a scaffold for the facilitation of the orientation and/orflip flopmovement of the helical
TM domains of newly inserted membrane proteins.
Proteins are also integrated into the outer membrane (OM) of Gram-negative bacteria (and the analogous membranes of
eukaryotic organelles). Proteins that are destined to enter this membrane still use the Sec translocase and, therefore, must
bypass lateral integration into the plasma membrane. To this end, the part of these proteins that crosses the OM in the
integrated state contains characteristic all-β-barrels (with only one known exception the all-α-helical-barrel structured
Wza protein that exports capsular polysaccharides). A proteinaceous machinery is essential for this process. This
machinery consists of the Omp85 OM protein
131,132
, which is evolutionary conserved, as well as a subset of auxiliary
periplasmic chaperones
28
. Omp85 is required for the insertion of many, or all, of the β-barrel OM proteins that assemble
into the OM of E. coli (for a recent review see REF. 133).
RE V IE W S
842
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
channel of a single ribosome complexed with a nascent
polypeptide chain
13
. Using a flexible fitting technique
and the high-resolution mjSecYEG structure
5
, a ‘front-
to-frontarrangement of the two ecSecYEG trimers was
proposed (FIG. 2c, right). In this model, the two almost
juxtaposed lateral gates potentially form a transient
larger pore. This model would be compatible with
fluorescence-quenching analyses of eukaryotic Sec61p,
which are suggestive of the presence of a large, >40 Å,
pore opening during co-translational translocation
48
. In
both cases, allosteric communication between the two
SecYEG trimers could optimize translocation
44
.
From other EM studies of the homologous Sec61p
complex in eukaryotic cells
7–9
, and ecSecYEG
10,11
, it was
proposed that a pore is formed by 2–4 SecYEG trimers.
In eukaryotes, translocation is mainly co-translational
and the Sec61p structures were solved in the presence
of ribosomes
8,9
. Although it cannot be excluded that a
monomeric SecYEG is a peculiarity of M. jannaschii, the
fact that the EM studies were performed in the presence
of ribosomes
7,8
, or SecA and translocating pre-protein
10,11
implies that the recruitment of more than one SecYEG
might be more physiologically relevant. Obviously,
within these SecYEG oligomers, each of the SecYEG
trimers could represent an individual translocation
channel. Alternatively, larger pores might translocate
bulkier substrates or act as enclaves to trap auxiliary
factors. Another interesting possibility was proposed
after a recent crosslinking analysis of SecA-mediated
pre-protein delivery to SecYEG. These results support
the idea that in a SecYEG dimer there is division of
labour: one SecYEG complex provides the pore that is
used for pre-protein transport, whereas the other acts as
a docking station for SecA
49
.
In summary, a single SecYEG trimer can form an
export pore for exiting polypeptides. However, SecYEG
complexes also form dimeric or higher-order assemblies.
These might facilitate the docking of other translocase
subunits and the ribosome or have other roles that
remain poorly understood.
The SecA ATPase. SecA provides the chemo–mechanical
energy conversion that is required for translocation. The
domain organization of this molecular motor has been
revealed by biochemical analyses
50–54
and by the crystal-
lographic structures of SecA from Bacillus subtilis
14–16
,
Mycobacterium tuberculosis
18
, Thermus thermophilus
19
and E. coli
17
.
Each SecA protomer contains a central core — the
DEAD (DEAD-box) or helicase motor (FIG. 3a). This is
structurally homologous to the corresponding domains
in DEAD and DExD/H superfamily II proteins, which
include helicases (BOX 3). DEAD motors do not share
any significant primary-sequence similarity, other
than nine motifs, and are formed by two RecA-like fold
domains (BOX 3). In SecA, these domains are called NBD
(nucleotide-binding domain) and IRA2 (intramolecular
regulator of ATPase 2)
51,53
(FIG. 3b). A single nucleotide
binds at the NBD–IRA2 interface, which is lined by heli
-
case motifs that have been suggested to link ATP bind-
ing and hydrolysis to translocation
52,55
(BOX 3). Nuclear
magnetic resonance (NMR) spectra have revealed that
the DEAD motor is dynamic and that the C domain
partially suppresses this inherent flexibility
55
.
DEAD motors acquire specificity for different sub-
strates from non-homologous substrate-specificity
domains
56
. Such domains are usually fused at the N- or
C-terminal region of the DEAD motor or they protrude
from internal loops without disturbing the overall motor
structure (BOX 3). SecA has two specificity domains
(FIG. 3a–d): the C domain and the pre-protein-binding
domain (PBD). The α-helical C domain is fused at
the C terminus of IRA2 and contains 4 substructures
Nature Reviews | Microbiology
a
b
c
Periplasm
Cytoplasm
Periplasm
Cytoplasm
Bottom Front Side Top
Front Front Side
‘Back-to-back’ ‘Front-to-front’
Top
TM4
TM7
TM2
Plug
TM6–TM10
TM1–TM5
Figure 2 | Structure of the SecYEG protein-conducting channel. a | A space-filling
model of SecYEG from Methanococcus jannaschii (mjSecYEG; Protein Data Bank code:
1RH5)
5
. Four views are presented: bottom (from the cytoplasm); front; side; and top (from
the periplasm). SecY is in grey, SecE is in violet and SecG is in orange. In the bottom and
top representations the plug is removed to visualize the putative translocation pore.
b | A front-view space-filling model and ribbon diagrams of mjSecYEG
5
. The colours for
SecY, SecE and SecG are the same as in a, except that transmembrane helices 2 and 7
(TM2 and TM7), which are proposed to form the lateral gate of the SecY channel, are
coloured dark blue and red, respectively. TM4 is in cyan and the plug is green. In the
structure on the extreme right a dotted line reveals the pseudosymmetry of the two
subdomains of SecY (TM1–TM5 and TM6–TM10). c | A schematic representation of two of
the proposed modes of SecYEG dimerization: ‘back-to-back(REF.12) andfront-to-
front(REF.13). The two SecYEG trimers (dark and light grey space-filling models) were
arranged manually to allow visualization of the two proposed arrangements. A top view
is shown (as in b; right). The last TM helices of SecE (purple cylinders), SecG (orange;
shown as a filled circle in this view), and TM2, TM7 and TM4 of SecY (dark blue, red and
cyan cylinders, respectively) are presented. In the back-to-back arrangement, the second
TM helices of the two SecEs are juxtaposed and the lateral TM2–TM7 gates of SecY face
outwards (left), whereas in the front-to-front arrangement they are proximal to one
another (right). The presumed protein-conducting pore is in white.
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
843
(FIG. 3b): the scaffold domain, which is a 46-residue
helix that acts as a molecular staple to control the open-
ing and closing of the DEAD motor
14,51,55,57
; the wing
domain (WD); the conserved helix–loop–helix IRA1
switch; and a short region of low conservation called the
C-terminal domain (CTD) or C tail, which is located at
the extreme C terminus and is not essential for cataly-
sis
58
. The structure of only the first half of the CTD in
one SecA structure has been determined by crystallog-
raphy
14
; its extreme 25 residues have been shown to bind
zinc ions by NMR
59,60
and biochemical studies
61
. The
PBD protrudes from the NBD between helicase motifs
II and III, binds pre-proteins
52
(BOX 3) and contains an
anti-parallel β-strand (stem) and a bilobate globular
domain (bulb) (FIG. 3c). Residues in (or from) the stem
and bulb appear to bind pre-protein segments
50,52,62
, with
the signal peptide binding mainly in a shallow groove
in the bulb
63
.
In the available SecA structures, the PBD bulb under-
goes a remarkable approximately 75° swivelling motion
between a closed (FIG. 3c) and an open (FIG. 3d) state. In
the open state, a large cleft is formed between the WD,
IRA1 and bulb structures. Part of this opening could
accommodate pre-proteins or regions of SecYEG
15,17
.
Order–disorder transitions of the DEAD motor are
allosterically transmitted to the PBD
55
. If pre-proteins
are bound, the PBD ‘talks backconformationally to the
DEAD motor
64
. Importantly, the two specificity domains
bind to one another, with the major contacts being pro-
vided by both IRA1 and the PBD bulb
65
. These physical
contacts ensure allosteric communication between these
domains, and between these domains and the DEAD
motor. This conformational flow lies at the core of
SecA-mediated catalysis
55,64,65
.
SecA oligomerization. SecA exists in solution in a
monomer to dimer equilibrium (the estimated equilib-
rium dissociation constant (K
d
) is approximately 1 µM
at 8°C in the presence of 300 mM K
+
acetate
66
) and, in
concentrated solutions, SecA is routinely isolated as a
dimer
51,66–69
. SecA is thought to partition equally between
the cytoplasm and the membrane. At the concentra-
tions that are believed to be found in the cell cytoplasm
(2–5 µM
66,69,70
), ecSecA is expected to be mainly dimeric.
Remarkably, the molecular determinants of SecA dimer
-
ization remain a conundrum. Despite the fact that all
the SecA protomers in the crystal structures are similar, the
dimers of the various SecAs present different dimeric
interfaces
14,16–19
(FIG. 3e). Moreover, B. subtilis SecA
crystallized as an apparent monomer in one study
15
, in
another it formed symmetrically related crystallographic
dimers
14
and in yet another it formed dimers in the
crystal unit cell
16
. Biochemical analysis of the different
domains of SecA favours possibilities for assembly
51,71,72
that are incompatible with most of the structures. For
example, an isolated C-domain polypeptide is stably
dimeric
51,72
. However, in none of the crystal structures is
the C domain used as an exclusive dimerization interface.
Finally, truncation of nine extreme N-terminal residues
that were thought to be important for dimerization only
in the symmetrically related dimeric bsSecA structure
14,15
C domain
Bulb
‘Closed’ ‘Open’
Stem
DEAD
motor
NBD
CTD
WD
IRA1
IRA2
PBD
PBD
SD
a b
c d
e
E. coli
B. subtilis I B. subtilis II
T. thermophilus M. tuberculosis
Figure 3 | The structure of SecA. Space-filling (a) and ribbon-representation (b)
models of the Escherichia coli SecA protomer (ecSecA) with its pre-protein-binding
domain (PBD) (modelled into Protein Data Bank (PDB) code 2FSF
as described in
REF. 17). The DEAD-motor domains are dark blue (nucleotide-binding domain
(NBD); residues 1–220 and 377–416) and mauve (intramolecular regulator of
ATPase 2 (IRA2); residues 417–621), the C domain (residues 622–901) is green and
the PBD (residues 221–377) is magenta. In b, the sub-structures of the C domain
are indicated: SD, scaffold domain (dark green); WD, wing domain (light green);
IRA1 (yellow); CTD, carboxy-terminal domain (orange). The swivelling of the bulb
sub-structure of the PBD in ecSecA from the ‘closed’ to the ‘open’ state is shown
against a space-filling model of the C domain (c and d). The closed state of ecSecA
was derived from structural modelling using Bacillus subtilis SecA (bsSecA)
14
as a
template (PDB code: 1M6N). The ‘open’ state was derived from the crystallographi-
cally determined structure
17
. In ecSecA, the stem comprises residues 221–227 and
370–377 and the bulb comprises residues 228–269. A comparison of SecA dimers is
also provided (e). Ribbon representation models are shown in which one protomer
is coloured as in a and the other is grey. The coloured protomer of each SecA
(shown in a top view looking down on the DEAD motor) was structurally aligned to
visualize the different orientations of the second protomer. The PDB codes that were
used are: B. subtilis SecA I (bsSecA I), 2BIM
16
; B. subtilis SecA II (bsSecA II), 1M6N
14
;
ecSecA, 2FSF
17
; Mycobacterium tuberculosis SecA (mtSecA), 1NL3 (REF. 18) and Thermus
thermophilus SecA (ttSecA), 2IPC
19
.
RE V IE W S
844
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
Box 3 | Superfamily II helicases
SecA is a member of the superfamily II (SF2) DExH/D proteins. Most of the members of
this family are nucleic acid helicases
56
. Several other functions, such as the disassembly
of protein–RNA complexes and stabilization of RNA conformations, have been described.
Helicases use the energy from ATP hydrolysis to bind and unwind nucleic acid substrates,
whereas SecA uses this energy to bind and move along polypeptide substrates. As shown
in the figure all helicases share a structurally highly conserved domain that contains two
RecA-like subdomains that form the helicase or DEAD (Asp-Glu-Ala-Asp) motor. The
motor comprises nine conserved motifs, many of which line the ATP-binding cleft.
Several of these motifs are directly involved in nucleotide binding, whereas others
appear to be important in conveying conformational cross-talk with the substrates.
In the figure, panel a shows a ribbon diagram of the Escherichia coli SecA (ecSecA) DEAD
motor bound to ATP
17
. The helicase motifs are coloured according to panel b. Only the
beginning of the stem of the pre-protein-binding domain (PBD) is visible (pink). The side
chains of the residues discussed in the main text are shown in red. Gate 1 comprises D217
and R566, joined by a salt bridge
64
. Panel b shows a schematic representation of the nine
conserved motifs of SF2 DEAD helicases aligned with those of the ecSecA DEAD motor.
The site of insertion of the PBD is indicated, as well as the number of the first and the last
residue of each motif in SecA. Capital letters indicate more than 80% homology; lower-
case letters indicate 50–79% conservation. The ‘o’ represents threonine or serine. IRA2,
intramolecular regulator of ATPase2; NBD, nucleotide-binding domain; Q, glutamine.
did not affect ecSecA dimerization and function
73,74
(for
an alternative view, see REF. 75). However, truncation of
the nonapeptide could disfavour the formation of one
dimeric state but promote the formation of an alterna-
tive dimeric state. One possibility is that the revealed
dimerization interfaces correspond to different SecA
dimers that have specialized functions. Another possibil-
ity is that the extreme conditions used for crystallization,
such as elevated salt concentration, dissociate the SecA
dimer and favour different but physiologically irrelevant
arrangements of the protomers in the crystal lattice. In a
remarkable recent breakthrough, the Kalodimos labora-
tory has acquired a solution structure of the full-length
dimeric SecA, a protein of 204 kDa, using NMR spec
-
troscopy
63
. NMR studies will provide direct information
on the native state of dimeric SecA in solution.
The SecA–SecYEG holoenzyme. SecYEG and SecA form
the active holoenzyme
31
. SecA binds to the membrane with
low affinity at acidic phospholipids but with high affinity
(5–40 nM) at SecYEG
24
. Several cytoplasm-exposed loops
of SecY are available for a possible interaction with SecA
and/or the ribosome
5,12,13,32,76,77
(FIG. 2a,b). SecYEG makes
three connections with the ribosome two between
ribosomal-RNA hairpins and the cytoplasmic SecY loops
and one between ribosomal proteins and the cytoplasmic
region of SecG
13
. Potential SecA-interacting residues in
the SecY cytoplasmic loops have also been identified
genetically, and biochemical data suggest that TM4 of
SecY, which is largely inaccessible from the cytoplasm,
might interact with SecA
77
(FIG. 2b,c). SecG, although not
essential for the initial binding of SecA to SecYEG, might
interact at a later stage of the translocation reaction
78,79
.
Several SecA domains might be important for binding to
SecYEG
65,80,81
. The DEAD motor was shown by biochemi-
cal methods to interact with SecYEG
49,65,80
. As the affinity
of wild-type SecA is four times higher than that of a SecA
that is missing its C domain
65,81
, the C domain might
also contribute to binding either directly or by allosteric
modulation of the DEAD motor.
The stoichiometry of the interaction between SecA
and SecYEG is largely unresolved. Several experi-
ments point to dimeric SecA being the active state
of the enzyme
67,82,83
. Nevertheless, the crystallization of
monomeric SecA
15
raised the possibility that SecA acts as
a monomer. The results of FRET and detergent-induced
monomerization experiments were interpreted as an
indication that SecA forms monomers during lipid bind-
ing
15,75,84–86
. In addition, SecA can bind to a single SecYEG
that is trapped in Nanodisc particles
46
. Nanodisc-
trapped SecA acquires an unusually elevated (several
fold) ATPase activity in the absence of a translocating
pre-protein. This is not detected when SecA is bound
to membrane-embedded SecYEG
64,87
. It has not been
reported whether Nanodisc-trapped SecA is still active
for the pre-protein interaction and translocation under
the same conditions. Monomeric SecA can accommo-
date two SecYEG trimers if it binds with its open PBD
conformation
15,49
(FIG. 4c). In other studies, all possible
combinations of SecA monomers and dimers that could
interact with SecYEG monomers or dimers have been
proposed
10,11,85,88
. Although a single SecA appears to be
competent in binding to a single SecY, the data that have
been obtained with low-resolution methods are not suf-
ficient to provide strong support for either combination
of SecA–SecY oligomeric assembly. The dimensions of
both SecYEG dimers
13,89
(FIG. 2c) are comparable to those
of both the monomeric and dimeric forms of SecA. As
an example, three possible arrangements of SecA bound
to dimeric SecY are shown in FIG. 4 (monomeric (part a)
and dimeric (parts b and c)). Addressing the assembly and
disassembly of a dynamic system at a membrane interface
is technically notoriously difficult.
Auxiliary components of the translocase. Another het-
erotrimeric protein complex called SecDFYajC has also
been found to be associated with the core SecYEG trimer
at the membrane
90
(FIG. 1a) under certain purification
Nature Reviews | Microbiology
NBD
NBD
ATP
Q
I
Ia
R509
R566
D217
Gate 1
PBD
PBD
IRA2
R574
V
IV
VI
Ib
III
II
GG
SF2 DEAD
SECA E. coli
Motifs
Q GG TPGR1 VlDEaDxm SAT liFxxo LvaTdvaaRGlD YxHRi GRogRxGAxoGoGKT PTRELAxQ
Q PG TNNEY LVDEVDSI TGT LVGTIS TIATNMAGRGTD ESRRIDNQLRGRSGRQGMKTGEGKT VNDYLAQR
Q I Ia GG Ib II III IV V VI
87
102
109
131
138
159
160
178
182
207
214
391
393
453
458
501
512
563
579
IRA2
C domain
a
b
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
845
conditions. This complex is not required for the in vitro
reconstitution of the translocation reaction, but has a
stimulatory effect on translocation activity in the absence
of SecG
90
. However, viability and protein export are seri-
ously compromised in vivo when cells are depleted of
SecD or SecF
91
, although the same is not true for YajC.
SecDFYajC might be involved in the release of translo-
cated proteins from the membrane
92
or the regulation of
SecA membrane cycling
93
. Another membrane protein,
YidC (FIG. 1a), is involved in the biogenesis of both the Sec-
dependent and Sec-independent membrane proteins
94–96
,
and associates with SecDFYajC
97
(BOX 2).
The translocase binds multiple ligands
Translocase and phospholipids. Anionic phospholipids
are a significant component of the E. coli membrane
(20% phosphatidylglycerol (PG) and 5% cardiolipin
(CL)) and are necessary for protein secretion both in vivo
and in vitro
87,98
. Long-chain phospholipid analogues and
acidic phospholipids were shown to promote SecA mon-
omerization
84,86
. Acidic phospholipids bound at SecYEG
contribute to the binding affinity of SecA and the subse-
quent activation of its ATPase activity
46,87
, and can even
restore the activity of cold-sensitive SecA mutants
99
. In
contrast to PG, which gives rise to organized bilayers,
phosphatidylethanolamine (PE) the major lipid com-
ponent (75%) of the E. coli membrane — cannot form
bilayer structures in vitro
98
. In the absence of PE, mem-
branes are functional for translocation only if they have
highly increased amounts of PG and CL, which, together
with divalent cations, promote non-bilayer structures
that are similar to those promoted by PE
100
.
The pre-protein–SecB complex. SecB is a general cellular
chaperone
20
that appears to be important for the export of
several secretory proteins, such as the outer-membrane
porin LamB
4
. It is not essential for viability and is not
present in all bacteria, but has a dual role: it maintains
secretory pre-proteins in a translocation-competent state
and interacts specifically with membrane-bound SecA
(K
d
= 10–30 nM)
24
.
SecB crystal structures
101–103
revealed structural fea-
tures that could allow this chaperone to bind to extended
polypeptides and SecA. SecB is a stable tetramer (FIG. 5a)
that probably binds to pre-proteins by recognizing
exposed hydrophobic surfaces
104
. A groove has been
identified in the structures that could constitute the
pre-protein binding site. Each SecB
4
contains two such
grooves, which, as these sites are mainly solvent exposed,
could accommodate peptides without disturbing the
stable tetramer (FIG. 5c, pink cylinder). The grooves are
lined by several aromatic residues at one end and hydro-
phobic, but not aromatic, residues at the opposing end
(FIG. 5c, part 2). The aromatic-residue-rich sub-site of SecB
could accommodate the proposed SecB-binding motif
in pre-proteins, which, it is suggested, is a nonapeptide
that contains several aromatic and basic residues
105
. In
an acidic region on the top surface of the SecB tetramer,
several residues interact with the 25 C-terminal residues
of the SecA CTD
106,107
(FIG. 5b). The symmetry of SecB
4
(a dimer of dimers) favours the binding of 1 SecA dimer
per SecB
4
(FIG. 5b). Site-directed spin labelling in combi-
nation with electron paramagnetic resonance spectros-
copy was used to investigate how SecA and a pre-protein
docks on SecB
108
and the results partially confirmed the
sites that had already been proposed by crystallography
(FIG. 5b,c). Several routes for how the polypeptide wraps
its extended aminoacyl chain around SecB have been
proposed, as would be expected for a chaperone that
binds to numerous diverse substrates.
SecA–pre-protein complex. When SecA recognizes the
SecB-delivered pre-protein, it releases SecB in an ATP-
regulated reaction
107
. These events are thought to occur
at the SecYEG-bound form of SecA. However, some
evidence suggests that pre-protein–SecA complexes
might also form at the ribosome exit site
109,110
. SecA rec-
ognizes pre-proteins, at least in part, through its PBD
region
50,52,63
(FIG. 3a). The hydrophobic core of the signal
peptide acquires an α-helical conformation after binding
to SecA
111
. Signal peptides bind at an interfacial groove
between the PBD bulb and the IRA1 hairpin of the C
domain
63
. As part of the CTD occupies this groove
63
, this
could provide a mechanism to shield this region from
the premature binding of signal peptides to cytoplasmic
SecA. Mutational
112,113
and biochemical
52,114
studies with
a model mature domain suggest that SecA might bind
mature pre-protein segments or regions through its bulb
region.
SecA–nucleotide binding. The NBD of SecA contains
a high-affinity nucleotide-binding site (see motifs in
BOX 3). Most of the residues that contact the nucleotide
directly are located in the NBD. A few IRA2 residues
(BOX 3) directly contact ATP in the ecSecA crystal struc-
ture
17
. The putative arginine finger in IRA2 R574 is
indispensable for catalysis
53,55
, and faces away from the
pocket (BOX 3), implying that ecSecA crystallized in an
inactive state. However, it is clear from NMR studies of
ecSecA that the arginine fingers are important for nucle-
otide binding and for the nucleotide-dependent stabi-
lization of IRA2
(REF. 55). Other interactions between
Nature Reviews | Microbiology
~6.7 nm
~5 nm
~8.7 nm
Cytoplasm
Periplasm
a b c
SecYEG
Figure 4 | Hypothetical SecA–SecYEG interaction models. The approximate
dimensions for the SecYEG dimer (yellow) and the lipid bilayer are indicated. Hypothetical
interactions with a monomeric SecA (Bacillus subtilis SecA)
15
(a) and two different dimeric
SecAs (Mycobacterium tuberculosis SecA
18
(b) and Escherichia coli SecA (ecSecA)
17
(which
is approximately 15 nm x 8 nm) (c), are shown as examples. The ecSecA dimer
17
is the only
one in which both DEAD motors face the same side of the dimer and would thus both be
available for binding to a SecYEG dimer. In SecA, the DEAD-motor domains are dark blue
and mauve, the C domain is green and the pre-protein-binding domain is magenta.
RE V IE W S
846
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
conserved residues that reside in opposing sites of the
motor ensure nucleotide-independent intra-domain
communication between NBD and IRA2
(REFS 14,17,65).
The crystal structures did not reveal major nucleotide-
driven conformational changes
14,15,17,18
. However,
nucleotides and other translocation ligands regulate the
equilibrium between the ordered and disordered states of
the DEAD motor
55
. In addition, a highly conserved salt
bridge (see gate 1 in BOX 3) controls the catalytic cycle
of SecA by regulating the opening and closing of the
nucleotide cleft
64
and is under the control of the PBD.
All of the above interactions between the trans-
locase and its peripherally associated ligands must be
ordered and sequential, and energy must be spent during
the ensuing step-wise mechanical work. The process of
energy generation is discussed in the next section.
Energizing the translocase machine
Stage two of the translocation reaction (FIG. 1b, part 2)
requires energy input in the form of both ATP and the
PMF. These energy sources function at different stages
of translocation
42,115
. ATP is essential for the initiation of
the translocation reaction
115
. Cytosolic SecA rapidly
hydrolyses ATP to ADP
51
. SecAADP is thermally stabi-
lized
53,116
and the release of ADP from the DEAD motor
is the rate-limiting step for the subsequent catalysis
53,116
.
SecA increases its ATP turnover in a pre-protein- and
SecYEG-dependent manner. This stimulated activity
is known as the translocation ATPase activity
64,87
. SecA
undergoes insertion and de-insertion cycles at SecYEG
during its catalytic cycle
93,117
. The inserted state is a con-
formation in which significant parts of SecA become
membrane integral and inaccessible to proteases. The
ATP- and ADP-bound states are thought to correspond
to inter-convertible extended and compact conforma-
tions that are able to insert and de-insert at the mem-
brane, respectively. The energy that is released during
these states is converted to mechanical work so that
pre-protein segments are inserted into the SecYEG pore
and eventually become translocated to the trans side of
the membrane after several SecA cycles of insertion and
de-insertion. When pre-protein translocation is substan
-
tially underway, the PMF can further drive the reaction
and even complete it in the apparent absence of ATP
115
.
A high concentration of SecA seems to render transloca-
tion largely PMF-independent
118
. Although the role of
the PMF during protein translocation is not completely
clear, it seems to promote SecA de-insertion either by
stimulating the release of ADP from the motor
119
or
by stimulating conformational changes in SecY
120,121
. In
the absence of a PMF and SecA, pre-proteins can slip
back towards the cis side of the membrane, which can
lead to uncoupled ATP consumption during translo-
cation
115
, and thus the PMF prevents back-slippage of
the translocating pre-protein
115,122
. Finally, some of the
energy that is released during export might be directed
towards active unfolding of pre-protein domains
122
.
Pre-protein translocation by the Sec translocase
Using the available information, we propose a mecha-
nistic model that describes SecA-dependent bacterial
pre-protein translocation in twelve stages (FIG. 6).
In the first, quiescent stage, SecA resides in the cyto-
plasm in its inactive and stable ADP-bound state
53,116
.
This is followed by a priming stage (stage 2), during
which SecA–ADP binds to SecYEG, mainly through
the DEAD motor
49,65,81
, and probably to the cytosolic
loops of SecY
5,76,77
. Priming leads to conformational
changes in the SecA C domain and, possibly, the partial
Figure 5 | The structure of SecB. a | A ribbon
representation of the Escherichia coli SecB (ecSecB)
tetramer, which is a dimer of dimers
103
(Protein Data Bank
code: 1QYN). The two dimers are green–yellow and
red–dark blue. b The SecB–SecA interaction. A ribbon and
space-filling model representation of the ecSecB tetramer
(grey)
103
in complex with the carboxy (C)-terminal peptide
of Haemophilus influenzae SecA (yellow)
101,102
is shown.
Two SecA peptides are in complex with one SecB tetramer.
The surface that binds the SecA C-terminal peptide is in
red. Other SecB residues that seem to interact with
SecA are in green
108
. c | The SecB–pre-protein complex.
A ribbon representation of the ecSecB
4
tetramer is shown
(protomers a–b, dark grey; c–d light grey) (1). The pink
cylinder marks the proposed peptide-binding channel that
is formed by two grooves on the same side of the tetramer
(a–d or b–c). A space-filling model of 1 is shown in 2, with
hydrophobic residues that line the 4 grooves (green) and
residues that seem to interact with a pre-protein indicated
(galactose-binding protein precursor; red)
134
. A space-
filling model of 1 is also shown in 3, but only protomers a
and d are shown so that the continuity of 1 peptide-
binding channel can be viewed. Each tetramer contains
two such channels.
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
847
detachment of the scaffold domain
55,57,65
and conforma-
tional changes in the PBD
63,64
. These changes result in
moderate membrane ATPase activity and, importantly,
favour high-affinity binding of the SecB–pre-protein
complex
24,87
to the CTD
102,107
and/or the N-terminal
106,108
regions of primed SecA. In the third, triggering stage
(stage 3), SecA binds the signal peptide and the N-terminal
regions of the mature pre-protein through the PBD
50,52,62
and, possibly, the C domain. This binding alleviates
C-domain suppression and opens the gate 1 salt bridge at
the base of the DEAD motor
64
(BOX 3). As a result, IRA2
becomes disordered and detaches from the NBD, which
leads to loss of affinity of the DEAD motor for ADP
(stage 4)
51,53,55,64,65
. The loss of ADP drives SecA to the
pre-insertion stage (stage 5) in which the PBD, together
with the bound pre-protein, undergoes conformational
changes
52,55
and SecB is released from SecA
107
.
ATP binding to the empty DEAD motor (the ATP
stroke stage; stage 6) drives additional conformational
changes, and regions of SecAATP penetrate deeper
into the membrane plane at SecYEG
93,117
. Such regions
of SecA could include the two specificity domains the
PBD and the C domain
117
as well as the DEAD motor.
SecA membrane insertion could have two effects. On the
one hand it might prepare the SecYEG channel by initi-
ating the dilation of the SecY ‘diaphragm(stage 7). On
the other hand, however, it might deliver the pre-protein
to the primed channel. The release of the signal peptide to
the lateral gate of SecY, where it becomes stably bound
by its hydrophobic core
35,38,49
, causes the plug to flip out
away from the pore, thereby leading to a fully activated
channel (stage 8)
6,34,39
. Signal-peptide penetration in the
membrane drags along the N-terminal mature region
of the pre-protein
115
, and the hairpin that it forms
123
becomes positioned at SecY in such a way that the mature
region is threaded into the open SecY pore
35,49,124
. ATP
then becomes hydrolysed and the pre-protein is partially
released into the channel
49,115
(pre-protein-release stage;
stage 9). SecA returns to its previous ADP-bound state in
a reaction that is facilitated by the PMF
119
, the C domain
re-associates with the DEAD motor and the PBD is free
to move further down the pre-protein chain (de-inserted
state; stage 10)
93,117
. The attachment of the PBD onto the
succeeding pre-protein segment (stage 11), will therefore
trigger ADP release and a new round of catalysis (stages
4–11). The open and closed PBD states that were detected
in the crystal structures (FIG. 3c,d) and by NMR
63
might
correspond to pre-protein binding and release confor-
mations. SecA must perform multiple catalytic rounds,
alternating between the SecA–ADP and SecA–ATP
states and the conformational states of the holoenzyme
that they represent, in order for the whole length of the
polymeric pre-protein substrate to cross the channel in
segments of approximately 30 residues at a time
115,124
.
Several repeats of these events (stages 4–11) would lead
to the translocation of a complete pre-protein chain, with
SecA functioning as a typical processive motor. The cata-
lytic cycle is facilitated by the PMF, which can drive the
completion of translocation when the pre-protein is not
attached to SecA
115
. Finally, at the maturation stage (stage
12), the mature polypeptide chain, which is generated
after signal-peptide cleavage by signal peptidase, acquires
Cytoplasm
Periplasm
Nature Reviews | Microbiolog
y
D
D
SecB
SecA
1 2 3 4 5 6 7
9 10 11
12
8
D D
T
T
D D
D
T
D
Plug
Figure 6 | SecA-dependent pre-protein translocation — a model for the sequence of events. See main text for
details. For simplicity, SecA and SecY are shown as monomers and the carboxyl terminus of SecA, where SecB is thought to
bind, is only shown in stages 1–4 and in stage 12. Only the initial stages of the process are shown, with a pre-protein
containing a signal peptide. However, similar multiple catalytic turnovers are expected to drive the complete secretion of
the mature region after signal-peptide cleavage. Irregular contours indicate conformational changes. The thick orange line
represents the secretory pre-protein and the orange rectangle represents the N-terminal signal peptide. D, ADP; T, ATP.
RE V IE W S
848
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
its native three-dimensional structure at the trans side
of the membrane.
Conclusions
Despite recent progress in our understanding of the
mechanisms that are involved in bacterial protein secre-
tion, many fundamental questions remain and several
aspects of the catalytic cycle (FIG. 6) are unresolved. How
does the channel open to accommodate a passing pre-
protein? Which of the oligomeric states of SecYEG
and SecA defines an active channel? What are the
conformational changes that follow SecA–pre-protein-
complex binding to SecYEG and trigger translocation
through the channel? What is the role of the SecA-specificity
domains during the translocation reaction? What is the
role of the PMF? What is the molecular basis of SecA
processivity? To address these and other questions, high-
resolution structures of complexes must be determined
and biophysical real-time experiments, such as single-
molecule studies, should be performed. Bacterial protein
secretion remains an exciting research area and a unique
paradigm for macromolecular membrane transport.
1. Economou, A. et al. Secretion by numbers: protein
traffic in prokaryotes. Mol. Microbiol. 62, 308–319
(2006).
2. Holland, I. B. Translocation of bacterial proteins — an
overview. Biochim. Biophys. Acta 1694, 5–16 (2004).
3.
Luirink, J. & Sinning, I. SRP-mediated protein
targeting: structure and function revisited. Biochim.
Biophys. Acta 1694, 17–35 (2004).
4. Randall, L. L. & Hardy, S. J. SecB, one small
chaperone in the complex milieu of the cell. Cell. Mol.
Life Sci. 59, 1617–1623 (2002).
5.
Van den Berg, B. et al. X-ray structure of a protein-
conducting channel. Nature 427, 36–44 (2004).
The first crystal structure of the membrane-
embedded pre-protein-conducting channel. It
provided hypotheses for the channel-dilation
mechanism and its gating.
6.
Li, W. et al. The plug domain of the SecY protein
stabilizes the closed state of the translocation channel
and maintains a membrane seal. Mol. Cell 26,
511–521 (2007).
7.
Hanein, D. et al. Oligomeric rings of the Sec61p
complex induced by ligands required for protein
translocation. Cell 87, 721–732 (1996).
8.
Beckmann, R. et al. Architecture of the protein-
conducting channel associated with the translating
80S ribosome. Cell 107, 361–372 (2001).
9.
Beckmann, R. et al. Alignment of conduits for the
nascent polypeptide chain in the ribosome-Sec61
complex. Science 278, 2123–2126 (1997).
10. Manting, E. H., van Der Does, C., Remigy, H., Engel, A.
& Driessen, A. J. SecYEG assembles into a tetramer to
form the active protein translocation channel. EMBO J.
19, 852–861 (2000).
11.
Scheuring, J. et al. The oligomeric distribution of
SecYEG is altered by SecA and translocation ligands.
J. Mol. Biol.
354, 258–271 (2005).
12. Breyton, C., Haase, W., Rapoport, T. A., Kuhlbrandt, W.
& Collinson, I. Three-dimensional structure of the
bacterial protein-translocation complex SecYEG.
Nature 418, 662–665 (2002).
13.
Mitra, K. et al. Structure of the E. coli protein-
conducting channel bound to a translating ribosome.
Nature 438, 318–324 (2005).
14. Hunt, J. F. et al.
Nucleotide control of interdomain
interactions in the conformational reaction cycle of
SecA. Science 297, 2018–2026 (2002).
The first crystal structure of the SecA motor
subunit of the translocase. Provided high-resolution
information for the structural and functional
dissection of the motor.
15. Osborne, A. R., Clemons, W. M. Jr & Rapoport, T. A.
A large conformational change of the translocation
ATPase SecA. Proc. Natl Acad. Sci. USA 101,
10937–10942 (2004).
16. Zimmer, J., Li, W. & Rapoport, T. A. A novel dimer
interface and conformational changes revealed by an
X-ray structure of B. subtilis SecA. J. Mol. Biol.
364,
259–265 (2006).
17.
Papanikolau, Y. et al. Structure of dimeric SecA, the
Escherichia coli preprotein translocase motor. J. Mol.
Biol. 366, 1545–1557 (2007).
18.
Sharma, V. et al. Crystal structure of Mycobacterium
tuberculosis SecA, a preprotein translocating ATPase.
Proc. Natl Acad. Sci. USA 100, 2243–2248 (2003).
19. Vassylyev, D. G. et al.
Crystal structure of the
translocation ATPase SecA from Thermus
thermophilus reveals a parallel, head-to-head dimer.
J. Mol. Biol.
364, 248–258 (2006).
20. Ullers, R. S. et al.
SecB is a bona fide generalized
chaperone in Escherichia coli. Proc. Natl Acad. Sci.
USA 101, 7583–7588 (2004).
21. Schierle, C. F. et al.
The DsbA signal sequence directs
efficient, cotranslational export of passenger proteins
to the Escherichia coli periplasm via the signal
recognition particle pathway. J. Bacteriol. 185,
5706–5713 (2003).
22.
Sijbrandi, R. et al. Signal recognition particle (SRP)-
mediated targeting and Sec-dependent translocation
of an extracellular Escherichia coli protein. J. Biol.
Chem. 278, 4654–4659 (2003).
23. Randall, L. L. Translocation of domains of nascent
periplasmic proteins across the cytoplasmic
membrane is independent of elongation. Cell 33,
231–240 (1983).
24. Hartl, F. U., Lecker, S., Schiebel, E., Hendrick, J. P. &
Wickner, W. The binding cascade of SecB to SecA to
SecY/E mediates preprotein targeting to the E. coli
plasma membrane. Cell 63, 269–279 (1990).
25. Scotti, P. A. et al.
SecA is not required for signal
recognition particle-mediated targeting and initial
membrane insertion of a nascent inner membrane
protein. J. Biol. Chem. 274, 29883–29888 (1999).
26.
Neumann-Haefelin, C., Schafer, U., Muller, M. & Koch,
H. G. SRP-dependent co-translational targeting and
SecA-dependent translocation analyzed as individual
steps in the export of a bacterial protein. EMBO J. 19,
6419–6426 (2000).
27. Paetzel, M., Karla, A., Strynadka, N. C. & Dalbey, R. E.
Signal peptidases. Chem. Rev. 102, 4549–4580
(2002).
28. Mogensen, J. E. & Otzen, D. E. Interactions between
folding factors and bacterial outer membrane proteins.
Mol. Microbiol. 57, 326–346 (2005).
29. Nakamoto, H. & Bardwell, J. C. Catalysis of disulfide
bond formation and isomerization in the Escherichia
coli periplasm. Biochim. Biophys. Acta 1694,
111–119 (2004).
30. Gruber, C. W., Cemazar, M., Heras, B., Martin, J. L. &
Craik, D. J. Protein disulfide isomerase: the structure
of oxidative folding. Trends Biochem. Sci. 31,
455–464 (2006).
31. Brundage, L., Hendrick, J. P., Schiebel, E., Driessen,
A. J. & Wickner, W. The purified E. coli
integral
membrane protein SecY/E is sufficient for
reconstitution of SecA-dependent precursor protein
translocation. Cell 62, 649–657 (1990).
In vitro functional reconstitution of the translocase
from purified components. This study identified
the components of the core channel and opened the
floodgates for subsequent biochemical and
biophysical studies.
32.
Bostina, M., Mohsin, B., Kuhlbrandt, W. & Collinson, I.
Atomic model of the E. coli membrane-bound protein
translocation complex SecYEG. J. Mol. Biol.
352,
1035–1043 (2005).
33. Schatz, P. J., Bieker, K. L., Ottemann, K. M., Silhavy,
T. J. & Beckwith, J. One of three transmembrane
stretches is sufficient for the functioning of the SecE
protein, a membrane component of the E. coli
secretion machinery. EMBO J. 10, 1749–1757
(1991).
34. Saparov, S. M. et al.
Determining the conductance of
the SecY protein translocation channel for small
molecules. Mol. Cell 26, 501–509 (2007).
35. Cannon, K. S., Or, E., Clemons, W. M. Jr, Shibata, Y. &
Rapoport, T. A. Disulfide bridge formation between
SecY and a translocating polypeptide localizes the
translocation pore to the center of SecY. J. Cell Biol.
169, 219–225 (2005).
36. Bieker, K. L., Phillips, G. J. & Silhavy, T. J. The
sec and
prl genes of Escherichia coli. J. Bioenerg. Biomembr.
22, 291–310 (1990).
37. Derman, A. I., Puziss, J. W., Bassford, P. J. Jr &
Beckwith, J. A signal sequence is not required for
protein export in prlA mutants of Escherichia coli.
EMBO J. 12, 879–888 (1993).
38. Plath, K., Mothes, W., Wilkinson, B. M., Stirling, C. J.
& Rapoport, T. A. Signal sequence recognition in
posttranslational protein transport across the yeast
ER membrane. Cell 94, 795–807 (1998).
39. Tam, P. C., Maillard, A. P., Chan, K. K. & Duong, F.
Investigating the SecY plug movement at the SecYEG
translocation channel. EMBO J. 24, 3380–3388
(2005).
40. Maillard, A. P., Lalani, S., Silva, F., Belin, D. & Duong, F.
Deregulation of the SecYEG translocation channel
upon removal of the plug domain. J. Biol. Chem. 282,
1281–1287 (2007).
41.
Junne, T., Schwede, T., Goder, V. & Spiess, M. The plug
domain of yeast Sec61p is important for efficient
protein translocation, but is not essential for cell
viability. Mol. Biol. Cell 17, 4063–4068 (2006).
42.
Tani, K., Tokuda, H. & Mizushima, S. Translocation of
ProOmpA possessing an intramolecular disulfide
bridge into membrane vesicles of Escherichia coli.
Effect of membrane energization. J. Biol. Chem. 265,
17341–17347 (1990).
43. De Keyzer, J., Van Der Does, C. & Driessen, A. J.
Kinetic analysis of the translocation of fluorescent
precursor proteins into Escherichia coli membrane
vesicles. J. Biol. Chem. 277, 46059–46065 (2002).
44.
Gumbart, J. & Schulten, K. Molecular dynamics
studies of the archaeal translocon. Biophys. J. 90,
2356–2367 (2006).
45.
Nishiyama, K., Suzuki, T. & Tokuda, H. Inversion of the
membrane topology of SecG coupled with SecA-
dependent preprotein translocation. Cell 85, 71–81
(1996).
46. Alami, M., Dalal, K., Lelj-Garolla, B., Sligar, S. G. &
Duong, F. Nanodiscs unravel the interaction between
the SecYEG channel and its cytosolic partner SecA.
EMBO J. 26, 1995–2004 (2007).
47.
Mori, H. et al. Fluorescence resonance energy transfer
analysis of protein translocase. SecYE from Thermus
thermophilus HB8 forms a constitutive oligomer in
membranes. J. Biol. Chem. 278, 14257–14264
(2003).
48. Hamman, B. D., Chen, J. C., Johnson, E. E. & Johnson,
A. E. The aqueous pore through the translocon has a
diameter of 40–60 Å during cotranslational protein
translocation at the ER membrane. Cell 89, 535–544
(1997).
49. Osborne, A. R. & Rapoport, T. A. Protein translocation
is mediated by oligomers of the SecY complex with
one SecY copy forming the channel. Cell 129, 97–110
(2007).
Describes how in dimeric SecYEG one trimer
provides the active exit pore and SecA docks on the
other trimer.
50.
Baud, C. et al. Allosteric communication between
signal peptides and the SecA protein DEAD motor
ATPase domain. J. Biol. Chem. 277, 13724–13731
(2002).
51.
Karamanou, S. et al. A molecular switch in SecA
protein couples ATP hydrolysis to protein translocation.
Mol. Microbiol. 34, 1133–1145 (1999).
52.
Papanikou, E. et al. Identification of the preprotein
binding domain of SecA. J. Biol. Chem.
280,
43209–43217 (2005).
53.
Sianidis, G. et al. Cross-talk between catalytic and
regulatory elements in a DEAD motor domain is
essential for SecA function. EMBO J. 20, 961–970
(2001).
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
849
54. Price, A., Economou, A., Duong, F. & Wickner, W.
Separable ATPase and membrane insertion domains
of the SecA subunit of preprotein translocase. J. Biol.
Chem. 271, 31580–31584 (1996).
55.
Keramisanou, D. et al. Disorder-order folding
transitions underlie catalysis in the helicase motor
of SecA. Nature Struct. Mol. Biol. 13, 594–602
(2006).
56. Cordin, O., Banroques, J., Tanner, N. K. & Linder, P.
The DEAD-box protein family of RNA helicases. Gene
367, 17–37 (2006).
57.
Mori, H. & Ito, K. The long alpha-helix of SecA is
important for the ATPase coupling of translocation.
J. Biol. Chem.
281, 36249–36256 (2006).
58.
Breukink, E. et al. The C terminus of SecA is involved
in both lipid binding and SecB binding. J. Biol. Chem.
270, 7902–7907 (1995).
59. Matousek, W. M. & Alexandrescu, A. T. NMR structure
of the C-terminal domain of SecA in the free state.
Biochim. Biophys. Acta 1702, 163–171 (2004).
60. Dempsey, B. R. et al.
Solution NMR structure and
X-ray absorption analysis of the C-terminal zinc-
binding domain of the SecA ATPase. Biochemistry 43,
9361–9371 (2004).
61. Fekkes, P., de Wit, J. G., Boorsma, A., Friesen, R. H.
& Driessen, A. J. Zinc stabilizes the SecB binding site
of SecA. Biochemistry 38, 5111–5116 (1999).
62. Musial-Siwek, M., Rusch, S. L. & Kendall, D. A.
Selective photoaffinity labeling identifies the signal
peptide binding domain on SecA. J. Mol. Biol.
365,
637–648 (2007).
63.
Gelis, I. et al. Structural basis for signal sequence
recognition by the 204kDa translocase motor SecA
determined by NMR. Cell (in the press).
This study presents the first structure of a pre-
protein segment with a translocase component and
reveals movements in the PBD that could underlie
the mechanism of translocation. This is the first
solution structure of SecA. Owing to its size, this
was a major challenge for NMR-based methods.
64.
Karamanou, S. et al. Preprotein-controlled catalysis in
the helicase motor of SecA. EMBO J. 26, 2904–2914
(2007).
First insight into how pre-proteins control ATP-
driven catalysis by the SecA motor through the
PBD.
65.
Vrontou, E., Karamanou, S., Baud, C., Sianidis, G. &
Economou, A. Global co-ordination of protein
translocation by the SecA IRA1 switch. J. Biol. Chem.
279, 22490–22497 (2004).
66. Woodbury, R. L., Hardy, S. J. & Randall, L. L. Complex
behavior in solution of homodimeric SecA. Protein Sci.
11, 875–882 (2002).
67. Driessen, A. J. SecA, the peripheral subunit of the
Escherichia coli precursor protein translocase, is
functional as a dimer. Biochemistry 32, 13190–13197
(1993).
68.
Shilton, B. et al. Escherichia coli SecA shape and
dimensions. FEBS Lett. 436, 277–282 (1998).
69.
Shinkai, A., Akita, M., Matsuyama, S. & Mizushima, S.
Quantitative renaturation from a guanidine-denatured
state of the SecA dimer, a 200 KDa protein involved in
protein secretion in Escherichia coli. Biochem.
Biophys. Res. Commun. 172, 1217–1223 (1990).
70.
Akita, M., Shinkai, A., Matsuyama, S. & Mizushima, S.
SecA, an essential component of the secretory
machinery of Escherichia coli, exists as homodimer.
Biochem. Biophys. Res. Commun. 174, 211–216
(1991).
71. Dempsey, B. R., Economou, A., Dunn, S. D. & Shilton,
B. H. The ATPase domain of SecA can form a tetramer
in solution. J. Mol. Biol. 315, 831–843 (2002).
72.
Hirano, M., Matsuyama, S. & Tokuda, H. The carboxyl-
terminal region is essential for Sec-A dimerization.
Biochem. Biophys. Res. Commun. 229, 90–95
(1996).
73.
Karamanou, S. et al. Escherichia coli SecA truncated
at its termini is functional and dimeric. FEBS Lett.
579, 1267–1271 (2005).
74.
Mori, H. et al. Amino-terminal region of SecA is
involved in the function of SecG for protein
translocation into Escherichia coli membrane vesicles.
J. Biochem. (Tokyo)
124, 122–129 (1998).
75.
Or, E., Boyd, D., Gon, S., Beckwith, J. & Rapoport, T.
The bacterial ATPase SecA functions as a monomer in
protein translocation. J. Biol. Chem. 280, 9097–9105
(2005).
76.
Mori, H. & Ito, K. Different modes of SecY–SecA
interactions revealed by site-directed in vivo
photo-cross-linking.
Proc. Natl Acad. Sci. USA 103,
16159–16164 (2006).
77. van der Sluis, E. O. et al.
Identification of two
interaction sites in SecY that are important for the
functional interaction with SecA. J. Mol. Biol.
361,
839–849 (2006).
78.
Nagamori, S., Nishiyama, K. & Tokuda, H. Membrane
topology inversion of SecG detected by labeling with a
membrane-impermeable sulfhydryl reagent that
causes a close association of SecG with SecA.
J. Biochem. (Tokyo)
132, 629–634 (2002).
79.
Nishiyama, K., Mizushima, S. & Tokuda, H.
Preferential interaction of Sec-G with Sec-E stabilizes
an unstable Sec-E derivative in the
Escherichia coli
cytoplasmic membrane. Biochem. Biophys. Res.
Commun. 217, 217–223 (1995).
80.
Snyders, S., Ramamurthy, V. & Oliver, D. Identification
of a region of interaction between Escherichia coli
SecA and SecY proteins. J. Biol. Chem. 272,
11302–11306 (1997).
81.
Dapic, V. & Oliver, D. Distinct membrane binding
properties of N- and C-terminal domains of
Escherichia coli SecA ATPase. J. Biol. Chem. 275,
25000–25007 (2000).
82.
de Keyzer, J. et al. Covalently dimerized SecA is
functional in protein translocation. J. Biol. Chem. 280,
35255–35260 (2005).
83. Jilaveanu, L. B. & Oliver, D. SecA dimer cross-linked at
its subunit interface is functional for protein
translocation. J. Bacteriol. 188, 335–338 (2006).
84.
Or, E., Navon, A. & Rapoport, T. Dissociation of the
dimeric SecA ATPase during protein translocation
across the bacterial membrane. EMBO J. 21,
4470–4479 (2002).
85.
Duong, F. Binding, activation and dissociation of the
dimeric SecA ATPase at the dimeric SecYEG
translocase. EMBO J. 22, 4375–4384 (2003).
86.
Benach, J. et al. Phospholipid-induced
monomerization and signal-peptide-induced
oligomerization of SecA. J. Biol. Chem.
278,
3628–3638 (2003).
87.
Lill, R., Dowhan, W. & Wickner, W. The ATPase activity
of SecA is regulated by acidic phospholipids, SecY, and
the leader and mature domains of precursor proteins.
Cell 60, 271–280 (1990).
88.
Tziatzios, C. et al. The bacterial protein-translocation
complex: SecYEG dimers associate with one or two
SecA molecules. J. Mol. Biol. 340, 513–524 (2004).
89.
Collinson, I. et al. Projection structure and oligomeric
properties of a bacterial core protein translocase.
EMBO J. 20, 2462–2471 (2001).
90.
Duong, F. & Wickner, W. Distinct catalytic roles of the
SecYE, SecG and SecDFyajC subunits of preprotein
translocase holoenzyme. EMBO J. 16, 2756–2768
(1997).
91. Pogliano, J. A. & Beckwith, J. SecD and SecF facilitate
protein export in Escherichia coli. EMBO J. 13,
554–561 (1994).
92.
Matsuyama, S., Fujita, Y. & Mizushima, S. SecD is
involved in the release of translocated secretory
proteins from the cytoplasmic membrane of
Escherichia coli. EMBO J. 12, 265–270 (1993).
93. Economou, A., Pogliano, J. A., Beckwith, J., Oliver,
D. B. & Wickner, W. SecA membrane cycling at SecYEG
is driven by distinct ATP binding and hydrolysis events
and is regulated by SecD and SecF. Cell 83,
1171–1181 (1995).
94. Yi, L. & Dalbey, R. E. Oxa1/Alb3/YidC system for
insertion of membrane proteins in mitochondria,
chloroplasts and bacteria. Mol. Membr. Biol. 22,
101–111 (2005).
95.
Serek, J. et al. Escherichia coli YidC is a membrane
insertase for Sec-independent proteins. EMBO J. 23,
294–301 (2004).
96.
Beck, K. et al. YidC, an assembly site for polytopic
Escherichia coli membrane proteins located in
immediate proximity to the SecYE translocon and
lipids. EMBO Rep. 2, 709–714 (2001).
97. Nouwen, N. & Driessen, A. J. SecDFyajC forms a
heterotetrameric complex with YidC. Mol. Microbiol.
44, 1397–1405 (2002).
98.
van Dalen, A. & de Kruijff, B. The role of lipids in
membrane insertion and translocation of bacterial
proteins. Biochim. Biophys. Acta 1694, 97–109
(2004).
99.
Suzuki, H., Nishiyama, K. & Tokuda, H. Increases in
acidic phospholipid contents specifically restore protein
translocation in a cold-sensitive secA or secG null
mutant. J. Biol. Chem. 274, 31020–31024 (1999).
100. Rietveld, A. G., Killian, J. A., Dowhan, W. &
de Kruijff, B. Polymorphic regulation of membrane
phospholipid composition in Escherichia coli. J. Biol.
Chem. 268, 12427–12433 (1993).
101.
Xu, Z. et al. Crystal structure of the bacterial protein
export chaperone SecB. Nature Struct. Biol. 12,
1172–1177 (2000).
102.
Zhou, J. & Xu, Z. Structural determinants of SecB
recognition by SecA in bacterial protein
translocation. Nature Struct. Biol. 10, 942–947
(2003).
103.
Dekker, C., de Kruijff, B. & Gros, P. Crystal structure
of SecB from Escherichia coli. J. Struct. Biol. 144,
313–319 (2003).
104. Randall, L. L. et al.
The interaction between the
chaperone SecB and its ligands: evidence for multiple
subsites for binding. Protein Sci. 7, 2384–2390
(1998).
105. Knoblauch, N. T. et al.
Substrate specificity of the
SecB chaperone. J. Biol. Chem. 274, 34219–34225
(1999).
106. Woodbury, R. L. et al.
Complexes between protein
export chaperone SecB and SecA. Evidence for
separate sites on SecA providing binding energy
and regulatory interactions. J. Biol. Chem. 275,
24191–24198 (2000).
107. Fekkes, P., van der Does, C. & Driessen, A. J. The
molecular chaperone SecB is released from the
carboxy-terminus of SecA during initiation of
precursor protein translocation. EMBO J. 16,
6105–6113 (1997).
108. Crane, J. M. et al.
Mapping of the docking of SecA
onto the chaperone SecB by site-directed spin
labeling: insight into the mechanism of ligand transfer
during protein export. J. Mol. Biol. 353, 295–307
(2005).
109. Karamyshev, A. L. & Johnson, A. E. Selective SecA
association with signal sequences in ribosome-bound
nascent chains: a potential role for SecA in ribosome
targeting to the bacterial membrane. J. Biol. Chem.
280, 37930–37940 (2005).
110. Eisner, G., Koch, H. G., Beck, K., Brunner, J. & Muller, M.
Ligand crowding at a nascent signal sequence. J. Cell
Biol. 163, 35–44 (2003).
111. Chou, Y. T. & Gierasch, L. M. The conformation of a
signal peptide bound by Escherichia coli preprotein
translocase SecA. J. Biol. Chem.
280, 32753–32760
(2005).
112.
Khatib, K. & Belin, D. A novel class of secA alleles that
exert a signal-sequence-dependent effect on protein
export in Escherichia coli. Genetics 162, 1031–1043
(2002).
113. Kourtz, L. & Oliver, D. Tyr-326 plays a critical role in
controlling SecA-preprotein interaction. Mol.
Microbiol. 37, 1342–1356 (2000).
114.
Baud, C. et al. Purification of a functional mature
region from a SecA-dependent preprotein. Protein
Expr. Purif. 40, 336–339 (2005).
115. Schiebel, E., Driessen, A. J., Hartl, F. U. & Wickner, W.
Δμ
H
+
and ATP function at different steps of the
catalytic cycle of preprotein translocase. Cell 64,
927–939 (1991).
Biochemical dissection of the distinct energetic
roles of ATP and the PMF in promoting
translocation.
116. Fak, J. J. et al.
Nucleotide exchange from the
high-affinity ATP-binding site in SecA is the rate-
limiting step in the ATPase cycle of the soluble
enzyme and occurs through a specialized
conformational state. Biochemistry 43,
7307–7327 (2004).
117.
Economou, A. & Wickner, W. SecA promotes
preprotein translocation by undergoing ATP-driven
cycles of membrane insertion and deinsertion. Cell 78,
835–843 (1994).
Demonstrated that ATP energy is used for
defined SecA conformational events that
correlate with the defined movements of short
pre-protein segments described by Schiebel and
colleagues in reference 115.
118.
Yamada, H., Matsuyama, S., Tokuda, H. &
Mizushima, S. A high concentration of SecA allows
proton motive force-independent translocation
of a model secretory protein into Escherichia coli
membrane vesicles. J. Biol. Chem. 264,
18577–18581 (1989).
119.
Nishiyama, K., Fukuda, A., Morita, K. & Tokuda, H.
Membrane deinsertion of SecA underlying proton
motive force-dependent stimulation of protein
translocation. EMBO J. 18, 1049–1058 (1999).
120.
Bessonneau, P., Besson, V., Collinson, I. & Duong, F.
The SecYEG preprotein translocation channel is a
conformationally dynamic and dimeric structure.
EMBO J. 21, 995–1003 (2002).
RE V IE W S
850
|
NOVEMBER 2007
|
VOLUME 5 www.nature.com/reviews/micro
121. Nouwen, N., de Kruijff, B. & Tommassen, J.
prlA suppressors in Escherichia coli relieve the proton
electrochemical gradient dependency of translocation
of wild-type precursors. Proc. Natl Acad. Sci. USA 93,
5953–5957 (1996).
122. Arkowitz, R. A., Joly, J. C. & Wickner, W. Translocation
can drive the unfolding of a preprotein domain. EMBO J.
12, 243–253 (1993).
123. Engelman, D. M. & Steitz, T. A. The spontaneous
insertion of proteins into and across membranes: the
helical hairpin hypothesis. Cell 23, 411–422 (1981).
124. Joly, J. C. & Wickner, W. The SecA and SecY subunits
of translocase are the nearest neighbors of a
translocating preprotein, shielding it from
phospholipids. EMBO J. 12, 255–263 (1993).
First direct demonstration that pre-proteins travel in a
proteinaceous environment that is composed of SecA
and SecY and form distinct interactions with each.
125. Economou, T. & Dalbey, R. E. (eds) Special issue of
Biochimica et Biophysica Acta — Molecular Cell
Research on Protein Export/Secretion in Bacteria
Vol. 1694, 1–333 (2004).
126. White, S. H. & von Heijne, G. Transmembrane helices
before, during, and after insertion. Curr. Opin. Struct.
Biol. 15, 378–386 (2005).
127.
Kiefer, D. & Kuhn, A. Hydrophobic forces drive
spontaneous membrane insertion of the
bacteriophage Pf3 coat protein without topological
control. EMBO J. 18, 6299–6306 (1999).
128.
Gallusser, A. & Kuhn, A. Initial steps in protein
membrane insertion. Bacteriophage M13 procoat
protein binds to the membrane surface by
electrostatic interaction. EMBO J. 9, 2723–2729
(1990).
129. Urbanus, M. L. et al.
Sec-dependent membrane
protein insertion: sequential interaction of nascent
FtsQ with SecY and YidC. EMBO Rep. 2, 524–529
(2001).
130.
van der Laan, M., Bechtluft, P., Kol, S., Nouwen, N.
& Driessen, A. J. F
1
F
0
ATP synthase subunit c is a
substrate of the novel YidC pathway for membrane
protein biogenesis. J. Cell Biol. 165, 213–222
(2004).
131.
Clantin, B. et al. Structure of the membrane protein
FhaC: a member of the Omp85-TpsB transporter
superfamily. Science 317, 957–961 (2007).
132.
Kim, S. et al. Structure and function of an essential
component of the outer membrane protein
assembly machine. Science 317, 961–964 (2007).
133. Bos, M. P., Robert, V. & Tommassen, J. Biogenesis
of the Gram-negative bacterial outer membrane.
Annu. Rev. Microbiol. 7 June 2007 (doi:10.1146/
annurev.micro.61.080706.093245).
134. Crane, J. M. et al.
Sites of interaction of a precursor
polypeptide on the export chaperone SecB mapped
by site-directed spin labeling. J. Mol. Biol. 363,
63–74 (2006).
Acknowledgements
We are grateful to the Kalodimos laboratory for an exciting
collaboration and to B. Kalodimos and T. Pugsley for stimu-
lating discussions. Research in our laboratory is supported
by grants from the European Union (LSHG-CT-2005-037586),
the Greek General Secretariat of Research and the European
Regio nal Deve lopm ent Fun d (01 AKMO N46 and
PENED03ED623).
DATABASES
Entrez Genome Project: http://www.ncbi.nlm.nih.gov/
entrez/query.fcgi?db=genomeprj
Bacillus subtilis | Escherichia coli | Methanococcus jannaschii |
Mycobacterium tuberculosis
| Thermus thermophilus
Entrez Protein: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=protein
ecSecYEG
Protein Data Bank: http://www.rcsb.org/pdb/home/home.do
bsSecA | bsSecA I | bsSecA II | ecSecA | ecSecB | mjSecYEG |
mtSecA | ttSecA
FURTHER INFORMATION
Anastassios Economou’s homepage:
http://ecoserver.imbb.forth.gr
ALL LINKS ARE ACTIVE IN THE ONLINE PDF
RE V IE W S
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 5
|
NOVEMBER 2007
|
851
... They also protect their parent bacteria by acting as decoys for antibiotics and bacteriophages [6]. Recently, MVs were found to be involved in bacteria-host interactions, such as host immunomodulation [7] and the delivery of toxins and pathogenic components to the host [8]. ...
... Outer membrane vesicles (OMVs) derived from an outer membrane of gram-negative bacteria have long been studied [3,8]. OMVs including pathogenic components as well as membrane proteins and lipopolysaccharide (LPS) have been applied as vaccines owing to their unique components and biological functions [9,10]. ...
Article
Full-text available
Nanosized membrane vesicles (MVs) released by bacteria play important roles in both bacteria–bacteria and bacteria–host interactions. Some gram-positive lactic acid bacteria produce MVs exhibiting immunoregulatory activity in the host. We found that both bacterial cells and MVs of Limosilactobacillus antri JCM 15950, isolated from the human stomach mucosa, enhance immunoglobulin A production by murine Peyer’s patch cells. However, the thick cell walls of gram-positive bacteria resulted in low MV production, limiting experiments and applications using MVs. In this study, we evaluated the effects of glycine, which inhibits cell wall synthesis, on the immunostimulatory MV productivity of L. antri. Glycine inhibited bacterial growth while increasing MV production, with 20 g/L glycine increasing MV production approximately 12-fold. Glycine was most effective at increasing MV production when added in the early exponential phase, which indicated that cell division in the presence of glycine increased MV production. Finally, glycine increased MV productivity approximately 16-fold. Furthermore, glycine-induced MVs promoted interleukin-6 production by macrophage-like J774.1 cells, and the immunostimulatory activity was comparable to that of spontaneously produced MVs. Our results indicate that glycine is an effective agent for improving the production of MVs with immunostimulatory activity in gram-positive lactic acid bacteria, which can be applied as mucosal adjuvants and functional foods.
... It is a highly conserved pathway present in all classes of bacteria [15]. Sec-routed proteins have a signal peptide at their N-terminus that directs them towards the SecYEG membrane protein channel, after which they are driven stepwise across the membrane by the ATPase molecular motor SecA [16]. The transported protein then folds on the trans side of the membrane. ...
... The transported protein then folds on the trans side of the membrane. In many Gram-negative bacteria, SecB stabilises and targets the unfolded protein to SecA, while in Gram-positive and other Gram-negative bacteria, general chaperones maintain the protein in an unfolded state [16]. Another common secretion pathway is the Twin Arginine Translocation (Tat) pathway, in which proteins are exported in a folded state [17], however, not all bacterial species have a Tat pathway [18]. ...
Article
Full-text available
Staphylococcus aureus is a major human pathogen that utilises many surface-associated and secreted proteins to form biofilms and cause disease. However, our understanding of these processes is limited by challenges of using fluores-cent protein reporters in their native environment, because they must be ex-ported and fold correctly to become fluorescent. Here, we demonstrate the feasibility of using the monomeric superfolder GFP (msfGFP) exported from S. aureus. By fusing msfGFP to signal peptides for the Secretory (Sec) and Twin Arginine Translocation (Tat) pathways, the two major secretion pathways in S. aureus, we quantified msfGFP fluorescence in bacterial cultures and cell-free supernatant from the cultures. When fused to a Tat signal peptide, we detect-ed msfGFP fluorescence inside but not outside bacterial cells, indicating a fail-ure to export msfGFP. However, when fused to a Sec signal peptide, msfGFP fluorescence was present outside cells, indicating successful export of the msfGFP in the unfolded state, followed by extracellular folding and maturation to the photoactive state. We applied this strategy to study coagulase (Coa), a secreted protein and a major contributor to the formation of a fibrin network in S. aureus biofilms that protects bacteria from the host immune system and increases attachment to host surfaces. We confirmed that a genomically inte-grated C-terminal fusion of Coa to msfGFP does not impair the activity of Coa or its localisation within the biofilm matrix. Our findings demonstrate that msfGFP is a good candidate fluorescent reporter to consider when studying proteins secreted by the Sec pathway in S. aureus.
... Sec-dependent pathway including three components (a motor protein, a membrane integrated conducting channel and the SecYEG translocase) is considered one of the most highly conserved protein secretion pathways ubiquitous in all domains of bacteria life [50,54]. The effectors secreted by phytoplasmas can directly transit into host cells via Sec-dependent pathway [55]. ...
... According to the bacteria secreted system [50], secretory pathway with highly efficient protein secretion mechanism in Ca. P. ziziphi was presented (Fig. 5A) [51][52][53], in which secreted proteins were first recognized by ffh and then bound to its receptor Ftsy. ...
Article
Full-text available
Phytoplasmas are obligate cell wall-less prokaryotic bacteria that primarily multiply in plant phloem tissue. Jujube witches’ broom (JWB) associated with phytoplasma is a destructive disease of jujube (Ziziphus jujuba Mill.). Here we report the complete ‘Candidatus Phytoplasma ziziphi’ chromosome of strain Hebei-2018, which is a circular genome of 764,108-base pairs with 735 predicted CDS. Notably, extra 19,825 bp (from 621,995 to 641,819 bp) compared to the previously reported one complements the genes involved in glycolysis, such as pdhA, pdhB, pdhC, pdhD, ackA, pduL and LDH. The synonymous codon usage bias (CUB) patterns by using comparative genomics analysis among the 9 phytoplasmas were similar for most codons. The ENc-GC3s analysis among the 9 phytoplasmas showed a greater effect under the selection on the CUBs of phytoplasmas genes than mutation and other factors. The genome exhibited a strongly reduced ability in metabolic synthesis, while the genes encoding transporter systems were well developed. The genes involved in sec-dependent protein translocation system were also identified.The expressions of nine FtsHs encoding membrane associated ATP-dependent Zn proteases and Mn-SodA with redox capacity in the Ca. P. ziziphi was positively correlated with the phytoplasma concentration. Taken together, the genome will not only expand the number of phytoplasma species and provide some new information about Ca. P. ziziphi, but also contribute to exploring its pathogenic mechanism.
... However, gram positive bacteria vary a bit, and they follow Sec/Tat pathway for translocation of proteins across single membrane. This mechanism varies slightly in pathogens (Papanikou et al., 2007;Muller M, 2005;Albers et al., 2006). Gram negative bacteria such as E. coli is of great interest to biotechnology-based industries. ...
Article
Full-text available
Protein secretion is an important process in any living organism, and this is mediated by numerous steps and several cellular components. This review provides an overview of protein production from nascent polypeptide synthesis stage to secretion of these proteins into extracellular space by vesicular sorting. This review also provides insights on the factors involved in ER and Golgi in this pathway and their role. It also covers responsibility of translocons and chaperones in stress response pathways such as UPR and ERAD. Beyond all these a series of quality control checks are performed by cellular machinery to ensure the quality of protein delivered. So essentially this review covers all the pathways involved from initiation of protein expression to secretion.
... This was the case for several genes related to the Sec system, which is responsible for the insertion, translocation, and secretion across the membrane of unfolded polypeptides, which carry a removable N-terminal signal sequence. Thus, genes coding for the membrane-embedded SecYEG translocon, the SecA ATPase and piloting factors such as the signal recognition particle (SRP) Ffh or the SecB chaperone, which maintain newly synthesized proteins in an unfolded conformation and drive them to the membrane (Papanikou et al., 2007), exhibited increased expression during predation. In contrast to the Sec translocon, the Tat system is responsible for exporting previously folded proteins which have a particular signal peptide containing a recognizable twin-arginine motif (Pickering and Oresnik, 2010;Palmer and Berks, 2012). ...
Article
Full-text available
Bacterial predation impacts microbial community structures, which can have both positive and negative effects on plant and animal health and on environmental sustainability. Myxococcus xanthus is an epibiotic soil predator with a broad range of prey, including Sinorhizobium meliloti, which establishes nitrogen-fixing symbiosis with legumes. During the M. xanthus-S. meliloti interaction, the predator must adapt its transcriptome to kill and lyse the target (predatosome), and the prey must orchestrate a transcriptional response (defensome) to protect itself against the biotic stress caused by the predatory attack. Here, we describe the transcriptional changes taking place in S. meliloti in response to myxobacterial predation. The results indicate that the predator induces massive changes in the prey transcriptome with up-regulation of protein synthesis and secretion, energy generation, and fatty acid (FA) synthesis, while down-regulating genes required for FA degradation and carbohydrate transport and metabolism. The reconstruction of up-regulated pathways suggests that S. meliloti modifies the cell envelop by increasing the production of different surface polysaccharides (SPSs) and membrane lipids. Besides the barrier role of SPSs, additional mechanisms involving the activity of efflux pumps and the peptide uptake transporter BacA, together with the production of H2O2 and formaldehyde have been unveiled. Also, the induction of the iron-uptake machinery in both predator and prey reflects a strong competition for this metal. With this research we complete the characterization of the complex transcriptional changes that occur during the M. xanthus-S. meliloti interaction, which can impact the establishment of beneficial symbiosis with legumes.
... Proteins secreted by the Sec pathway depend on a signal located in the aminoterminal region of the proteins. This signal comprises twenty amino acids divided into a positively charged N-terminal region, a hydrophobic center, and a polar C-terminal region (Papanikou et al., 2007). As a signal peptide, the proteins secreted by the Tat pathway have a sequence of two arginines in their N-terminal region (Robinson & Bolhuis, 2004). ...
Preprint
Full-text available
Understanding protein secretion pathways are of paramount importance in studying diseases caused by bacteria and their respective treatments. Most such paths must signal ways to identify secretion. However, some proteins, known as non-classical secreted proteins, do not have signaling ways. This study aims to classify such proteins from predictive machine-learning techniques. Guided by the literature, we collected a set of physical-chemical characteristics of amino acids from the AA index site bolding know protein motifs, like hydrophobicity. In this work, we developed a six steps method (Alignment, Preliminary classification, mean outliers, two Clustering algorithms, and Random choice) to filter data from raw genomes and compose a negative dataset in contrast to a positive dataset of 141 proteins also gathered from the literature. Using a conventional Random Forest machine-learning algorithm, we obtained an accuracy of ~91% on classifying non-classical secreted proteins in a validation dataset with 14 positive and 92 negatives proteins - sensitivity and specificity of 91 and ~86%, respectively, performance compared to state of the art for non-classical secretion classification, but a less sophisticated algorithm allows us to classify bacterial proteins concerning secretion by non-classical pathways more rapidly. Therefore, this research has shown that selecting an appropriate descriptors' set and an expressive training dataset compensates for not using an advanced machine learning algorithm for the secretion by non-classical pathways purpose. The data and software from this work, available at https://github.com/santosardr/non-CSPs, can be downloaded for standalone use without needing third-party software.
... SignalP predicts only conserved signal sequences of proteins transported predominantly by the Sec or Tat pathways [41]. However, non-canonical secretion by the natural host cannot be excluded since some proteins are apparently secreted with so far unknown export mechanisms which sometimes are species-specific or limited to only a few proteins [42,43]. For example, a study focused on the secretome of B. subtilis showed that the genome-based annotations mirrored the real composition of the extracellular proteome to only 50% [44]. ...
Article
Full-text available
Background Plant cell walls represent the most plentiful renewable organic resource on earth, but due to their heterogeneity, complex structure and partial recalcitrance, their use as biotechnological feedstock is still limited. Results In order to identify efficient enzymes for polysaccharide breakdown, we have carried out functional screening of metagenomic fosmid libraries from biogas fermenter microbial communities grown on sugar beet pulp, an arabinan-rich agricultural residue, or other sources containing microbes that efficiently depolymerize polysaccharides, using CPH (chromogenic polysaccharide hydrogel) or ICB (insoluble chromogenic biomass) labeled polysaccharide substrates. Seventy-one depolymerase-encoding genes were identified from 55 active fosmid clones by using Illumina and Sanger sequencing and dbCAN CAZyme (carbohydrate-active enzyme) annotation. An around 56 kb assembled DNA fragment putatively originating from Xylanivirga thermophila strain or a close relative was analyzed in detail. It contained 48 ORFs (open reading frames), of which 31 were assigned to sugar metabolism. Interestingly, a large number of genes for enzymes putatively involved in degradation and utilization of arabinose-containing carbohydrates were found. Seven putative arabinosyl hydrolases from this DNA fragment belonging to glycoside hydrolase (GH) families GH51 and GH43 were biochemically characterized, revealing two with endo-arabinanase activity and four with exo-α- l -arabinofuranosidase activity but with complementary cleavage properties. These enzymes were found to act synergistically and can completely hydrolyze SBA (sugar beet arabinan) and DA (debranched arabinan). Conclusions We screened 32,776 fosmid clones from several metagenomic libraries with chromogenic lignocellulosic substrates for functional enzymes to advance the understanding about the saccharification of recalcitrant lignocellulose. Seven putative X. thermophila arabinosyl hydrolases were characterized for pectic substrate degradation . The arabinosyl hydrolases displayed maximum activity and significant long-term stability around 50 °C. The enzyme cocktails composed in this study fully degraded the arabinan substrates and thus could serve for arabinose production in food and biofuel industries.
... Alternatively, they can be transported to the periplasm before crossing the outer membrane in a two-step transport (43). In this case, the transport across the inner membrane is facilitated by the general export pathway (Sec or Tat pathways) (126,127). ...
Article
Bacteria can adhere to almost any surface and form resistant biofilm communities. These pose a major clinical burden as they can form on medical devices such as catheters and are hard to prevent. Catheter- associated urinary tract infections (CAUTIs) are the most common nosocomial infection as they represent ~40% of all nosocomial infections. Hook et al., previously developed a series of polymers that prevent biofilm formation. Following a screen of 20,000 polymers and copolymers and subsequent scale-up, the hit acrylate polymer, EGdPEA has been clinically approved for the coating of urinary tract catheters and presents superior antibiofilm properties compared with silver-embedded silicone catheters. The bacterial biofilm resistance mechanism and surface sensing pathways involved are under investigation. In P. aeruginosa surface sensing includes the flagella, type IV pili and the wsp system. The present study sought to investigate SadB, an essential gene for biofilm formation and CheA, a histidine kinase of the chemotaxis system for their involvement in surface sensing. SadB is involved in the switch from reversible to irreversible surface attachment. Expression of sadB under a constitutive promoter overcomes the resistance of EGdPEA to biofilm formation. CheA is responsible for transmitting environmental signals to the cell and activating Pch phosphodiesterase, which is important in surface sensing. Biofilm and motility assays and monitoring on-surface gene expression were used as a test of potential involvement in surface sensing. In agreement with the literature, SadB was found to be essential for biofilm formation and to affect swarming motility. sadB was also found to be differentially expressed on the anti and pro-biofilm surfaces. In contrast, CheA affected swarming but did not affect biofilm formation on most polymer surfaces and glass. The promoter of cheA also controls the cheI operon with 13 more genes. The expression of the operon was found to be different on different surfaces but CheA is unlikely to be important in surface sensing. The literature and findings of the present study highlighted the importance of sadB in surface sensing, but the regulatory pathway was still unknown. Whole transcriptome analysis (RNAseq) was employed to investigate the downstream regulation of sadB in an untargeted manner. RT-PCR and phenotypic assays were used to validate the data. SadB was found to be a global regulator controlling up to 50% of the genome. sadB was shown to negatively affect the pqs system and subsequently the rhl system, rhamnolipid and pyocyanin production and the denitrification pathway. The levels of the PQS signal, pyocyanin and rhamnolipids were also shown to be affected by sadB. The denitrification pathway downregulation in the absence of sadB was shown to affect growth under microaerophilic conditions. Furthermore, sadB positively affected biofilm related targets such as Psl and c-di-GMP production. Advances were made in the present study in discovering the downstream targets of sadB that explain some of the observed phenotypes, however the function of SadB is still unknown. A study from Muriel et al., reported that a sadB homolog in P. fluorescens binds c-di- GMP. The sadB homolog of P. aeruginosa was successfully purified and tested using SPR and thermal shift assay. The results showed no binding of c-di-GMP. Additionally, attempts were made to resolve the crystal structure of SadB. SadB crystals were obtained after screening commercial crystal screens and optimisation of the conditions but the resolution of the diffraction was not sufficient. Finally, SadB was found to localise in the cytoplasm using an anti-SadB antibody and Western blot and a SadB fluorescent fusion and super-resolution microscopy.
Preprint
Protein- O -mannosylation (POM) is a form of O -glycosylation that is ubiquitous throughout all domains of life and has been extensively characterized in eukaryotic systems. However, in prokaryotes this process has only been investigated in terms of pathogenicity (in Mycobacterium tuberculosis ) even though there are many non-pathogenic bacteria that are known to regularly carry out POM. To date, there is no consensus on what benefit POM imparts to the non-pathogenic bacteria that can perform it. Though the generation of a POM deficient mutant of Corynebacterium glutamicum – a widely utilized and known mannosylating actinobacteria – this work shows that even closely related actinobacterial GT-39s (the enzymes responsible for the initiation of POM) can potentially have different activities and substrate specificities for targets of POM. Moreover, presented here is evidence that POM does not only occur in a SEC-dependent manner; POM also occurs with TAT and non-SEC secreted substrates in a specific and likely tightly regulated manner. Together these results highlight the need for further biochemical characterization of POM in these and other bacterial species to help elucidate the true nature of its biological functions. Importance Both the mechanism and overall cellular function of protein- O -mannosylation, a ubiquitous subset of O -glycosylation, is poorly understood in bacterial systems. In Mycobacterium tuberculosis and other pathogenic actinobacteria, numerous secreted virulence factors were identified as mannoproteins, with protein- O -mannosylation deficient mutants displaying a less virulent phenotype due to these proteins lacking the modification. However, these findings do not offer any explanations as to why non-pathogenic strains of actinobacteria also perform this modification as in these organisms it is seemingly dispensable. Corynebacterium glutamicum is a widely utilized, industrially relevant actinobacteria that also performs protein- O -mannosylation. This manuscript describes the utilization of C. glutamicum as a Gram-positive recombinant host for the in vivo study of actinobacterial protein- O -mannosylation and demonstrates the distinct lack of first-hand biochemical data of the process in prokaryotes.
Article
Full-text available
Life-like systems need to maintain a basal metabolism, which includes importing a variety of building blocks required for macromolecule synthesis, exporting dead-end products, and recycling cofactors and metabolic intermediates, while maintaining steady internal physical and chemical conditions (physicochemical homeostasis). A compartment, such as a unilamellar vesicle, functionalized with membrane-embedded transport proteins and metabolic enzymes encapsulated in the lumen meets these requirements. Here, we identify four modules designed for a minimal metabolism in a synthetic cell with a lipid bilayer boundary: energy provision and conversion, physicochemical homeostasis, metabolite transport, and membrane expansion. We review design strategies that can be used to fulfill these functions with a focus on the lipid and membrane protein composition of a cell. We compare our bottom-up design with the equivalent essential modules of JCVI-syn3a, a top-down genome-minimized living cell with a size comparable to that of large unilamellar vesicles. Finally, we discuss the bottlenecks related to the insertion of a complex mixture of membrane proteins into lipid bilayers and provide a semiquantitative estimate of the relative surface area and lipid-to-protein mass ratios (i.e., the minimal number of membrane proteins) that are required for the construction of a synthetic cell.
Article
Full-text available
The major route for protein export or membrane integration in bacteria occurs via the Sec-dependent transport apparatus. The core complex in the inner membrane, consisting of SecYEG, forms a protein-conducting channel, while the ATPase SecA drives translocation of substrate across the membrane. The SecYEG complex from Escherichia coli was overexpressed, purified and crystallized in two dimensions. A 9 Å projection structure was calculated using electron cryo-microscopy. The structure exhibits P121 symmetry, having two asymmetric units inverted with respect to one another in the unit cell. The map shows elements of secondary structure that appear to be transmembrane helices. The crystallized form of SecYEG is too small to comprise the translocation channel and does not contain a large pore seen in other studies. In detergent solution, the SecYEG complex displays an equilibrium between monomeric and tetrameric forms. Our results therefore indicate that, unlike other known channels, the SecYEG complex can exist as both an assembled channel and an unassembled smaller unit, suggesting that transitions between the two states occur during a functional cycle.
Article
Disulfide bond formation is a catalyzed process in vivo. In prokaryotes, the oxidation of cysteine pairs is achieved by the transfer of disulfides from the highly oxidizing DsbA/DsbB catalytic machinery to substrate proteins. The oxidizing power utilized by this system comes from the membrane-embedded electron transport system, which utilizes molecular oxygen as a final oxidant. Proofreading of disulfide bond formation is performed by the DsbC/DsbD system, which has the ability to rearrange non-native disulfides to their native configuration. These disulfide isomerization reactions are sustained by a constant supply of reducing power provided by the cytoplasmic thioredoxin system, utilizing NADPH as the ultimate electron source.
Article
Recent studies have shown that there is a pathway that is evolutionarily conserved for the insertion of proteins into the membrane in mitochondria, chloroplasts, and bacteria. In this pathway, the Oxa1/Alb3/YidC proteins are believed to function as membrane insertases that play an important role in the membrane protein biogenesis of respiratory and energy transduction proteins. Additional roles of the Oxa1/Alb3/YidC members may be in the lateral integration of proteins into the lipid bilayer, and in the folding and assembly of proteins into membrane protein complexes.
Article
Size exclusion chromatography of the cytosolic fraction of SecA-overproducing cells of suggested that SecA, an essential component of the secretory machinery, exists as an oligomer. The subunit structure of SecA was then studied using a purified specimen. Estimation of the molecular mass by means of ultracentrifugation and chemical cross-linking analysis revealed that SecA exists as a homodimer. The purified SecA was denatured in 6 M guanidine-HCl and renatured to a dimer, which was fully active in terms of translocation, even in the presence of 1 mM dithiothreitol. It is suggested that the dimeric structure is not critically maintained by disulfide bonding between the two subunits, each of which contains four cysteine residues.
Article
SecA is an essential component of the protein secretory machinery of . SecA denatured in 6 M guanidine hydrochloride was quantitatively renatured through dilution and dialysis. The renatured SecA was the same as native SecA as to the CD spectrum, fluorescence spectrum for tryptophan residues and dimeric structure. It was as functionally active as native SecA as to interactions with ATP and presecretory proteins, and translocation. SecA-N95, which lacks the carboxyl-terminal 70 amino acid residues including three of four cysteine residues and yet is as active as intact SecA as to translocation, was also renatured to an active form from the guanidine solution. Furthermore, the renaturation of SecA took place in the presence of 1 mM dithiothreitol. It is concluded that disulfide bridges, both intra- and intermolecular ones, do not play a role in the folding and functioning of the SecA molecule.
Article
SecA is an essential ATP-dependent motor protein that interacts with the preprotein and translocon to drive protein translocation across the eubacterial plasma membrane. A region containing residues 267–340 has been proposed to comprise the preprotein binding site of Escherichia coli SecA. To elucidate the function of this region further, we isolated mutants using a combination of region-specific polymerase chain reaction (PCR) mutagenesis and a genetic and biochemical screening procedure. Although this region displayed considerable plasticity based on phylogenetic and genetic analysis, Tyr-326 was found to be critical for SecA function. secA mutants with non-conservative substitutions at Tyr-326 showed strong protein secretion defects in vivo and were completely defective for SecA-dependent translocation ATPase activity in vitro. The SecA-Y326 mutant proteins were normal in their membrane, SecYE and nucleotide-binding properties. However, they exhibited a reduced affinity for preprotein and were defective in preprotein release, as assessed by several biochemical assays. Our results indicate that the region containing Tyr-326 functions as a conformational response element to regulate the preprotein binding and release cycle of SecA.
Article
SecA, the motor subunit of bacterial polypeptide translocase, is an RNA helicase. SecA comprises a dimerization C-terminal domain fused to an ATPase N-terminal domain containing conserved DEAD helicase motifs. We show that the N-terminal domain is organized like the motor core of DEAD proteins, encompassing two subdomains, NBD1 and IRA2. NBD1, a rigid nucleotide-binding domain, contains the minimal ATPase catalytic machinery. IRA2 binds to NBD1 and acts as an intramolecular regulator of ATP hydrolysis by controling ADP release and optimal ATP catalysis at NBD1. IRA2 is flexible and can undergo changes in its -helical content. The C-terminal domain associates with NBD1 and IRA2 and restricts IRA2 activator function. Thus, cytoplasmic SecA is maintained in the thermally stabilized ADP-bound state and unnecessary ATP hydrolysis cycles are prevented. Two DEAD family motifs in IRA2 are essential for IRA2–NBD1 binding, optimal nucleotide turnover and polypeptide translocation. We propose that translocation ligands alleviate C-terminal domain suppression, allowing IRA2 to stimulate nucleotide turnover at NBD1. DEAD motors may employ similar mechanisms to translocate different enzymes along chemically unrelated biopolymers.