Download Protein transport across the endoplasmic reticulum membrane

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

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

Document related concepts

Flagellum wikipedia , lookup

Model lipid bilayer wikipedia , lookup

Protein wikipedia , lookup

Membrane potential wikipedia , lookup

Cytosol wikipedia , lookup

P-type ATPase wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Lipid raft wikipedia , lookup

Magnesium transporter wikipedia , lookup

Cytokinesis wikipedia , lookup

Mechanosensitive channels wikipedia , lookup

Theories of general anaesthetic action wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

JADE1 wikipedia , lookup

SNARE (protein) wikipedia , lookup

Signal transduction wikipedia , lookup

Thylakoid wikipedia , lookup

Ribosome wikipedia , lookup

Cell membrane wikipedia , lookup

Western blot wikipedia , lookup

List of types of proteins wikipedia , lookup

Trimeric autotransporter adhesin wikipedia , lookup

Endomembrane system wikipedia , lookup

Transcript
THE SIR HANS KREBS LECTURE
Protein transport across the endoplasmic reticulum
membrane
Delivered on 8 July 2007 at the 32nd FEBS Congress in Vienna,
Austria
Tom A. Rapoport
Howard Hughes Medical Institute and Department of Cell Biology, Harvard Medical School, Boston, MA, USA
Keywords
endoplasmic reticulum; lipid; membrane
integration; protein-conducting channel;
protein translocation; ribosome; Sec
complex; Sec61; SecA; SecY
Correspondence
T. A. Rapoport, Howard Hughes Medical
Institute and Department of Cell Biology,
Harvard Medical School, 240 Longwood
Avenue, Boston, MA 02115, USA
Fax: +1 617 432 1190
Tel: +1 617 432 0676
E-mail: [email protected]
A decisive step in the biosynthesis of many eukaryotic proteins is their partial or complete translocation across the endoplasmic reticulum membrane.
A similar process occurs in prokaryotes, except that proteins are transported across or are integrated into the plasma membrane. In both cases,
translocation occurs through a protein-conducting channel that is formed
from a conserved, heterotrimeric membrane protein complex, the Sec61 or
SecY complex. Structural and biochemical data suggest mechanisms that
enable the channel to function with different partners, to open across the
membrane and to release laterally hydrophobic segments of membrane
proteins into lipid.
(Received 21 May 2008, accepted 18 June
2008)
doi:10.1111/j.1742-4658.2008.06588.x
Introduction
Protein translocation across the eukaryotic endoplasmic reticulum (ER) membrane is a decisive step in
the biosynthesis of many proteins [1]. These include
soluble proteins, such as those ultimately secreted from
the cell or localized to the ER lumen, and membrane
proteins, such as those in the plasma membrane or in
other organelles of the secretory pathway. Soluble
proteins cross the membrane completely and usually
have cleavable N-terminal signal sequences, whose
major feature is a segment of approximately seven to
12 hydrophobic amino acids. Integral membrane
proteins have one or more transmembrane (TM)
segments, each containing approximately 20 hydrophobic amino acids, with intervening hydrophilic regions
on either side of the membrane. Both types of proteins
use the same machinery for transport across the
membrane: a protein-conducting channel. This channel
allows polypeptides to cross the membrane and permits hydrophobic TM segments of membrane proteins
to exit laterally into the lipid phase. In bacteria, the
translocation of secretory and membrane proteins
occurs through a homologous channel in the plasma
membrane, employing signal and TM sequences that
are similar to those in eukaryotes.
The translocation channel is formed from an evolutionarily conserved heterotrimeric membrane protein
Abbreviations
EM, electron microscopy; ER, endoplasmic reticulum; SRP, signal recognition particle; TM, transmembrane.
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
4471
Protein transport across the ER membrane
T. A. Rapoport
complex, called the Sec61 complex in eukaryotes and
the SecY complex in bacteria and archaea [2]. The
a-subunit forms the channel pore, as originally demonstrated by systematic crosslinking experiments [3]. In
addition, reconstitution experiments show that the
Sec61 ⁄ SecY complex is the only essential membrane
component for protein translocation in mammals and
bacteria [4–6]. The channel has an aqueous interior,
as demonstrated by electrophysiology experiments
and measurements of the fluorescence life-time of
probes incorporated into a translocating polypeptide
chain [7–9].
Different modes of translocation
The channel must associate with partners that provide
the driving force for translocation. Depending on the
channel partner, there are different known translocation modes [1]. In co-translational translocation, the
major partner is the ribosome. This translocation
mode is found in all cells of all species, and it is used
for the translocation of both secretory and membrane
proteins. Co-translational translocation begins with a
targeting phase, during which a ribosome-nascent
chain complex is directed to the membrane by the
signal recognition particle (SRP). The ribosome ⁄ SRP ⁄ nascent chain complex is bound to the
membrane, first by an interaction between SRP and its
membrane receptor (SR), and then by an interaction
between the ribosome and the translocation channel
(Fig. 1). The elongating polypeptide chain subsequently moves directly from a tunnel inside the ribosome into the associated membrane channel. GTP
hydrolysis during translation provides the energy for
translocation.
In most, if not all cells, some proteins are translocated
after their completion. This post-translational translocation occurs by different mechanisms in eukaryotes
and bacteria. In yeast (and probably in all eukaryotes),
translocation occurs by a ratcheting mechanism and
involves, as channel partners, the tetrameric Sec62 ⁄ 63
membrane protein complex and the ER luminal
protein BiP, a member of the Hsp70 family of ATPases
(Fig. 2) [10]. Following a transient interaction of BiPATP with the J-domain of Sec63p, ATP is hydrolyzed
and the peptide-binding pocket of BiP closes around
the translocation substrate. BiP serves as a Brownian
ratchet, preventing the bound polypeptide from sliding
ribosome
SRP
signal
sequence
Fig. 1. Model of co-translational translocation. The polypeptide chain moves from
the tunnnel inside the ribosome into the
membrane channel. The energy for
translocation is provided by GTP hydrolysis
during translation. Figure adapted from [1].
cytosol
SRP receptor
Sec61/SecY
complex
signal
sequence
Sec62/63
complex
cytosol
Sec61
complex
J-domain
ADP
ADP
ATP
BiP
ATP
ATP
ADP
ADP
ADP
ADP
ATP
4472
Fig. 2. Model of post-translational translocation in eukaryotes. A Brownian ratcheting
mechanism is responsible for moving a polypeptide chain through the membrane. Translocation might be mediated by oligomers of
the Sec61p complex, as in the other modes
of translocation. Figure adapted from [1].
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
T. A. Rapoport
Protein transport across the ER membrane
back into the cytosol, but allowing polypeptide movement in the forward direction. When moved sufficiently, the next BiP molecule binds, and this process
is repeated until the entire polypeptide chain is translocated. Finally, nucleotide exchange of ADP for ATP
opens the peptide binding pocket and releases BiP
from the polypeptide.
In eubacterial post-translational translocation, polypeptides are ‘pushed’ through the channel by its partner, the cytosolic ATPase SecA (Fig. 3) [11]. SecA has
two nucleotide binding domains (NBD1 and NBD2),
which bind the nucleotide between them and move
relative to one another during the ATP hydrolysis
cycle. Exactly how these movements are used to ‘push’
a polypeptide chain through the channel remains
unknown. The size of SecA makes it unlikely that it
inserts deeply into the SecY channel, as proposed previously [11]. Bacterial translocation in vivo requires an
electrochemical gradient across the membrane, but the
mechanism by which the gradient is utilized is unclear.
Archaea probably have both co- and post-translational translocation. Although co-translational translocation is likely to be similar to that in eukaryotes
and eubacteria, it is not known how post-translational
translocation occurs because archaea lack SecA, the
Sec62 ⁄ 63 complex and BiP.
A
γ
back
front
lateral gate
hinge
B
cytosol
pore
ring
plug
Structure and function of the
translocation channel
The crystal structure of an archaeal SecY complex provides much insight into channel function [2]. The structure is likely representative of all species, as indicated
by sequence conservation and by the similarity to a
lower resolution structure of the Escherichia coli SecY
complex determined by electron microscopy (EM)
from 2D crystals [12]. The a-subunit consists of two
halves, TMs 1–5 and TMs 6–10, which form a lateral
gate at the front and are clamped together at the back
by the c-subunit (Fig. 4A). The 10 helices of the a-subunit form an hourglass-shaped pore that consists of
pore
ring
β
plug
Fig. 4. Structure of the translocation channel. (A) View from the
cytosol of the X-ray structure of the SecY complex from Methanococcus jannaschii. The two halves of SecY are colored blue (TM
1–5) and red (TM 6–10). The plug is shown in yellow and pore ring
residues are shown in green. The purple arrow indicates how the
lateral gate opens. The black arrow indicates how the plug moves
to open the channel across the membrane. (B) Cross-section of the
channel from the side. Figure adapted from [1].
cytoplasmic and external funnels whose tips meet
approximately half way across the membrane
(Fig. 4B). The cytoplasmic funnel is empty, whereas
the external funnel is filled by a short helix, the ‘plug’.
The constriction of the hourglass is formed by a ‘pore
signal
sequence
SecA
ADP
ATP
NBF1
cytosol
Fig. 3. Model of post-translational translocation in bacteria. The ATPase SecA uses the
energy of ATP hydrolysis to push a polypeptide through the channel. Figure adapted
from [1].
SecY
complex
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
4473
Protein transport across the ER membrane
T. A. Rapoport
ring’ of hydrophobic amino acid residues that project
their side chains radially inward. The crystal structure
represents a closed state of the channel, but biochemical data indicate how it can open to translocate
proteins (see below).
Opening the channel across the
membrane
The crystal structure indicates that a single copy of the
Sec61 ⁄ SecY complex forms the pore [2]. The translocation of a secretory protein begins with insertion of a
loop into the channel, such that the signal sequence
is intercalated into the walls of the channel and the
segment distal to it is inserted into the pore proper
(Figs 1–3). In a first step, the binding of a channel
partner (i.e. the ribosome, the Sec62 ⁄ 63p complex, or
SecA) likely weakens interactions that keep the plug in
the center of the Sec61 ⁄ SecY molecule, as indicated by
an increased ion conductance when nontranslating
ribosomes are bound to the channel [7]. Next, the
hydrophobic segment of a signal sequence intercalates
into the lateral gate, between TM2b and TM7, as indicated by photo-crosslinking experiments [13]. This further destabilizes plug interactions, causing the plug to
be displaced from the center of Sec61 ⁄ SecY, as shown
by disulfide bridge crosslinking [14,15]. During subsequent translocation, the signal sequence remains
stationary, whereas the rest of the polypeptide moves
through the pore from the cytoplasmic funnel through
the pore ring into the extracellular funnel (Figs 1–3),
as indicated by systematic disulfide crosslinking experiments [16]. The aqueous interior of the channel and its
shape help to minimize the energy required for the
translocation of a polypeptide through the membrane.
The plug can only return to the center of Sec61 ⁄ SecY
when the polypeptide chain has left the pore.
The diameter of the pore ring, as observed in the
crystal structure, has to increase during translocation,
probably by movements of the helices to which the
pore ring residues are attached. The pore ring is indeed
flexible, as shown by molecular dynamics simulations
and electrophysiology experiments [17,18]. The maximum size of the pore could be 15 · 20 Å, which is
much smaller than the pore size estimated from fluorescence quenching experiments (40–60 Å) [19]. These
data could be reconciled with the crystal structure if
two or more Sec61 ⁄ SecY complexes associated at their
front surfaces, opened their lateral gates, and fused
their pores to form a larger channel. However, disulfide bridge crosslinking experiments argue against
fusion of different pores because they show that,
during SecA-mediated translocation, both the signal
4474
sequence and the mature region of a polypeptide chain
are located in the same SecY molecule [20]. In addition, a detergent-solubilized translocation intermediate
also contains just one copy of SecY associated with
one SecA and one translocation substrate molecule
[21]. Two SecY molecules in a nearly front-to-front
orientation were proposed to be associated with a
translating E. coli ribosome [22]. However, this conclusion was based on a low-resolution ( 15 Å) EM
structure, and the docking of the crystal structure
required its drastic modification. Furthermore, the
position and orientation of both SecY molecules are
different from that of the single SecY molecule
observed in recent EM structures of nontranslating
ribosome-SecY complexes [23] (Fig. 5). It should be
noted that, in the fluorescence quenching experiments,
the fluorescent probes were located deep inside the
ribosome, and therefore the same large diameter
(40–60 Å) must be assumed for the ribosome tunnel, a
size that does not agree with that seen in ribosome
structures (< 20 Å) determined by crystallography
or cryo-EM [24]. Taken together, it is likely that the
translocation pore is formed by just one copy of the
Sec61 ⁄ SecY complex.
Oligomeric translocation channels
Although the pore is formed by only one Sec61 ⁄ SecY
molecule, translocation of a polypeptide chain appears
to be mediated by oligomers. This conclusion is based
on the observation that a SecY molecule defective in
SecA-mediated translocation can be rescued by linking
it covalently with a wild-type SecY copy [20]. Disulfide
bridge crosslinking showed that SecA interacts through
its NBD1 with a nontranslocating SecY copy and
moves the polypeptide chain through a neighboring
SecY copy. The Sec61 ⁄ SecY complex probably forms
oligomers during co-translational translocation as well.
When a ribosome ⁄ nascent chain ⁄ SRP complex binds
to the SRP receptor, a domain of SRP undergoes a
conformational change, exposing a site on the ribosome to which a single Sec61 ⁄ SecY molecule could
bind [25]. This is likely to be the molecule seen in
recent EM structures of complexes of nontranslating
ribosomes with either the SecY or the Sec61 complex
[23,23a]. The bound SecY ⁄ Sec61 molecule is close to
the point where a polypeptide exits the ribosome
and could thus become the translocating copy (Fig. 5).
At a later stage of translocation, SRP completely
detaches from the ribosome, and an additional copy
of the Sec61 ⁄ SecY complex may associate (Fig. 1),
as suggested by crosslinking and freeze-fracture EM
experiments [26,27]. These copies could stabilize the
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
T. A. Rapoport
Protein transport across the ER membrane
SecY
A
tunnel
exit
B
Fig. 5. Structure of the E. coli ribosome-associated SecY channel. (A) Bottom view showing the single copy of the SecY complex that is
bound to a nontranslating ribosome. The docked X-ray structure is indicated by the ribbons. The electron density of the channel is shown in
transparent pink. (B) Comparison of the single-SecY model (left) with a model [22] in which two SecY copies are bound in a near front-tofront orientation to a translating ribosome (right). The lateral gate of the SecY channel is indicated by an arrow, and the tunnel exit by a star.
The smaller picture above shows the orientation of the ribosome.
ribosome–channel junction and possibly recruit other
components, such as signal peptidase and oligosaccharyl transferase, or the translocon-associated protein
complex. Upon termination of translocation, dissociation of the Sec61 ⁄ SecY oligomers could facilitate
the release of the ribosome from the membrane.
Dissociable oligomers may also allow the Sec61 ⁄ SecY
complex to change channel partners and modes of
translocation.
The EM structures of detergent-solubilized ribosome–channel complexes suggested the presence of
three or four Sec61 molecules [28,29]. However, it now
appears that only one Sec61 molecule is present and
that the additional density can be attributed to lipid
and ⁄ or detergent.
Membrane protein integration
During the synthesis of a membrane protein, hydrophobic TM segments move from the aqueous interior
of the channel through the lateral gate into the lipid
phase. The lateral gate may continuously open and
close, exposing polypeptide segments located in the
aqueous channel to the surrounding hydrophobic lipid
phase. Alternatively, there may be a ‘window’ in the
lateral gate that would allow the hydrocarbon chains
of lipids to make contact with a translocating polypeptide at the same time as preventing charged head
groups from entering the channel. Polypeptide segments inside the channel would partition between the
aqueous and hydrophobic environments. This model is
supported by photo-crosslinking experiments [30] and
by the close correlation between a hydrophobicity scale
and the tendency of a peptide to span the membrane
[31]. Hydrophilic segments between the TMs would
alternately move from the ribosome through the aqueous channel to the external side of the membrane, or
emerge into the cytosol between the ribosome and
channel through a ‘gap’ that can be visualized in EM
structures [28,29].
The first TM segment of a membrane protein can
have its N-terminus on either side of the membrane,
depending on the amino acid sequence of the protein,
which often determines the orientation of subsequent
TMs. If the first TM is long and the preceding
sequence not retained in the cytosol by positive
charges or by its folding, the N-terminus can flip
across the channel and subsequently exit laterally into
the lipid phase. When the N-terminus is retained in the
cytosol and the polypeptide chain is further elongated,
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
4475
Protein transport across the ER membrane
T. A. Rapoport
the C-terminus can translocate across the channel,
inserting the polypeptide as a loop, as in the case of a
secretory protein.
Maintaining the permeability barrier
The channel must prevent the free movement of small
molecules, such as ions or metabolites. The crystal
structure suggests a simple model for the maintenance
of the membrane barrier. The resting channel would
be closed, which is consistent with electrophysiology
experiments showing that, in the absence of other components, the SecY channel is impermeable to ions and
water [18]. In the active channel, the pore ring would
fit around the translocating polypeptide chain like a
gasket to restrict the passage of small molecules. The
seal would not be expected to be perfect, but leakage
could be compensated for by powerful ion pumps.
During the synthesis of a multi-spanning membrane
protein, the seal would be provided in an alternating
manner by either the nascent chain in the pore or,
once the chain has left the pore, by the plug returning
to the center of Sec61 ⁄ SecY. Although this model
needs further experimental verification, it would
explain how the membrane barrier can be maintained
in both co- and post-translational translocation.
Surprisingly, plug deletion mutants are viable in
Saccharomyces cerevisiae and E. coli and have only
moderate translocation defects [32–34]. However, the
crystal structures of these mutants show that new plugs
are formed from neighboring polypeptide segments
[34]. The new plugs still seal the closed channel, but
they have lost many interactions that normally keep
the plug in the center of SecY. This results in continuous channel opening and closing, and permits polypeptides with defective or even missing signal sequences to
be translocated. The plug sequences are only poorly
conserved among Sec61 ⁄ SecY channels, supporting the
idea that promiscuous segments can seal the channel
and lock it in its closed state.
Perspective
We are beginning to understand protein translocation
across the eukaryotic ER and bacterial plasma membranes at the molecular level. In particular, progress
during the recent years has led to important insights
into the function of the Sec61 ⁄ SecY channel. Nevertheless, there are major questions in the field that remain
controversial and unresolved, and further progress will
require a combination of approaches. Electrophysiology experiments are needed to complement the fluorescence quenching method, particularly because the
4476
results obtained from the latter are difficult to reconcile with structural data. Important questions with
respect to co-translational translocation include how
the SRP receptor and channel collaborate, how many
Sec61 ⁄ SecY complexes participate in translocation,
and how the ribosome ultimately dissociates from the
channel. Both the precise role of the Sec62 ⁄ 63 components in post-translational translocation and the mechanism by which SecA moves polypeptides need to be
clarified. Membrane protein integration is still particularly poorly understood, and new methods are required
to follow the membrane integration of TMs. Several
other translocation components have been identified,
such as the TRAM protein and the translocon-associated protein complex in mammalian cells, or the YidC
and SecDF proteins in prokaryotes. These components
may be required as chaperones for the folding of TM
segments, or to increase the efficiency of translocation
of some substrates, but their precise functions remain
to be clarified. Much of the progress in the field will
hinge on structural data, with the ‘holy grail’ being
a picture of an active translocon, where a channel
associated with both a partner and a translocating
polypeptide chain is visualized at the atomic level. The
results obtained will likely serve as a paradigm for
other protein translocation systems, such as those in
mitochondria, chloroplasts and peroxisomes.
Acknowledgements
Work in the author’s laboratory was supported by
grants from the National Institute of Health. The
author is a Howard Hughes Medical Institute Investigator. Briana Burton and Sol Schulman are thanked
for critically reading the manuscript.
References
1 Rapoport TA (2007) Protein translocation across the
eukaryotic endoplasmic reticulum and bacterial plasma
membranes. Nature 450, 663–669.
2 van den Berg B, Clemons WM Jr, Collinson I, Modis
Y, Hartmann E, Harrison SC & Rapoport TA (2004)
X-ray structure of a protein-conducting channel. Nature 427, 36–44.
3 Mothes W, Prehn S & Rapoport TA (1994) Systematic
probing of the environment of a translocating secretory protein during translocation through the ER
membrane. EMBO J 13, 3937–3982.
4 Akimaru J, Matsuyama SI, Tokuda H & Mizushima S
(1991) Reconstitution of a protein translocation system
containing purified SecY, SecE, and SecA from Escherichia coli. Proc Natl Acad Sci USA 88, 6545–6549.
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
T. A. Rapoport
5 Brundage L, Hendrick JP, Schiebel E, Driessen AJM
& Wickner W (1990) The purified E. coli integral
membrane protein SecY ⁄ E Is sufficient for reconstitution of SecA-dependent precursor proteintranslocation.
Cell 62, 649–657.
6 Görlich D & Rapoport TA (1993) Protein translocation into proteoliposomes reconstituted from purified
components of the endoplasmic reticulum membrane.
Cell 75, 615–630.
7 Simon SM & Blobel G (1991) A protein-conducting
channel in the endoplasmic reticulum. Cell 65, 371–
380.
8 Crowley KS, Liao SR, Worrell VE, Reinhart GD &
Johnson AE (1994) Secretory proteins move through
the endoplasmic reticulum membrane via an aqueous,
gated pore. Cell 78, 461–471.
9 Crowley KS, Reinhart GD & Johnson AE (1993) The
signal sequence moves through a ribosomal tunnel into
a noncytoplasmic aqueous environment at the ER
membrane early in translocation. Cell 73, 1101–1115.
10 Matlack KE, Misselwitz B, Plath K & Rapoport TA
(1999) BiP acts as a molecular ratchet during posttranslational transport of prepro-a factor across the
ER membrane. Cell 97, 553–564.
11 Economou A & Wickner W (1994) SecA promotes
preprotein translocation by undergoing ATP-driven
cycles of membrane insertion and deinsertion. Cell 78,
835–843.
12 Breyton C, Haase W, Rapoport TA, Kuhlbrandt W &
Collinson I (2002) Three-dimensional structure of the
bacterial protein-translocation complex SecYEG.
Nature 418, 662–665.
13 Plath K, Mothes W, Wilkinson BM, Stirling CJ &
Rapoport TA (1998) Signal sequence recognition in
posttranslational protein transport across the yeast ER
membrane. Cell 94, 795–807.
14 Harris CR & Silhavy TJ (1999) Mapping an interface
of SecY (PrlA) and SecE (PrlG) by using synthetic
phenotypes and in vivo cross-linking. J Bacteriol 181,
3438–3444.
15 Tam PC, Maillard AP, Chan KK & Duong F (2005)
Investigating the SecY plug movement at the SecYEG
translocation channel. EMBO J 24, 3380–3388.
16 Cannon KS, Or E, Clemons WM Jr, Shibata Y &
Rapoport TA (2005) Disulfide bridge formation
between SecY and a translocating polypeptide localizes
the translocation pore to the center of SecY. J Cell
Biol 169, 219–225.
17 Gumbart J & Schulten K (2006) Molecular dynamics
studies of the archaeal translocon. Biophys J 90, 2356–
2367.
18 Saparov SM, Erlandson K, Cannon K, Schaletzky J,
Schulman S, Rapoport TA & Pohl P (2007) Determining the conductance of the SecY protein trans-
Protein transport across the ER membrane
19
20
21
22
23
23a
24
25
26
27
28
29
30
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS
location channel for small molecules. Mol Cell 26,
501–509.
Hamman BD, Chen JC, Johnson EE & Johnson AE
(1997) The aqueous pore through the translocon has a
diameter of 40-60Å during cotranslational protein
translocation at the ER membrane. Cell 89, 535–544.
Osborne AR & Rapoport TA (2007) Protein translocation is mediated by oligomers of the SecY
complex with one SecY copy forming the channel.
Cell 129, 97–110.
Duong F (2003) Binding, activation and dissociation
of the dimeric SecA ATPase at the dimeric SecYEG
translocase. EMBO J 22, 4375–4384.
Mitra K, Schaffitzel C, Shaikh T, Tama F, Jenni S,
Brooks CL 3rd, Ban N & Frank J (2005) Structure of
the E. coli protein-conducting channel bound to a
translating ribosome. Nature 438, 318–324.
Menetret JF, Schaletzky J, Clemons WM Jr, Osborne
AR, Skanland SS, Denison C, Gygi SP, Kirkpatrick
DS, Park E, Ludtke SJ et al. (2007) Ribosome binding
of a single copy of the SecY complex: implications for
protein translocation. Mol Cell 28, 1083–1092.
Menetret JF, Hedge RS, Aguiar M, Gygi SP, Park E,
Rapaport TA & Akey CW (2008) Single copies of
Sec61 and TRAP associate with a nontranslating
mammalian ribosome. Structure 16, 1126–1137.
Voss NR, Gerstein M, Steitz TA & Moore PB (2006)
The geometry of the ribosomal polypeptide exit tunnel.
J Mol Biol 360, 893–906.
Halic M, Gartmann M, Schlenker O, Mielke T, Pool
MR, Sinning I & Beckmann R (2006) Signal recognition particle receptor exposes the ribosomal translocon
binding site. Science 312, 745–747.
Schaletzky J & Rapoport TA (2006) Ribosome binding to and dissociation from translocation sites of the
endoplasmic reticulum membrane. Mol Biol Cell 17,
3860–3869.
Hanein D, Matlack KES, Jungnickel B, Plath K,
Kalies K-U, Miller KR, Rapoport TA & Akey CW
(1996) Oligomeric rings of the Sec61p complex induced
by ligands required for protein translocation. Cell 87,
721–732.
Beckmann R, Spahn CM, Eswar N, Helmers J, Penczek PA, Sali A, Frank J & Blobel G (2001) Architecture of the protein-conducting channel associated with
the translating 80S ribosome. Cell 107, 361–372.
Menetret JF, Hegde RS, Heinrich SU, Chandramouli
P, Ludtke SJ, Rapoport TA & Akey CW (2005)
Architecture of the ribosome-channel complex derived
from native membranes. J Mol Biol 348, 445–457.
Heinrich SU, Mothes W, Brunner J & Rapoport TA
(2000) The Sec61p complex mediates the integration of
a membrane protein by allowing lipid partitioning of
the transmembrane domain. Cell 102, 233–244.
4477
Protein transport across the ER membrane
T. A. Rapoport
31 Hessa T, Kim H, Bihlmaier K, Lundin C, Boekel J,
Andersson H, Nilsson I, White SH & von Heijne G
(2005) Recognition of transmembrane helices by the
endoplasmic reticulum translocon. Nature 433,
377–381.
32 Junne T, Schwede T, Goder V & Spiess M (2006) 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.
4478
33 Maillard AP, Lalani S, Silva F, Belin D & Duong F
(2007) Deregulation of the SecYEG translocation
channel upon removal of the plug domain. J Biol
Chem 282, 1281–1287.
34 Li W, Schulman S, Boyd D, Erlandson K, Beckwith J
& Rapoport TA (2007) 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.
FEBS Journal 275 (2008) 4471–4478 ª 2008 The Author Journal compilation ª 2008 FEBS