Download ref. #27 of the TIBS article

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

Genetic code wikipedia , lookup

Lipid raft wikipedia , lookup

Membrane potential wikipedia , lookup

Lipid bilayer wikipedia , lookup

Bottromycin wikipedia , lookup

QPNC-PAGE wikipedia , lookup

Protein wikipedia , lookup

Magnesium transporter wikipedia , lookup

Protein adsorption wikipedia , lookup

Ribosomally synthesized and post-translationally modified peptides wikipedia , lookup

Theories of general anaesthetic action wikipedia , lookup

Model lipid bilayer wikipedia , lookup

List of types of proteins wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

NADH:ubiquinone oxidoreductase (H+-translocating) wikipedia , lookup

Protein–protein interaction wikipedia , lookup

Thylakoid wikipedia , lookup

Mechanosensitive channels wikipedia , lookup

Proteolysis wikipedia , lookup

Two-hybrid screening wikipedia , lookup

Protein domain wikipedia , lookup

Endomembrane system wikipedia , lookup

Cell membrane wikipedia , lookup

Cyclol wikipedia , lookup

SNARE (protein) wikipedia , lookup

Western blot wikipedia , lookup

Protein structure prediction wikipedia , lookup

Cell-penetrating peptide wikipedia , lookup

Trimeric autotransporter adhesin wikipedia , lookup

P-type ATPase wikipedia , lookup

Transcript
Article No. mb982217
J. Mol. Biol. (1998) 284, 1165±1175
Proline-induced Disruption of a Transmembrane
a -Helix in its Natural Environment
IngMarie Nilsson1,2, Annika SaÈaÈf1, Paul Whitley3, Guro Gafvelin3
Cecilia Waller1 and Gunnar von Heijne1*
1
Department of Biochemistry
Stockholm University, S-106 91
Stockholm, Sweden
2
Department of Biosciences
Karolinska Institute, NOVUM
S-141 57 Huddinge, Sweden
3
Deptartment of Laboratory
Medicine, Divsion of Clinical
Immunology, Karolinska
Hospital, S-171 76 Stockholm
Sweden
a-Helix formation in globular proteins has been studied both theoretically
and experimentally for decades, while a lack of both high-resolution
structures and suitable experimental techniques has hampered the study
of helices in membrane proteins. We have developed a new experimental
approach, glycosylation mapping, where the active site of the lumenally
exposed endoplasmic reticulum enzyme oligosaccharyl transferase is
used as a point of reference against which the position of a transmembrane segment in the membrane can be measured. Here, we report an
initial analysis of the helix-breaking properties of proline residues
inserted in a transmembrane helix. We ®nd that proline residues can
break a transmembrane helix, but only when inserted near the end, and
only when the helix is suf®ciently long. The glycosylation mapping technique may be generally useful for determining the position of transmembrane helices in the membrane.
# 1998 Academic Press
*Corresponding author
Keywords: membrane protein; protein structure; glycosylation;
transmembrane helix; proline
Introduction
Integral membrane proteins are thought to come
in two basic varieties, built, respectively, on a helix
bundle and a b-barrel architecture (von Heijne,
1996). The helix bundle design with its long,
hydrophobic transmembrane a-helices (TMHs) is
particularly simple, yet detailed studies of the conformational properties of individual amino acid
residues in such transmembrane segments and of
the effects of different residues on the segments
position in the lipid bilayer are dif®cult, since
neither X-ray crystallography nor NMR are easily
applied to membrane proteins.
Previous studies of helix formation in a non-aqueous environment have mainly been carried out
with puri®ed peptides dissolved in detergents or
incorporated into liposomes (Deber & Goto, 1996;
Li & Deber, 1994; Li et al., 1996; Liu et al., 1996;
Papavoine et al., 1997; Shon et al., 1991; Stopar et al.,
1996). We have developed an alternative approach
Abbreviations used: TMHs, transmembrane a-helices;
ER, endoplasmic reticulum; OST, oligosaccharyl
transferase; MGD, minimum glycosylation distance;
Lep, leader peptidase.
E-mail address of the corresponding author:
[email protected]
0022±2836/98/491165±11 $30.00/0
where the active site of the lumenally exposed
endoplasmic reticulum (ER) enzyme oligosaccharyl
transferase (OST) is used as a point of reference
against which positional alterations in a TMH
embedded in the ER membrane can be measured
with high precision. As a ®rst application of this
technique, we have investigated the conformational effects of proline residues placed in different
positions in a poly-Leu TMH. A calibration of the
method using the TMH from the H-subunit of the
Rhadobacter sphaeroides photosynthetic reaction center makes it possible to deduce the position of
other TMHs relative to the lipid bilayer from glycosylation mapping data.
Results
Glycosylation mapping
The basic idea behind the glycosylation mapping
technique is depicted in Figure 1(a). As shown by
(Nilsson & von Heijne (1993), the OST can transfer
a glycosyl moiety to an acceptor Asn residue in a
nascent membrane protein only when the Asn-X(Thr/Ser) acceptor site is placed a minimum number of residues away from either end of a TMH,
the ``minimum glycosylation distance'' (MGD). N
and C-terminal MGD values (MGDN, MGDC) can
# 1998 Academic Press
1166
Glycosylation Mapping of Transmembrane Helices
Figure 1. Glycosylation mapping
of constructs derived from leader
peptidase (Lep) (Wolfe et al., 1983).
(a) Potential Asn-Ser-Thr acceptor
sites for N-linked glycosylation
(Y ˆ glycosylated acceptor site, =
non-glycosylated; acceptor site) are
introduced in different positions
relative to the H1 and H2 transmembrane segments, as well as to
a segment (H3) inserted into the P2
domain by site-directed mutagenesis. Proteins are expressed in vitro
in the presence of dog pancreas
microsomes, and the degree of
glycosylation is determined as a
function of d, the number of residues between a chosen reference
residue at the end of the hydrophobic segment and the acceptor
Asn. The minimal glycosylation
distance (MGD) is the number of
residues required for half-maximal
glycosylation. (b) The gels show
the results obtained from in vitro
translation in the absence (ÿ) and
presence (‡) of rough microsomes (RM) of constructs with the n ˆ 23 poly-Leu transmembrane segment replacing
the H2 segment in Lep and with the acceptor-site Asn located at d ˆ 9, 10, 11, and 12 residues (counting from the
®rst Gln after the hydrophobic stretch). Filled and empty circles indicate the non-glycosylated and glycosylated forms
of the protein, respectively. The MGD value (de®ned as the value of d for which the glycosylation ef®ciency is 40%,
i.e. one-half of the maximal glycosylation ef®ciency which is 80%) is determined from the graph by interpolation
and is shown above the sequence of the transmembrane segment.
be measured with high precision using in vitro
translation in the presence of dog pancreas microsomes of mutant proteins with differently placed
glycosylation acceptor sites (Figure 1(b)). Changes
in the position of a TMH relative to the membrane
(or, more accurately, relative to the active site of
the membrane-bound OST enzyme) caused by
mutations in the hydrophobic stretch should be
re¯ected in proportional changes in the MGDN and
MGDC values, thus providing a sensitive assay for
studying the conformational role of different
amino acids in TMHs.
In the studies reported here, we have replaced
the ®rst (H1) and second (H2) transmembrane
domains in the well-characterized Escherichia coli
inner membrane protein leader peptidase (Lep;
Figure 1(a) left-hand panel), with either poly-Leu
based model transmembrane segments or with the
N-terminal TMH from the R. sphaeroides photosyn-
thetic reaction center. We have also inserted polyLeu sequences in the middle of the lumenal P2
domain (right-hand panel). Although of bacterial
origin, Lep integrates ef®ciently into dog pancreas
microsomes with the same topology as in E. coli
(Johansson et al., 1993; Nilsson & von Heijne,
1993).
To facilitate the analysis, a set of Lep-encoding
plasmids differing only in the position of the glycosylation acceptor site Asn-Ser-Thr upstream of H1
and H3 or downstream of H2 was constructed,
and MGD values were determined for each transmembrane segment by cloning it into the appropriate plasmids and determining the number of
residues between the N or C-terminal end of the
hydrophobic stretch and the acceptor site required
for half-maximal glycosylation. Since proline residues are known to reduce the ef®ciency of glycosylation when present either immediately upstream,
Glycosylation Mapping of Transmembrane Helices
1167
in the middle, or immediately downstream of an
Asn-X-Thr/Ser acceptor site (Mellquist et al., 1998;
Shakineshleman et al., 1996; our unpublished
observations), all such Pro residues in Lep were
changed to Gln in the engineered acceptor sites.
MGD is a function of the length of the
transmembrane segment
We have shown that MGDC values vary with
the length of the hydrophobic transmembrane segment (Nilsson et al., 1994). To quantify this effect,
H2 was replaced by a stretch of n leucine residues
and one valine side ¯anked by four N-terminal
lysine residues and a C-terminal Gln-Gln-Gln-Pro
segment, and MGDC (counting from the ®rst ¯anking Gln residue) was determined for different
n-values. As seen in Figure 2(a), there is a roughly
linear correlation between n and MGDC for
12 < n < 23 whereas MGDC 12.5 for n 4 12 and
9.5 for n 5 23.
Similarly, the variation of MGDN with the length
of the hydrophobic region was determined.
A stretch of n leucine residues (11 4 n 4 29) and
one valine residue, ¯anked by an N-terminal SerGln-Gln-Gln segment and four C-terminal lysine
residues was inserted in place of the N-terminal
transmembrane segment (H1) or in the middle of
the large P2 domain in Lep (H3). In both cases, a
roughly linear dependence of MGDN on n was
found for the entire range of n-values tested
(Figure 2(a)), although there is a slight upwards
curvature for the shortest H3-constructs. As was
noted (Nilsson & von Heijne, 1993), MGDN values
are consistently three to ®ve residues larger than
the corresponding MGDC values, suggesting that
the OST active site is oriented with the Asn-binding pocket closer to the membrane surface than the
Ser/Thr-binding pocket.
Since Lep uses the SRP/Sec61-translocon pathway for membrane assembly when expressed in
the microsomal in vitro system (Mothes et al.,
1997), we also wanted to test the relation between
MGD and n for a protein that does not use this
pathway. Synaptobrevin is known to insert a
hydrophobic C-terminal segment into the microsomal membrane by an as yet poorly understood
mechanism that does not involve the SRP/Sec61
machinery (Kutay et al., 1995), and it can be glycosylated on its lumenal, C-terminal tail even when
the single transmembrane domain is replaced by
poly-Leu stretches (Whitley et al., 1996a). We thus
determined MGDC vaules for four different constructs derived from human synaptobrevin 2 with
poly-Leu transmembrane segments (n ˆ 11, 14, 17
and 20); since synaptobrevin is not very ef®ciently
targeted to microsomes, all quanti®cations were
done with carbonate-extracted membrane pellets.
The results for the n ˆ 17 constructs are shown in
Figure 2(b). Although the MGDC values were
found to be increased by about 1.5 residues compared to the H2 data (Figure 2(a)), their dependence on n was virtually identical to that seen for
Figure 2. MGD values vary with the length of the
hydrophobic segment. (a) MGDN and MGDC values are
shown as a function of the number of leucine residues
(n) in poly-Leu transmembrane segments of the
general sequence LISQQQLnVKKKKHM (H1, H3) or
PGLIKKKKLnVQQQP (H2). MGD-values are counted
from the ®rst Gln residue before (MGDN) or after
(MGDC) the hydrophobic stretch. Filled circles,
MGDN-values for H1-poly-Leu segments; open squares,
MGDC-values for H2-poly-Leu segments; open circles,
MGDN-values for H3-poly-Leu segments; ®lled squares,
MGDC-values for poly-Leu segments in the C-terminal
tail of synaptobrevin. For synaptobrevin, MGDC-values
were based on 30% glycosylation since the maximal glycosylation ef®ciency is only 60±65%. (b) Glycosylation of
synaptobrevin-derived constructs with a transmembrane
segment of the composition . . . KKKKLnVQQQP . . . and
glycosylation acceptor sites placed 11, 13, and 15 residues downstream of the hydrophbobic stretch (counting
from the ®rst ¯anking Gln). Microsomes were subjected
to a carbonate wash procedure to remove soluble and
extrinsic membrane proteins before loading onto the gel.
Open and ®lled crcles indicate the non-glycosylated and
glycosylated forms of the protein, respectively.
the Lep-constructs. The MGDN,C f(n) relationships thus appear to be independent of the mode
of membrane assembly, and presumably re¯ect
interactions between the transmembrane segments
and the lipid bilayer rather than translocon-speci®c
interactions. In all cases, an increase in n by four
residues leads to a reduction in MGD by about one
residue. A possible interpretation of this effect is
presented in the Discussion; for now we simply
1168
Glycosylation Mapping of Transmembrane Helices
take the MGDN,C f(n) relationships as calibration
curves (see below).
Helix-breaking effects of C-terminal
proline residues
Figure 3. Proline residues disrupt transmembrane
helices. (a) MGDC values (counting from Gln ÿ 1) for
H2 transmembrane segments with n ˆ 8, 11, 14, 20, 23,
and 29 leucine residues and with proline residue replacement mutations in the indicated positions. The no P
value is for a segment without proline reidues (cf.
Figure 2(a)), and 23L P6 indicates the MGD-value for a
23L double mutant with proline residues in positions 3
and 6. Position ÿ1 is the Gln residue immediately
C-terminal to the hydrophobic segment, and position 1
is the Val residue at the C-terminal end of this segment
(i.e. proline residue positions are counted in the C to
N-terminal direction). The position in the membrane of
the n ˆ 23 (Leu5 ! Pro) mutant relative to the n ˆ 18
construct is shown below the graph (see the text).
To study the conformational effects of proline
residues introduced near the C-terminal end of a
TMH, a number of residues throughout the H2
n ˆ 23 poly-Leu transmembrane stretch were individually substituted by proline, and MGDC was
determined for each mutant. As shown in
Figure 3(a), the exchange of Gln ÿ 1 or Val1 for
Pro had little effect on MGDC, whereas the
Leu2 ! Pro substitution led to a substantial
decrease in MGDC from 9.7 to 8.5 residues. The
effect on MGDC persisted six residues into the
hydrophobic segment with a minimum value of
MGDC ˆ 7.2 for the Leu5 ! Pro substitution, but
then quickly disappeared as the proline was
moved beyond Leu6.
A ®rst interpretation of these results is that the
stretch of residues between the end of the hydrophobic segment and the OST active site has a ¯exible, extended conformation that is not affected by
the introduction of a helix-breaking proline (mutants Gln ÿ 1 ! Pro, Val1 ! Pro). The
Leu2 ! Pro substitution leads to termination of
the TMH at the proline, thus pushing one residue
out of the membrane to become part of the ¯exible
segment. When the proline is inserted deep into
the transmembrane segment (beyond Leu6), there
is little or no effect on MGDC and hence the proline
does not break the helix.
The interpretation is somewhat more complicated for the Leu3 ! Pro to Leu6 ! Pro mutants,
since MGDC varies with n for transmembrane segments with n 4 23 (Figure 2(a)). Thus, if the helix
in the Leu5 ! Pro mutant were to end at Leu6,
this would correspond to a TMH with n ˆ 18.
From Figure 2(a), the n ˆ 18 construct has
MGDC ˆ 11.5 residues (counting from Gln ÿ 1).
The Leu5 ! Pro mutant has MGDC ˆ 7.2 residues
(or 12.2 counting from Pro5), suggesting that the
helix indeed terminates around Pro5, as illustrated
in Figure 3(a).
As a further test, we made a double mutant with
Pro both in positions 3 and 6 (denoted 23LP6 in
Figure 3(a)); in this case, we found MGDC ˆ 5.1
(or MGDC ˆ 11.1 counting from Pro6). From
Residues assumed to be in a helical conformation are
shown in upper case, those in a more ¯exible, extended
conformation in lower case. (b) MGDN values (counting
from Gln ÿ 1) for H1 and H3 transmembrane segments
with n ˆ 23 leucine residues and with proline replacement mutations in the indicated positions. The no P
value is for a segment without proline (cf. Figure 2(a)).
Position ÿ1 is the Gln residue immediately N-terminal
to the hydrophobic segment, and position 1 is the Leu
residue at the N-terminal end of this segment (i.e. proline residue positions are counted in the N to C-terminal
direction).
Glycosylation Mapping of Transmembrane Helices
Figure 2(a), if the transmembrane segment ends at
Leu7 in this case (i.e. n ˆ 17), we should have
MGDC 11.7. It is thus apparent that the entire 23residue long TMH is either beginning to re-form in
the Leu6 ! Pro mutant, or that the short hydrophobic stretch between Leu5 and Val1 somehow
dips back into the membrane; in either case, the
introduction of the additional Pro3 mutation
obviously pushes the whole stretch towards the
lumen.
In summary, we conclude that a proline inserted
into the C-terminal 1.5 turns of the n ˆ 23 TMH
has a strongly disrupting effect, whereas more deeply buried proline residues do not break the helix.
MGDC-values were also measured for proline
substitutions in H2 transmembrane segments composed of one valine and n ˆ 8, 11, 14, 20, and 29
leucine residues, (Figure 3(a)); i.e. starting from the
shortest poly-Leu stretch known to function as an
ef®cient signal-anchor sequence in the microsomal
system (Sakaguchi et al., 1992). In all cases except
n ˆ 8 and 11, the MGDC-values dropped by about
one residue in the Val1 ! Pro mutant and by
about two residues in the Leu2 ! Pro mutant, consistent with a disruption of the TMH. Interestingly,
Leu ! Pro substitutions deeper into the hydrophobic stretch were more disruptive in the longer
segments: the Leu3 ! Pro mutation had only a
weak disruptive effect for n ˆ 14, whereas a signi®cant disruptive effect was apparent even for the
Leu9 ! Pro substitution in the n ˆ 29 construct.
Leu ! Pro substitutions had negligible effects on
MGDC for n ˆ 8.
The observation that proline residues have no
effect on MGDC for n ˆ 8 and only strongly affect
MGDC for the n ˆ 14 segment in positions Val1
and Leu2 but not Leu3 further suggests that a
Ê long)
``core helix'' of about 13 residues (i.e. 20 A
is always present, irrespective of amino acid comÊ
position. This ®ts nicely with the roughly 20 A
wide central hydrophobic core evident in the crystal stuctures of the photosynthetic reaction center
and mitochondrial cytochrome c oxidase (Wallin
et al., 1997), where disruption of the helical conformation would be very costly (Popot & Engelman,
1990). It also ®ts well with a recent molecular
dynamics simulation of a transmembrane poly-Ala
helix embedded in a DMPC bilayer, where the central 12 residues were found to form a very stable
a-helix (Shen et al., 1997).
Helix-breaking effects of N-terminal
proline residues
We also tested the effect on MGDN of a proline
residue inserted near the N-terminal end of a
transmembrane segment. A stretch of 23 leucine
residues and one valine, ¯anked by an N-terminal
Ser-Gln-Gln-Gln segment and four C-terminal
lysine residues was inserted in place of the
N-terminal transmembrane segment (H1) in
Lep (Figure 1(a)). As shown in Figure 3(b), the
Gln ÿ 1 ! Pro substitution reduced MGDN by
1169
about two residues, suggesting that the helix normally extends almost to the N-terminal end of the
Ser-Gln-Gln-Gln segment: conversion of four helical residues to an extended conformation will
increase
the
length
spanned
by
about
Ê , which corresponds to a
4 (3.3 ÿ 1.5) ˆ 7.2 A
movement of the glycosylation acceptor site by
about 8/3.3 ˆ 2.2 residues relative to the OST
active site; also, Ser is a common N-cap and Gln a
common N3 residue in helices in globular proteins
(Harper & Rose, 1993). The MGDN value was
reduced by another 1.6 residues in the Leu1 ! Pro
mutant (which effectively shortens the hydrophobic stretch by one residue), and then increased
as the proline was moved further into the hydrophobic stretch.
Again, a detailed interpretation requires that the
variation in MGDN with the length of a transmembrane segment is taken into account. As shown in
Figure 2(a), MGDN increases by about 0.25 residues for every leucine residue removed from the
transmembrane segment. If the Leu1 ! Pro
mutation were to lead to termination of the helix at
the proline residue, the expected increase in MGDN
compared to the Gln ÿ 1 ! Pro mutant would
thus be 1.25 residues, in reasonable agreement
with the observed value. For the Leu3 ! Pro and
Leu6 ! Pro mutations, the observed values would
suggest that the helix ends near Leu2 in both cases.
This is consistent with the obsevation that proline
resudes are often found in the three most N-terminal positions in a-helices both in globular and integral membrane proteins but not further in
(Richardson & Richardson, 1988; Wallin et al.,
1997). Thus, N-terminal proline residues also seem
to cause a break in the n ˆ 23 TMH when present
in the ®rst one to two helical turns.
Similar MGDN values were obtained when proline residue insertions were made in a 23-residues
long poly-leucine transmembrane stretch (H3)
placed in the middle of the large C-terminal
domain, Figure 3(b), demonstrating that the
location within the protein does not much in¯uence the results.
Helix-breaking effects of N-terminal proline
residues assayed in E. coli
Since the immediate environment of a TMH in
the ER membrane at the time of glycosylation is
dif®cult to characterize in detail (see Discussion),
we also carried out protease-protection experiments in E. coli. To this end, we expressed a Lepderived construct where H1 has been replaced by a
20-residues long (Ala, Leu) stretch (Whitley et al.,
1994) and where H2 has the composition L7ML15,
together with two further mutants in H2 with the
substitutions Leu5 ! Pro and Leu10 ! Pro (counting from the N-terminal end of the hydrophobic
stretch), and probed the location of the N-terminal
end of H2 relative to the inner membrane by proteinase K digestion of spheroplasts (Figure 4). In
this case, there is little doubt that the TMH is
1170
Figure 4. Disruption of a poly-Leu transmembrane helix by proline residues assayed in E. coli. Lepderived constructs with an ``inverted'' topology, H1
replaced by a 20-residues long (Ala, Leu) stretch
(Whitley et al., 1994), and H2 replaced either by the
sequence . . . N3L7ML15K4 . . . (lanes 1, 4), by this sequence
with a Leu5 ! Pro replacement (counting from the
N-terminal leucine; lanes 2, 5), or a Leu10 ! Pro replacement (lanes 3, 6) were expressed in E. coli and labelled
for one minute by [35S]Met. Cells were converted to
spheroplasts, digested with proteinase K, lysed, and
subjected to immunoprecipitation with a Lep antibody.
The protease-protected fragments (lanes 4-6) represent
the H2-P2 domain. The faint lower Mr band in lanes 1-3
probably represents an endogenous proteolysis product
cleaved in the periplasmic loop.
embedded in the lipid bilayer when the experiment
is performed. All three Lep constructs expose a
short, protease-sensitive loop between H1 and H2
to the periplasm, and proteinase K digestion will
produce a protected fragment composed of the H2P2 domain (Whitley et al., 1994). Signi®cantly, the
protease-protected fragment in the Leu5 ! Pro
mutant was considerably smaller than in the
construct without proline residues or in the
Leu10 ! Pro mutant (Figure 4; lanes 4-6), indicating that the N-terminal part of H2 becomes more
accessible to protease added from the periplasmic
side when the proline is placed in the ®rst two
turns of the helix. Although much less precise than
the glycosylation mapping measurements, the protease accessibility assay thus yields similar results,
strengthening the assumption that the TMH is at
least partly embedded in the lipid bilayer when
glycosylation takes place.
The photosynthetic reaction center H-subunit
transmembrane helix
In order to calibrate the glycosylation mapping
results against a TMH with known 3D structure,
we inserted the N-terminal TMH (plus some ¯ank-
Glycosylation Mapping of Transmembrane Helices
ing residues) from the H-subunit of the R. sphaeroides photosynthetic reaction center (Allen et al.,
1987; Rees et al., 1989) in place of the H1 or H2
transmembrane segment. Judging from the X-ray
structure of the reaction center (Figure 5(a)), the
H-subunit helix does not interact extensively with
the other transmembrane helices and extends
roughly between Leu12 and Thr33, which are
probably located in the lipid headgroup regions
(Wallin et al., 1997).
As illustrated in Figure 5(b), the N-terminal
MGDN was found to be 13.2 residues (counting
from Asp11) and the C-terminal MGDC was found
to be 10.1 residues (counting from Glu34). To
assess the effect of proline insertions on MGDC, we
changed Tyr30 to Pro; for this mutant we found an
MGDC value of only 4.6 residues (8.6 residues
counting from Pro30), i.e., a decrease of 5.5 residues. The helix-breaking effect of the proline is
thus even stronger in this context, possibly because
the residues C-terminal to the proline residue are
less hydrophobic in the H-subunit helix than in the
poly-Leu helices and thus more easily pushed out
of the membrane.
In order to provide an independent check on
the relation between MGD and the length of the
transmembrane segment (Figure 2(a)), two further
mutations were made in the H-subunt helix
placed in the H1 location: one insertion and one
deletion, each of four hydrophobic residues. The
MGDN values were found to be 12.1 and 13.7
residues, respectively (data not shown), corresponding to a decrease in MGDN of 0.8 residues
for every four residues added to the transmembrane segment. This is in reasonable agreement
with the value found above for the H1 poly-Leu
segment (decrease of one residue for every four
added; Figure 2(a)), and suggests that the results
obtained with the poly-Leu model sequences are
representative also for natural transmembrane
segments.
The results for the H-subunit helix allow the
n ˆ 20 poly-Leu helix to be positioned relative to
the H-subunit helix (which contains a 20-residue
long hydrophobic stretch between Leu12 and
Leu31; Figure 5(b) (bottom). The MGDN-vaule for
the H1 n ˆ 20 construct is 16.3, but, as shown
above, this value should be reduced by about
two to correct for the assumed N-terminal extension of the helix through the Ser-Gln-Gln-Gln segment. This predicts that the N-terminal end of
the n ˆ 20 poly-Leu stretch is located at about
the same position in the membrane as Ala13 in
the H-subunit helix. The MGDC vaule for the
n ˆ 20 construct is 11.1 residues, roughly placing
the valine at the end of the poly-Leu stretch in
an equivalent position to Leu31 in the H-subunit
helix. The location of the poly-Leu transmembrane segment relative to the H-subunit helix
deduced from the glycosylation mapping data
is thus as expected from their hydrophobicity
pro®les.
1171
Glycosylation Mapping of Transmembrane Helices
Discussion
We have developed a new method, glycosylation
mapping, that makes it possible to study the position of transmembrane helices in a natural membrane. Our results do not directly address the
question of the molecular environment of the TMH
at the time of glycosylation, although they are consistent with a predominantly non-polar milieu.
Since glycosylation near the MGDN of a transmembrane hydrophobic stretch is observed even when
truncated mRNA transcripts lacking a stop-codon
are used (Whitley et al., 1996b), it is clear that the
modi®cation can take place within the context of
the ER translocon, which is known to include both
the Sec61 translocation complex and the oligosaccharyl transferase (GoÈrlich & Rapoport, 1993;
Rapoport et al., 1996). The recent demonstration
that hydrophobic transmembrane segments in general (and the Lep H1 segment in particular)
become exposed to lipids very soon after entering
the translocation channel (Martoglio et al., 1995;
Mothes et al., 1997) nevertheless suggests that the
MGD-values measured here pertain to a situation
where the TMH is in a predominantly lipidic
environment. This conclusion is further supported
by the observation that the variation in MGD
vaules with the length of the poly-Leu transmembrane stretch is essentially the same, irrespective of
whether or not membrane assembly proceeds
through the SRP/Sec61-pathway (Figure 2(a)).
A ®nal argument in favor of a lipidic environment
is the observation that Leu ! Pro mutations have
similar effects on the position of a transmembrane
segment in the membrane as probed by glycosylation mapping in the ER membrane or by protease
trimming in the inner membrane of E. coli
(Figures 3 and 4).
In a ®rst detailed application of the glycosylation
mapping approach, we have found that proline
residues break a poly-Leu TMH when inserted as
far as one to two helical turns from the ends of the
hydrophobic stretch. In contrast to what has been
found for globular proteins (Barlow & Thornton,
1988), proline residues appear not to have any
gross conformational effects when placed more
centrally in the TMH, as also observed in prolinecontaining transmembrane helices in proteins of
known 3D structure such as the photosynthetic
reaction center, bacteriorhodopsin, and cytochrome
c oxidase (Iwata et al., 1995; Sansom, 1992;
Tsukihara et al., 1996; von Heijne, 1991; Wallin
Figure 5. MGD-determination for the R. sphaeroides
reaction center H-subunit transmembrane helix.
(a) Location of the H-subunit transmembrane helix in
the reaction center. Leu12 and Thr33 are shown as CPK
models. Co-ordinates are from the PDB ®le 2RCR
(Grigorieff et al., 1996) and the plot was made using
MOLSCRIPT (Kraulis, 1991). (b) Glycosylation ef®ciency
for acceptor sites located at different distances d from
the H-subunit helix placed in the H2 position and for
the indicated Tyr30 ! Pro mutant (counting from
Glu34; note that the MGDC value for the latter is determined by extrapolation from the d ˆ 5 construct). The
position in the membrane relative to the H-subunit
transmembrane helix of the n ˆ 20 poly-Leu transmembrane helix is shown at the bottom. Residues assumed
to be in a helical conformation are shown in upper case,
those in a more ¯exible, extended conformation in lower
case.
1172
et al., 1997). The helix-disrupting effect is only seen
when the hydrophobic segment is longer than 13
residues, suggesting that a core helix of this length
will always be present in any transmembrane segment, irrespective of amino acid sequence.
By calibrating our measurements against the
TMH from the H-subunit of the photosynthetic
reaction center, we have found that a model transmembrane segment composed of one valine and 20
leucine residues is located in a position in the
membrane equivalent to that occupied by the
Ala13-Leu31 segment in the reaction center H-subunit (as shown in the accompanying paper; MonneÂ
et al., 1998), very similar results are obtained with
the M13 coat protein TMH which also has a
known position in the bilayer). This not only
reinforces our structural interpretation of the data,
but also suggests that glycosylation mapping may
be used as a general method for determining the
position of transmembrane helices in the membrane. It has recently been shown that analysis of a
series of glycosylation acceptor sites placed
throughout a loop between two transmembrane
helices can give a rough idea of where the helix
ends are located (Popov et al., 1997); with the
increased precision obtainable from a comparison
to the H-subunit and phage M13 coat protein
helices (see Monne et al., 1998), this kind of information may provide important constraints for
model building exercises.
What, ®nally, does the striking correlation
between the MGD-values and the length of the
hydrophobic transmembrane segment mean in
structural terms? As is clear from Figure 2(a) and
from the equivalent deletion/insertion experiment
on the H-subunit helix, the relation MGD 0.25n
(where n is the length of the hydrophobic segment)
holds over an extended range of n-values irrespective of whether MGDN or MGDC is measured, and
irrespective of whether the protein is inserted into
the membrane via the SRP/Sec61 pathway or not.
A possible explanation is based on the model
shown in Figure 6, which assumes that a helix
comprised of only 13 hydrophobic residues is
Ê wide hydrophobic core
enough to span the 20 A
of the lipid bilayer (cf. Figure 3(a)), whereas helices
composed of up to 25 hydrophobic residues can
still be accomodated within a reasonably hydroÊ thickness
phobic bilayer environment of 35 ± 40 A
(cf. (Wallin et al., 1997). As illustrated in the Figure,
this model predicts a relation of the form MGD
0.75/3.3n ˆ 0.23n, which is reasonably close to
the observed one. For very long helices that would
extend into the aqueous phase if perpendicular to
the bilayer, tilting would be a way to bury more of
the hydrophobic surface area. If this happens,
MGD would increase much more slowly, if at all,
with n, as seen for the H2 transmembrane segment
in Figure 2(a). However, for the H1 and H3 transmembrane segments, the linear decrease in the
MGDN-values seems to continue well beyond
n ˆ 23; at present, we have no good explanation
for this discrepancy.
Glycosylation Mapping of Transmembrane Helices
Figure 6. Model for the relation between MGD-values
and the length of a hydrophobic transmembrane helix.
For short segments, it is assumed that a perpendicular
orientation has the lowest free energy. An increase in
length by one hydrophobic residue extends the helix by
Ê ; if the locations of the two helix ends are symme1.5 A
Ê closer to the
trically affected, each end moves 0.75 A
membrane surface. To reduce the MGD-value by one
residue, the linker segment (assumed to be in a ¯exible,
extended conformation) between the transmembrane
Ê,
helix and the OST active site needs to move 3 ± 3.5 A
i.e. MGD 0.75/3.3n 0.23n. Y, glycosylated acceptor
site;
non-glycosylated acceptor site.
In summary, the glycosylation mapping technique has allowed us to make a detailed and comprehensive analysis of the helix-breaking effects of
proline residues in transmembrane helices of different lengths (a similar analysis of the effects of
charged residues is presented in the accompanying
paper). The approach also opens up a possibility to
deduce the position in the membrane of transmembrane segments of arbitrary amino acid sequence
relative to ``standards'' such as the reaction center
H-subunit TMH. Since glycosylation is in most
cases a co-translational event, the technique cannot
be applied to fully folded, multi-spanning membrane proteins; however, the position in the membrane of individual transmembrane helices is in
most cases unlikely to change dramatically upon
folding (Popot & Engelman, 1990).
Materials and Methods
Enzymes and chemicals
Unless otherwise stated, all enzymes were from Promega. T7 DNA polymerase, [35S]Met, ribonucleotides,
deoxyribonucleotides, dideoxyribonucleotides, and the
cap analog m7G(50 )ppp(50 )G were from Amersham-Pharmacia (Uppsala, Sweden). Plasmid pGEM1, DTT, BSA,
Sp6 RNA polymerase, RNasin and rabbit reticulocyte
lysate were from Promega. Spermidine was from Sigma.
Oligonucleotides were from Kebo Lab (Stockholm,
Sweden).
1173
Glycosylation Mapping of Transmembrane Helices
DNA manipulations
For cloning into and expression from the pGEM1
plasmid, the 50 end of the lep gene was modi®ed, ®rst,
by the introduction of an XbaI site and, second, by
changing the context 50 to the initiator ATG codon to a
``Kozak consensus'' sequence (Kozak, 1989). Thus, the
50 region of the gene was modi®ed to: . . . ATAACCCTCTAGAGCCACCATGGCGAAT . . . (XbaI site and
initiator codon underlined).
Replacement of the H2 segment in Lep was performed
by ®rst introducing BclI and NdeI restriction sites in
codons 59 and 80 ¯anking the H2 region, and then replacing the BclI-NdeI fragment with the appropriate doublestranded oligonucleotides. Residues 59-81 in H2 were
replaced by residues Val4-Glu38 from the R. sphaeroides
reaction center H-subunit or by poly-Leu sequences of
the general design PGLIKKKKLnVQQQP, where the
valine residue at the end of the poly-Leu stretch was
included to obtain a SpeI restriction site in this position.
The H1 segment was replaced by ®rst introducing BclI
and NdeI restriction sites in codons 4 and 22, and then
replacing the BclI-NdeI fragment with the appropriate
double-stranded oligonucleotides encoding either
residues 4-38 from the reaction center H-subunit
or poly-Leu segments of the general design
LISQQQLnVKKKKHM. The H3 segment was replaced
by ®rst introducing BclI and NdeI restriction sites in
codons 213 and 221, and then replacing the BclI-NdeI
fragment by the appropriate double-stranded oligonucleotides encoding poly-Leu segments of the general
design LISQQQLnVKKKKHM.
Site-speci®c mutagenesis used to add BclI and NdeI
restriction sites at the 30 and 50 ends of H2 in Lep and to
introduce Asn-Thr-Ser acceptor sites for N-linked glycosylation was performed according to the method of
Kunkel (Geisselsoder et al., 1987; Kunkel, 1985). Glycosylation acceptor sites were designed as described (Nilsson
& von Heijne, 1993; Nilsson et al., 1994), i.e. by replacing
three appropriately positioned codons upstream of H1
and H3 or downstream of H2 with codons for the acceptor tri-peptide Asn-Ser-Thr. In the glycosylation construct
Asn84-Ser85-Thr86
and Asn88-Ser89-Thr90
(numbering corresponding to the Lep wild-type
sequence), the ¯anking Pro residues were changed to
Gln since proline residues were found to reduce the ef®ciency of glycosylation, cf. (Gavel & von Heijne (1990).
To make the constructs with d ˆ 5, 6, and 7 (counting
from Glu34) for the Tyr30 ! Pro mutant in the H-subunit transmembrane helix, Ser38, Arg37, and Met36 in
the H-subunit segment (see Figure 5(b)) were progressively deleted from the d ˆ 8 construct. To make glycosylation sites six and ®ve residues downstream of Val1, the
SpeI-NdeI fragment was replaced by double-stranded oligonucleotides, deleting one or two glutamine residues.
All mutants were con®rmed by DNA sequencing of plasmid or single-stranded M13 DNA using T7 DNA polymerase.
The synaptobrevin series of constructs were made as
detailed by (Whitley et al., 1996a). The natural transmembrane segment M95MIILGVICAIILIIIIAYV was replaced
with M95MIKKKKLnVQQQPYV.
For expression and protease-protection experiments in
E. coli, the TM2 segment in construct 2 K/TM1/0 K/
TM2 (Whitley et al., 1994) between the NheI and KpnI
sites was replaced by double-stranded oligonucleotides
encoding, respectively, the sequences N3L7ML15K4,
N3L4PL2ML15K4, and N3L7MLPL13K4.
Expression in vitro
Synthesis of mRNA from pGEM1 by SP6 RNA polymerase and translation in reticulocyte lysate in the presence of dog pancreas microsomes was performed as
described (LiljestroÈm & Garoff, 1991). For the synaptobrevin constructs, the microsomes were subjected to a
sodium carbonate wash procedure (Sakaguchi et al.,
1987) before being loaded onto the gel. Proteins were
analyzed by SDS-PAGE and gels were quanti®ed on a
Fuji BAS1000 phosphoimager using the MacBAS 2.1 software. The extent of glycosylation of a given mutant was
calculated as the quotient between the intensity of the
glycosylated band divided by the summed intensities of
the glycosylated and non-glycosylated bands. In general,
the glycosylation ef®ciency varies by no more than 5%
between different experiments, and the precision in the
MGD determinations is 0.2 residues.
Expression in E. coli
Experiments were performed in E. coli strain MC1061
(lacX74, araD139, (ara, leu)7697, galU, galK, hsr, hsm,
strA; Dalbey & Wickner, 1986). Constructs were
expressed from the pING1 plasmid (Johnston et al., 1985)
by induction with arabinose.
MC1061 cells transformed with the pING1 vector carrying the relevant constructs under control of the arabinose promoter were grown at 37 C in M9 minimal
medium supplemented with 100 mg/ml ampicillin, 0.5%
(w/v) fructose, 100 mg/ml thiamine, and all amino acids
(50 mg/ml each) except methionine. An overnight culture
was diluted 1:25 in fresh medium, shaken for 3.5 hours
at 37 C, induced with arabinose (0.2% (w/v)) for ®ve
minutes, labeled with [35S]methionine (75 mCi/ml) for
one minute, and put on ice. Cells were spun at
14,000 rpm for two minutes, resuspended in ice-cold buffer (40% (w/v) sucrose, 33 mM Tris, pH 8.0), and incubated with lysozyme (5 mg/ml) and 1 mM EDTA for
15 minutes on ice. Aliquots of the cell suspension were
incubated 60 minutes on ice, either with no additions, or
with the addition of 400 mg/ml proteinase K. After
addition of PhMeSO2F (PMSF), samples were acid-precipitated with trichloroacetic acid (TCA, 10% (v/v) ®nal
concn), resuspended in 10 mM Tris/2% SDS, immunoprecipitated with antisera to Lep, washed, and analyzed
by SDS-PAGE. Gels were scanned in a FUJIX Bas 1000
phosphoimager.
Acknowledgments
This work was supported by grants from the Swedish
Cancer Foundation, the Swedish Natural and Technical
Sciences Research Councils, and the GoÈran Gustafsson
Foundation to G.v.H.
References
Allen, J., Feher, G., Yeates, T., Komiya, H. & Rees, D.
(1987). Structure of the reaction center from Rhodobacter sphaeroides R-26: the protein subunits. Proc.
Natl Acad. Sci. USA, 84, 6162± 6166.
Barlow, D. J. & Thornton, J. M. (1988). Helix geometry
in proteins. J. Mol. Biol. 201, 601± 619.
Dalbey, R. E. & Wickner, W. (1986). The role of the
polar, carboxyl-terminal domain of Escherichia coli
1174
leader peptidase in its translocation across the
plasma membrane. J. Biol. Chem. 261, 13844± 13849.
Deber, C. M. & Goto, N. K. (1996). Folding proteins into
membranes. Nature Struct. Biol. 3, 815± 818.
Gavel, Y. & von Heijne, G. (1990). Sequence differences
between glycosylated and non-glycosylated Asn-XThr/Ser acceptor sites ± implications for protein
engineering. Protein Eng. 3, 433±442.
Geisselsoder, J., Witney, F. & Yuckenberg, P. (1987). Ef®cient site-directed in vitro mutagenesis. BioTechniques, 5, 786± 791.
GoÈrlich, D. & Rapoport, T. A. (1993). Protein translocation into proteoliposomes reconstituted from puri®ed components of the endoplasmic reticulum
membrane. Cell, 75, 615± 630.
Grigorieff, N., Ceska, T. A., Downing, K. H., Baldwin,
J. M. & Henderson, R. (1996). Electron-crystallographic re®nement of the structure of bacteriorhodopsin. J. Mol. Biol. 259, 393± 421.
Harper, E. T. & Rose, G. D. (1993). Helix stop signals in
proteins and peptides: the capping box. Biochemistry, 32, 7605± 7609.
Iwata, S., Ostermeier, C., Ludwig, B. & Michel, H.
Ê resolution of cytochrome c
(1995). Structure at 2.8 A
oxidase from Paracoccus denitri®cans. Nature, 376,
660± 669.
Johansson, M., Nilsson, I. & von Heijne, G. (1993). Positively charged amino acids placed next to a signal
sequence block protein translocation more ef®ciently in Escherichia coli than in mammalian microsomes. Mol. Gen. Genet. 239, 251± 256.
Johnston, S., Lee, J. H. & Ray, D. S. (1985). High-level
expression of M13 gene II protein from an inducible
polycistronic messenger RNA. Gene, 34, 137± 145.
Kozak, M. (1989). Context effects and inef®cient
initiation at non-AUG codons in eucaryotic cell-free
translation systems. Mol. Cell. Biol. 9, 5073± 5080.
Kraulis, P. J. (1991). MOLSCRIPT: A program to produce both detailed and schematic plots of protein
structures. J. Appl. Crystallog. 24, 946± 950.
Kunkel, T. A. (1985). Rapid and ef®cient site-speci®c
mutagenesis without phenotypic selection. Proc.
Natl Acad. Sci. USA, 82, 488± 492.
Kutay,
U.,
Ahnerthilger,
G.,
Hartmann,
E.,
Wiedenmann, B. & Rapoport, T. A. (1995). Transport route for synaptobrevin via a novel pathway
of insertion into the endoplasmic reticulum membrane. EMBO J. 14, 217±223.
Li, S.-C. & Deber, C. M. (1994). A measure of helical
propensity for amino acids in membrane environments. Nature Struct. Biol. 1, 368± 373.
Li, S.-C., Goto, N. K., Williams, K. A. & Deber, C. M.
(1996). a-Helical but not b-sheet, propensity of proline is determined by peptide environment. Proc.
Natl Acad. Sci. USA, 93, 6676± 6681.
LiljestroÈm, P. & Garoff, H. (1991). Internally located
cleavable signal sequences direct the formation of
Semliki Forest virus membrane proteins from a
polyprotein precursor. J. Virol. 65, 147± 154.
Liu, L. P., Li, S. C., Goto, N. K. & Deber, C. M. (1996).
Threshold hydrophobicity dictates helical conformations of peptides in membrane environments.
Biopolymers, 39, 465± 470.
Martoglio, B., Hofmann, M. W., Brunner, J. &
Dobberstein, B. (1995). The protein-conducting
channel in the membrane of the endoplasmic reticulum is open laterally toward the lipid bilayer. Cell,
81, 207± 214.
Glycosylation Mapping of Transmembrane Helices
Mellquist, J. L., Kasturi, L., Spitalnik, S. L. &
ShakinEshleman, S. H. (1998). The amino acid following an Asn-X-Ser/Thr sequon is an important
determinant of N-linked core glycosylation ef®ciency. Biochemistry, 37, 6833± 6837.
MonneÂ, M., Nilsson, I., Elmhed, N., Johansson, M. &
von Heijne, G. (1998). Positively and negatively
charged residues have different effects on th position in the membrane of a model transmembrane
helix. J. Mol. Biol..
Mothes, W., Heinrich, S., Graf, R., Nilsson, I., von
Heijne, G., Brunner, J. & Rapoport, T. (1997). Molecular mechanisms of membrane protein integration
into the endoplasmic reticulum. Cell, 89, 523± 533.
Nilsson, I. & von Heijne, G. (1993). Determination of the
distance between the oligosaccharyltransferase
active site and the endoplasmic reticulum membrane. J. Biol. Chem. 268, 5798± 5801.
Nilsson, I., Whitley, P. & von Heijne, G. (1994). The
C-terminal ends of internal signal and signal-anchor
sequences are positioned differently in the ER translocase. J. Cell. Biol. 126, 1127± 1132.
Papavoine, C., Remerowski, M., Horstink, L., Konings,
R., Hilbers, C. & van de Ven, F. (1997). Backbone
dynamics of the major coat protein of bacteriophage
M13 in detergent micelles by 15N nucelar magnetic
resonance relaxation measurements using the
model-free approach and reduced spectral density
mapping. Biochemistry, 36, 4015± 4026.
Popot, J. L. & Engelman, D. M. (1990). Membrane
protein folding and oligomerization ± the 2-stage
model. Biochemistry, 29, 4031± 4037.
Popov, M., Tam, L. Y., Li, J. & Reithmeier, R. A. F.
(1997). Mapping the ends of transmembrane segments in a polytopic membrane protein ± scanning
N-glycosylation mutagenesis of extracytosolic loops
in the anion exchanger, Band 3. J. Biol. Chem. 272,
18325± 18332.
Rapoport, T. A., Jungnickel, B. & Kutay, U. (1996). Protein transport across the eukaryotic endoplasmic
reticulum and bacterial inner membranes. Annu.
Rev. Biochem. 65, 271±303.
Rees, D. C., Komiya, H., Yeates, T. O., Allen, J. P. &
Feher, G. (1989). The bacterial photosynthetic reaction center as a model for membrane proteins.
Annu. Rev. Biochem. 58, 607± 633.
Richardson, J. S. & Richardson, D. C. (1988). Amino acid
preferences for speci®c locations at the ends of
a-helices. Science, 240, 1648± 1652.
Sakaguchi, M., Mihara, K. & Sato, R. (1987). A short
amino-terminal segment of microsomal cytochrome
P-450 functions both as an insertion signal and as a
stop-transfer sequence. EMBO J. 6, 2425±2431.
Sakaguchi, M., Tomiyoshi, R., Kuroiwa, T., Mihara, K. &
Omura, T. (1992). Functions of signal and signalanchor sequences are determined by the balance
between the hydrophobic segment and the N-terminal charge. Proc. Natl Acad. Sci. USA, 89, 16 ±19.
Sansom, M. S. P. (1992). Proline residues in transmembrane helices of channel and transport proteins: a
molecular modelling study. Protein Eng. 5, 53 ± 60.
Shakineshleman, S. H., Spitalnik, S. L. & Kasturi, L.
(1996). The amino acid at the X position of an AsnX-Ser sequon is an important determinant of
N-linked core-glycosylation ef®ciency. J. Biol. Chem.
271, 6363± 6366.
Shen, L., Bassolino, D. & Stouch, T. (1997). Transmembrane helix structure, dynamics, and interactions:
Glycosylation Mapping of Transmembrane Helices
multi-nanosecond molecular dynamics simulations.
Biophys. J. 73, 3 ± 20.
Shon, K. J., Kim, Y. G., Colnago, L. A. & Opella, S. J.
(1991). NMR studies of the structure and dynamics
of membrane-bound bacteriophage-P¯ coat protein.
Science, 252, 1303± 1304.
Stopar, D., Spruijt, R. B., Wolfs, C. & Hemminga, M. A.
(1996). Local dynamics of the M13 major coat protein in different membrane-mimicking systems. Biochemistry, 35, 15467± 15473.
Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T.,
Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, R.,
Yaono, R. & Yoshikawa, S. (1996). The whole stucture of the 13-subunit oxidized cytochrome c oxiÊ . Science, 272, 1136± 1144.
dase at 2.8 A
von Heijne, G. (1991). Proline kinks in transmembrane
a-helices. J. Mol. Biol. 218, 499± 503.
von Heijne, G. (1996). Principles of membrane protein
assembly and structure. Prog. Biophys. Mol. Biol. 66,
113± 139.
1175
Wallin, E., Tsukihara, T., Yoshikawa, S., von Heijne, G.
& Elofsson, A. (1997). Architecture of helix bundle
membrane proteins: an analysis of cytochrome c
oxidase from bovine mitochondria. Protein Sci. 6,
808± 815.
Whitley, P., Nilsson, I. & von Heijne, G. (1994). De novo
design of integral membrane proteins. Nature Struct.
Biol. 1, 858±862.
Whitley, P., Grahn, E., Kutay, U., Rapoport, T. & von
Heijne, G. (1996a). A 12 residues long poly-leucine
tail is suf®cient to anchor syntaptobrevin to the ER
membrane. J. Biol. Chem. 271, 7583± 7586.
Whitley, P., Nilsson, I. M. & von Heijne, G. (1996b).
A nascent secretory protein may traverse the ribosome/ER translocase complex as an extended
chain. J. Biol. Chem. 271, 6241± 6244.
Wolfe, P. B., Wickner, W. & Goodman, J. M. (1983).
Sequence of the leader peptidase gene of Escherichia
coli and the orientation of leader peptidase in the
bacterial envelope. J. Biol. Chem. 258, 12073± 12080.
Edited by F. Cohen
(Received 15 May 1998; received in revised form 1 September 1998; accepted 1 September 1998)