You are on page 1of 13

Leading Edge

Review

Membrane Budding
_
James H. Hurley,1,* Evzen Boura,1 Lars-Anders Carlson,1 and Bartosz Rózycki 2
1Laboratory of Molecular Biology
2Laboratory of Chemical Physics
National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, MD 20892-0580, USA
*Correspondence: hurley@helix.nih.gov
DOI 10.1016/j.cell.2010.11.030

Membrane budding is a key step in vesicular transport, multivesicular body biogenesis, and envel-
oped virus release. These events range from those that are primarily protein driven, such as the
formation of coated vesicles, to those that are primarily lipid driven, such as microdomain-
dependent biogenesis of multivesicular bodies. Other types of budding reside in the middle of
this spectrum, including caveolae biogenesis, HIV-1 budding, and ESCRT-catalyzed multivesicular
body formation. Some of these latter events involve budding away from cytosol, and this unusual
topology involves unique mechanisms. This Review discusses progress toward understanding
the structural and energetic bases of these different membrane-budding paradigms.

Eukaryotic cells are defined by their compartmentalization into thermal energy (Bloom et al., 1991).This is important for biology
membrane-delimited structures. The protein and lipid content because events that require thermal energy of this magnitude
of these membranes is maintained and regulated by a constant (that is, of 100 kBT or greater) do not occur spontaneously.
flux of vesicular trafficking. Each vesicular trafficking event Biophysical studies of membrane budding, which offer the
involves the budding of a membrane vesicle from a donor promise of accounting for energetics, are typically carried out
membrane, typically followed by its regulated transport to, dock- in vesicles that are much larger than their counterparts in biolog-
ing to, and fusion with an acceptor membrane. Many viruses also ical systems. Fortunately, the energetic cost of bud formation is
have membrane envelopes and escape from host cells by to a first approximation independent of the size of the bud.
membrane-budding events. In pure lipid mixtures used in biophysical studies, vesicles are
Our laboratory has been characterizing the unusual microns in size, spreading the energetic cost over 106 or
membrane-budding reaction promoted by the ESCRTs, which more lipid molecules. In cells, however, membrane buds have
has led us to take a fresh look at how membrane lipid properties a diameter of 20–100 nm, thus involving as few as 103–104 lipid
might make protein-dependent, energetically expensive reac- molecules. This poses the question, how do a modest number of
tions easier. Several excellent reviews have covered the way protein-lipid interactions create the free energy that is needed for
proteins induce curvature in biological membranes (Farsad and budding, or alternatively, how do lipids themselves contribute to
De Camilli, 2003; McMahon and Gallop, 2005; Voeltz and Prinz, lowering the energy barrier?
2007) and the physical principles of membrane curvature
(Zimmerberg and Kozlov, 2006). This Review will take a different Coated Vesicle Budding
viewpoint and consider the comparative roles of proteins and Clathrin
lipids in select examples of vesicular budding events (Figure 1) The dominant mechanism of membrane budding into the cytosol
to discuss similarities and differences in budding events in and the paradigm for protein-directed budding is the formation
synthetic versus cellular contexts, the potential roles of proteins of coated vesicles (Figures 1F and 1G and Figure 2). Clathrin-
in orchestrating lipid phase changes, and the roles of lipids in coated vesicles (CCVs) are typically 60–100 nm in diameter
recruiting and regulating proteins. We also examine the implica- (Bonifacino and Lippincott-Schwartz, 2003; Brodsky et al.,
tions of the above for cell physiology. This article is not intended 2001). Clathrin can form baskets in vitro that resemble the
as a comprehensive review of all cellular budding events. Rather, CCVs in the absence of membranes, and the basket structure
we consider emerging mechanistic thinking in multivesicular has been characterized in molecular detail (Fotin et al., 2004).
body formation and virus budding, placing these in the context Clathrin itself binds neither membranes nor cargo but relies on
of the classical mechanisms underlying budding of coated adaptors for this function. Among the most comprehensively
vesicles. studied is adaptor protein complex 2 (AP-2 complex) (Robinson
and Bonifacino, 2001), which functions in clathrin-mediated
Energetics of Vesicle Budding endocytosis at the plasma membrane. The AP-2 adaptor
The formation of spherical vesicles from a flat membrane of complex opens up in the presence of cargo and the lipid phos-
typical biological composition and no intrinsic propensity to phatidylinositol (4,5)-bisphosphate (PI(4,5)P2) to form a flat plat-
curve entails a membrane-bending free energy (Helfrich, 1973), form capable of binding multiple PI(4,5)P2 and cargo molecules
DG = 8pk 250–600 kBT, given k 10–25 kBT, where kBT is (Jackson et al., 2010). The established role for PI(4,5)P2 in this

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 875


Figure 1. Proteins and Lipid Microdomains in Membrane Budding
(A) Budding of phase-separated lipid microdomains from GUVs (giant unilamellar vesicles) composed of synthetic lipids is an example of membrane budding in
the absence of any proteins. Reproduced by permission from Baumgart et al. (2003).
(B) Shiga toxin (black dots) acts from outside the plasma membrane to induce membrane buds and is an example of a protein triggering budding events that are
primarily driven by lipid microdomains. Image reproduced by permission from Macmillan Publishers Ltd: Nature, Römer et al. (2007), copyright 2007.
(C) Budding by caveolae represents a hybrid between a membrane microdomain and protein coat-driven mechanisms. Reproduced by permission from Mac-
millan Publishers Ltd: Nat. Rev. Mol. Cell. Biol., Parton and Simons (2007), copyright 2007.
(D) ESCRT-I and -II induce buds in synthetic GUVs. Reproduced by permission from Wollert and Hurley (2010). Proteins organize these structures but do not form
a coat, suggesting a possible role for microdomains.
(E) HIV-1 buds visualized by electron tomography (Carlson et al., 2008). The bud is organized by the HIV-1 capsid protein, heavily enriched in raft lipids, and
cleaved by ESCRT proteins.
(F) Deep etch visualization of clathrin-coated pits (image courtesy of J. Heuser). Clathrin assembles into baskets in the absence of membranes but is thought to be
too flexible to deform membranes on its own. For this, clathrin needs help from other membrane-deforming proteins and possibly from lipids.
(G) The COP II cage is an example of a protein structure that can form in the absence of lipids and can impose its shape on any simple bilayer-forming lipid mixture.
Reproduced by permission from Russell and Stagg (2010).

pathway is to recruit AP-2 and other proteins to the site of domain proteins, and PI(4,5)P2 constitute the minimum require-
budding. A role for PI(4,5)P2 clustering into microdomains has ments for membrane bud formation in this pathway.
been suggested on theoretical grounds (Liu et al., 2006) but The scission of the clathrin-coated bud to form a detached
has yet to be directly visualized. vesicle is a complex process in its own right, and the reader is
Clathrin is absolutely required for the budding of AP-2- and referred to recent reviews (Pucadyil and Schmid, 2009). Finally,
cargo-rich plasma membrane domains, which remain flat in its following scission, the clathrin coat is removed by the ATP-
absence (Hinrichsen et al., 2006). However, clathrin monomers dependent action of the molecular chaperone Hsc70 and its
are flexible, which gives clathrin the ability to form different types cofactor auxillin (Eisenberg and Greene, 2007). It is only following
of lattices and to adapt to various cargoes (Ehrlich et al., 2004). nucleotide hydrolysis that the energetic cost of clathrin-induced
Given the flexibility of clathrin monomers, the energy of clathrin membrane deformation is finally paid, making the full reaction
polymerization has been proposed on theoretical grounds to cycle—from flat membrane to uncoated vesicle—thermody-
be insufficient on its own to bend the membrane into a bud (Nos- namically irreversible.
sal, 2001). However, this concept has yet to be confirmed exper- COP I and COP II
imentally and is not universally accepted. Vesicles carrying cargo from the endoplasmic reticulum (ER) to
Cholesterol is important for clathrin-mediated endocytosis by the Golgi are coated by the COP II complex, which, like clathrin,
many (though not all) accounts (Rodal et al., 1999; Subtil et al., can form membrane-free baskets in vitro with vesicle-like dimen-
1999), although it is less sensitive to cholesterol depletion than sions (Stagg et al., 2006). COP II vesicles have a preferred size,
most coat-independent budding pathways (Sandvig et al., but as with clathrin, the flexibility of the COP II subunits allows
2008). Clathrin, cargo adaptors, and PI(4,5)P2 are necessary formation of expanded lattices that can accommodate large
but not sufficient on their own to induce membrane curvature. cargoes such as procollagen and large lipoprotein particles
The essential early endocytic factor epsin wedges its amphi- known as chylomicrons (Stagg et al., 2008).
pathic helix a0 into the membrane upon PI(4,5)P2 binding, COP II vesicle budding has been reconstituted in vitro from
promoting positive curvature (Ford et al., 2002). The cargo- purified proteins and synthetic lipids (Lee et al., 2005; Matsuoka
binding muniscin proteins FCHo1/2 (Syp1 in yeast) contain et al., 1998). A membrane consisting only of synthetic unsatu-
BAR domains that promote positive curvature very early in endo- rated phospholipids was capable of supporting budding
cytosis (Henne et al., 2010; Reider et al., 2009; Stimpson et al., (Matsuoka et al., 1998). COP II consists of the Sec23/24 sub-
2009; Traub and Wendland, 2010). In principle, the reagents complex, which binds lipids and cargo via a gently curved face
and concepts would appear to be in place to reconstitute (Bi et al., 2002), the Sec13/31 subcomplex, which forms an outer
clathrin-dependent membrane budding. Reconstitution of cage around the vesicle, and the membrane-bending GTPase
clathrin-mediated endocytosis using synthetic lipids and purified Sar1. The Sec23/24 and Sec13/31 subcomplexes in combina-
proteins would be an important step in determining whether tion are sufficient to form buds, with Sar1 strictly required only
clathrin, AP-2, one or more amphipathic helix and/or BAR for the scission of the buds. GTP hydrolysis by Sar1 provides

876 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


Figure 2. Coated Vesicle Budding
(A) Structure of a clathrin basket from cytoelectron microscopy; reproduced by
permission from Macmillan Publishers Ltd: Nature, Fotin et al. (2004), copy-
right 2004.
(B) COP II vesicles produced from purified components; reproduced by
permission from Lee et al. (2005).
(C) Structural parallels between clathrin, COP I, and COP II. Adapted from Lee
and Goldberg (2010).

energy input into the system, making the overall process (which
Figure 3. Membrane Microdomains and Budding
culminates in the uncoating of cargo-loaded vesicles) thermody-
(A) Coexistence of phases in model membranes visualized by atomic force
namically irreversible. microscopy in a supported bilayer (a membrane bilayer adsorbed onto a solid
COP I-coated vesicles are responsible for retrograde traffic support, usually glass). Reproduced with permission from Chiantia et al.
from the Golgi to the ER, and this reaction has also been recon- (2006).
(B) Phase transitions in a single-lipid membrane analyzed by molecular
stituted from purified proteins and synthetic lipids. The budding dynamics simulations. Reproduced with permission from Heller et al. (1993).
reaction requires the coatomer complex, GTP-bound Arf1, and Copyright 1993 American Chemical Society.
protein cargo tails tethered to the membrane but has no special (C) Schematic model of a raft-type membrane microdomain, including a model
of a myristoylated ESCRT-III subunit Vps20 as an example of protein that
lipid requirements (Bremser et al., 1999). Budding occurs even might anchor to rafts.
from vesicles composed of the pure synthetic phospholipid
DOPC doped with small amounts of a lipopeptide cargo.
Recently, a composite crystallographic structure of cage-form- phase, with the translational and conformational order of the lipid
ing components of coatomer consisting of the a, b0 , and 3 chains depending on their composition and the temperature. The
subunits has been determined and shown to resemble the cla- liquid phase is the more relevant to biology and can be subdi-
thrin triskelion (Lee and Goldberg, 2010). In sum, COP I and vided into liquid disordered (Ld) and liquid ordered (Lo) phases.
COP II provide some of the purest examples of protein-directed Lipids in the Ld phase have higher conformational freedom and
membrane budding, in which the protein coat imposes its shape diffusion coefficients than in the Lo phase. At biological temper-
upon the membrane with minimal dependence on its lipid atures, the Ld and Lo phases can coexist in membranes of mixed
composition. composition (Elson et al., 2010; Garcı́a-Sáez and Schwille,
2010).
Membrane Microdomains and Budding In general, phospholipids with unsaturated chains prefer the
Lipid Phase Separation as a Budding Mechanism Ld phase, whereas cholesterol, sphingolipids, and phospholipids
In contrast to the protein-dominated paradigm of coated vesicle with saturated chains prefer the Lo phase (Lingwood and
budding, phase separation in simple lipid mixtures can drive Simons, 2010). Typically, the energetic cost for contact between
budding on a micron scale in synthetic model membranes, in dissimilar lipids is small, 0.5 kBT (Garcı́a-Sáez and Schwille,
the absence of proteins (Baumgart et al., 2003) (Figure 1A and 2010), but becomes significant when summed over many lipids.
Figure 3). Membrane bilayers can adopt either a solid or a liquid The higher acyl chain order in the Lo phase results in their

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 877


Figure 4. Protein Structures that Cluster Raft Lipids
(A) Simian virus 40 VP1 pentamer bound to the membrane via the headgroup of the ganglioside GM1 (Neu et al., 2008).
(B) Cholera toxin B subunit pentamer bound to GM1 (Merritt et al., 1994).
(C) Composite model of the myristoylated HIV-1 matrix domain trimer bound to PI(4,5)P2 (Hill et al., 1996; Saad et al., 2006, 2008). In each case, lipid tails are
modeled. Images were generated with VMD 1.8.6.

elongation to their maximum extent, hence Lo membrane domains known as ‘‘rafts.’’ Why don’t rafts and other microdo-
domains are thicker than Ld domains. The height mismatch at mains coalesce on the micron scale in living cells, as they do
the phase boundary is energetically unfavorable because it in model membranes? The answer is not known, but the action
forces the polar headgroup region of the Ld domain into contact of the cytoskeleton and membrane traffic, and the large fraction
with the hydrophobic portion of the Lo domain. The free energy of protein in cellular membranes, are usually invoked. Indeed, it
cost per unit length is known as the line tension and has units is to be expected that cells would have mechanisms to
of force. In order to minimize the free energy associated with block the unchecked growth of microdomains, as the ensuing
line tension, membrane domains will coalesce with one another spontaneous vesiculation of cell membranes would be disas-
into circular zones. When circular domains reach a critical size at trous.
which the line tension energy term exceeds the Helfrich (curva- Soluble and lumenally anchored cargoes, viruses, and toxins
ture-dependent) energy of membrane deformation, the are selectively transported in vesicular carriers even though
membrane will deform out of plane in order to minimize the they have no direct communication with the cytosol to signal
zone of contact (Lipowsky, 1992). If the line tension is high their packaging and sorting. In some cases, transmembrane-
enough, the neck connecting the membrane bud can be sorting receptors serve as adaptors to link cargo to conventional
severed, leading to the formation of detached vesicles. In addi- cytosolic coat complexes. In other cases, membrane rafts make
tion to line tension effects, membrane microdomain formation the link. Simian virus 40 (SV40) and cholera toxin enter cells by
can bend membranes by concentrating lipids with distinct binding to multiple molecules of the ganglioside GM1 (Damm
intrinsic curvatures, and the contents of such microdomains et al., 2005; Kirkham et al., 2005), a raft-favoring lipid. The
can not only drive budding but dictate its direction (Bacia cholera toxin B subunit (Merritt et al., 1994) and the SV40 VP1
et al., 2005). protein (Neu et al., 2008) both bind to GM1 as pentamers.
The complex lipid mixture of the plasma membrane supports Cholera toxin pentamer binds GM1 (Figure 4) and thus induces
phase separation in micron-sized domains when reconstituted formation of an Lo microdomain in model membranes
in giant unilamellar vesicles (Baumgart et al., 2007). However, (Hammond et al., 2005) and in turn leads to budding (Bacia
in living cells, membrane microdomains are heterogeneous, et al., 2005; Ewers et al., 2010). Shiga toxin B subunit binds
highly dynamic nanoscale structures (Hancock, 2006; Lingwood the glycolipid Gb3 and appears to operate by a similar paradigm.
and Simons, 2010; Pike, 2006). In the most up-to-date biophys- In this case tubular vesicles are formed, and lipid compression
ical view, these nanoscale structures likely correspond to critical favoring negative curvature is thought to be the driving force
fluctuations (Veatch et al., 2007). Although the concepts of the (Römer et al., 2007). In each of these examples, it is clear that
Lo and Ld phases are oversimplifications of the variety of clustering of lipids leads to important changes in membrane
dynamic membrane substructures that exist in cells (Lingwood structure that contribute to budding. The proposed physical
and Simons, 2010), they will be used in this Review because mechanisms remain speculative, however. Revealing these
they are useful intuitive handles, deeply ingrained in the mechanisms remains a profound challenge to experimen-
literature, and helpful in relating model membrane studies to talists and thus is an area that will benefit from increasing
biology. Most, but not all, of the membrane microdomains impli- sophisticated computer simulations of membrane dynamics on
cated in cellular budding are the sterol- and sphingolipid-rich realistic timescales.

878 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


Caveolae brane-spanning a helices (Hemler, 2005). Tetraspanins have two
Caveolae (‘‘little caves’’) are flask-shaped 60–80 nm invagina- extracellular domains; the second, EC2, is the larger of the two.
tions of the plasma membrane that consist of raft lipids, caveo- The structure of the EC2 region of CD81 has been determined,
lins 1–3, and the caveolin-associated cavins 1–4 (Hansen and revealing an extensive dimerization interface (Kitadokoro et al.,
Nichols, 2010). Caveolins are structurally analogous to the retic- 2001). The minimal functional tetraspanin oligomer is probably
ulons and DP1/Yop1 proteins that maintain the curvature of ER a homodimer. These proteins are multiply palmitoylated on their
membrane tubules (Hu et al., 2008; Shibata et al., 2009) and to short intracellular loop and N- and C-terminal extensions, and
another plasma membrane raft protein, flotillin (Bauer and Pelk- these palmitoylations are central to their ability to form TEMs.
mans, 2006). Caveolins are pentahelical proteins, with two of the Tetraspanins bind to a wide range of potential cargo proteins
helices inserting deeply into the membrane, almost but not (Hemler, 2005), potentially coupling them to TEMs and thereby
completely spanning the bilayer. The other three helices are to microdomain-mediated budding. More extensive mechanistic
amphipathic and are thought to wedge themselves into the inter- analysis of the budding mechanism responsible for TEM traffic
facial region of the membrane (Parton et al., 2006). Caveolae will be eagerly awaited.
contain a consistent number of caveolin molecules, 144, which
suggests the formation of a highly organized coat (Pelkmans and Multivesicular Bodies
Zerial, 2005). The sorting of unneeded, damaged, or dangerous plasma
Posttranslational modification of caveolins is important to their membrane proteins to the lysosome for degradation is carried
function. Palmitoylation at multiple residues promotes their out by endosomes (Sorkin and von Zastrow, 2009). This pathway
constitutive association with cholesterol and other raft lipids. also is central to the biogenesis of the lysosome (or yeast
Caveolins also undergo phosphoregulation by multiple protein vacuole), as it carries newly synthesized lysosomal enzymes
kinases (Pelkmans and Zerial, 2005). For instance, when caveo- from the trans-Golgi to their destination. In the metazoa, the
lin-1 is phosphorylated at serine 80, which adjoins one of the endosomal pathways have many additional roles, with the
predicted interfacial a helices, its ability to induce curvature is most pertinent to this Review being the biogenesis of lyso-
turned off. Although the energetic book-keeping of caveolin- some-related organelles (Raposo and Marks, 2007) and
induced curvature has not been worked out, it is likely to differ exosomes. Multivesicular bodies (MVBs, also known as multive-
greatly from that of conventional coated vesicles. Insertion of sicular endosomes) are key intermediates in endolysosomal
caveolin into the membrane presumably shifts the intrinsic transport (Figure 5; Gruenberg and Stenmark, 2004; Piper and
curvature of the membrane such that the positively curved bud Katzmann, 2007). MVBs are formed by the invagination and
is the low-energy state and the flat caveolin microdomain is scission of buds from the limiting membrane of the endosome
the high-energy state. Thus, once the caveolin microdomain is into the lumen. MVB biogenesis is the main physiological
formed, energy input is probably needed to flatten the example of membrane budding away from the cytosol.
membrane rather than to curve it. ATP hydrolysis by protein ESCRTs and Multivesicular Bodies
kinases that phosphorylate caveolin might provide the thermo- Yeast (Saccharomyces cerevisiae) has a single MVB pathway
dynamic driving force for membrane flattening. Dephosphoryla- that drives the internalization of ubiquitinated transmembrane
tion by protein phosphatases would, in this speculative scheme, proteins into the lumens of early endosomes (Piper and Katz-
allow the membrane to spring back to its low-energy state. mann, 2007). The pathway is initiated by the presence of the lipid
Cavins are soluble proteins rich in predicted coiled-coil struc- phosphatidylinositol 3-phosphate (PI(3)P) and membrane-teth-
ture and basic residues but otherwise structurally uncharacter- ered ubiquitin moieties on the endosome surface. PI(3)P is
ized. They seem to be important for caveolar structure, but the synthesized by the class III PI 3-kinase Vps34, an enzyme essen-
precise role of these recently discovered factors in structuring tial for the progression of the endolysosomal pathway. PI(3)P is
the caveolar coat is not clear. Given that caveolae have a consis- the defining marker of early endosomes, autophagosomes,
tent amount of caveolin protomers, they could be viewed as highly and, in mammalian cells, phagosomes. PI(3)P signals are recog-
organized assemblies whose specialized structure and distinct nized by FYVE and PX domain-containing proteins (Misra et al.,
curvature are caveolin driven but lipid stabilized. Alternatively, if 2001). In the MVB pathway, the key FYVE domain protein is
viewed from the standpoint of their lipid content, caveolae could a subunit of the ESCRT-0 complex. ESCRT-0 contains five
be viewed as specialized, morphologically distinct membrane ubiquitin-binding domains (UBDs) (Ren and Hurley, 2010) and
microdomains, whose formation is driven by lipids but stabilized clusters ubiquitinated cargo in vitro (Wollert and Hurley, 2010).
by caveolin (Parton and Simons, 2007). The hybrid nature of Recruitment of ESCRT-0 to the early endosomal membrane
caveolae, seemingly at once both coated vesicle and membrane initiates the recruitment of the ESCRT-I, -II, and –III complexes
microdomain, makes them a particularly fascinating example of (Saksena et al., 2007; Williams and Urbé, 2007). Based on
the interplay between proteins and lipids in membrane budding. in vitro reconstitution, ESCRT-I and -II drive membrane budding,
Tetraspanin-Enriched Microdomains whereas ESCRT-III cleaves the bud necks to form intralumenal
Tetraspanin-enriched microdomains (TEMs), which are abun- vesicles (Hurley and Hanson, 2010; Wollert and Hurley, 2010;
dant in exosomes and in the intralumenal vesicles of immune Wollert et al., 2009). In vitro ESCRT budding reactions have
cell multivesicular bodies, are another potential example of been carried out with a mixture of saturated and unsaturated
a membrane microdomain involved in budding (Pols and Klum- phospholipids and cholesterol (Wollert and Hurley, 2010), but
perman, 2009). Tetraspanins are a family of at least 32 proteins the precise lipid requirements for the reaction have yet to be
in mammals and are defined by the presence of four transmem- analyzed in detail.

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 879


Figure 5. Multivesicular Bodies Bud via
Diverse Mechanisms
(A) Multivesicular bodies (MVBs) form from late en-
dosomes in animal cells. Their formation is depen-
dent on both ESCRT complexes and the unusual
lipid lysobisphosphatidic acid (LBPA).
(B) The conserved ESCRT-dependent MVB
biogenesis pathway from early endosomes in
yeast and animal cells. PI(3)P has been directly
visualized in these MVBs. Cholesterol has been
visualized in MVBs from animal cells, but it has
not been directly confirmed whether these are
ESCRT dependent or not.
(C) Specialized formation of MVBs containing
polymerized Pmel17.
(D) Ceramide-dependent MVBs bud from raft-like
and tetraspanin-enriched microdomains in animal
cells.

complexes in solution (Im and Hurley,


2008; Kostelansky et al., 2007). These
structures show that multiple membrane
and ESCRT-III attachment sites are sepa-
rated by rigid spacers of up to 18 nm
across, suggesting a mechanism to
Strikingly, ESCRT-I, -II, and -III all localize to the bud neck induce or at least stabilize formation of a membrane neck of
(Wollert and Hurley, 2010). ESCRT-III subunits assemble into roughly those dimensions. Subsequent recruitment and poly-
tubular structures in vitro and when overexpressed (Bajorek merization of ESCRT-III into spiral domes (Fabrikant et al.,
et al., 2009; Hanson et al., 2008; Lata et al., 2008). The 2009) would then narrow and sever the neck in the current model
ESCRT-III proteins coat the interior of lipid tubes created (Hurley and Hanson, 2010).
in vitro (Lata et al., 2008) and have diameters of 40–50 nm for The observation that the ESCRT complexes localize to the bud
lipid-free tubes, and 100 nm for lipid-coated tubes. These neck explains how they bud membranes away from the cytosol
tubes exceed the narrowest dimensions of bud necks in cells, without themselves being consumed in the bud. This mechanism
based on just a few observations that suggest a size closer to stands in sharp contrast to the familiar budding of coated vesi-
20 nm (Murk et al., 2003). However, the tubes taper to cles toward cytosol, described above. The thermodynamic
a dome at their ends (Fabrikant et al., 2009), which may repre- driving force for the pathway is the coupling of ESCRT-III solubi-
sent the tubes’ most important functional feature. Lipid tube lization and recycling to ATP hydrolysis by the dodecameric AAA
extrusion by ESCRT-III seems to have no special lipid require- ATPase Vps4 (Babst et al., 1998; Wollert et al., 2009). Although
ments, as it can be supported in vitro by a simple mixture of the overall thermodynamic driving force is clear, the energetic
the unsaturated phospholipids SOPC and DOPS (Lata et al., trajectory of neck-directed bud formation is currently unknown.
2008). Indeed, whereas most ESCRTs are unique to the eukarya, Theoretical analysis of the membrane mechanics of this process
ESCRT-III is conserved in a subset of Archaea, where it functions is urgently needed, as is a better understanding of the roles of
in the membrane abscission step of cell division (Lindås et al., lipids.
2008; Samson et al., 2008). Thus the Archaeal ESCRT-III ortho- All four ESCRT complexes are conserved between yeast and
logs can presumably function in membrane scission with metazoa. In its broad outlines, the ESCRT-dependent conver-
Archaeal lipids, which are radically different from eukaryotic sion of early endosomes into MVBs is the same in yeast and
lipids and rich in rigid, bilayer-spanning tetraether linkages metazoa (Raiborg and Stenmark, 2009). Intralumenal vesicles
(Koga and Morii, 2005). It is thought on theoretical grounds in mammalian cells are highly enriched in cholesterol and tetra-
that membrane tubes are induced by the binding of the curved spanins (Möbius et al., 2003; van der Goot and Gruenberg,
ESCRT-III polymer to the membrane (Lenz et al., 2009). 2006). However, at least some of the cholesterol- and tetraspa-
ESCRT-III polymerization governs the late stage of neck devel- nin-rich intralumenal vesicles in mammalian cells are part of
opment leading to scission, but it is not likely to be the main process that is distinct from the ESCRT pathway (Simons and
factor in the initial budding event. The initial formation of the Raposo, 2009). Raft markers such as long-chain sphingomyelins
bud is driven by the assembly of ESCRT-I and -II with one transit through MVBs (Koivusalo et al., 2007). Consistent with
another and with the endosome membrane (Wollert and Hurley, a possible ESCRT-sterol connection, defects in ESCRT function
2010). The structure of this assembly is unknown, and the nature block endosomal cholesterol transport in mammalian cells
of the assembly is a pressing question in the field. Composite (Bishop and Woodman, 2000; Peck et al., 2004). In yeast, ergos-
structures of the ESCRT-I and -II complexes have been devel- terol and, more speculatively, Sna3 (Piper and Katzmann, 2007)
oped on the basis of crystal structures of the separate compo- might replace the roles of cholesterol and tetraspanins in micro-
nents together with hydrodynamic information of the complete domain formation. Given that ESCRTs bud membranes without

880 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


a coat, and that most other coatless budding mechanisms rely
on membrane microdomains of some sort, it is tempting to
speculate that ESCRT-mediated budding could involve tetra-
spanin and cholesterol-rich domains. Very little is known about
how ESCRTs might couple to such microdomains.
Lipids modifications might also play a role. The ESCRT-III
subunit Vps20 must be myristoylated for full function (Babst
et al., 2002; Yorikawa et al., 2005). Yet even unmyristoylated
Vps20 has an affinity for membranes in the tens of nM, dropping
to low single digit nM range when bound to ESCRT-II (Im et al.,
2009), suggesting that myristoylation is required for another
reason than membrane targeting alone. The myristoyl moiety is
saturated and favors association with Lo phase microdomains
(Resh, 2006). Ubiquitination of tetraspanins (Lineberry et al.,
2008) and, in yeast, Sna3 (Stawiecka-Mirota et al., 2007) has
also been reported.
Another important question surrounds the nature of the PI(3)P
lipid that binds to ESCRT complexes through its headgroup. Figure 6. Lipids and ESCRTs in HIV-1 Assembly
Substantial levels of PI(3)P are found in the MVB lumen (Gillooly Apart from viral proteins, the release of HIV-1 requires both specific cellular
et al., 2000). A critical gap in understanding the formation of in- lipids and proteins, which are recruited to the budding site by the viral Gag
protein. Gag assembles into an imperfect hexagonal lattice on the plasma
tralumenal vesicles is the lack of data on the tail compositions
membrane (Briggs et al., 2009). It binds the plasma membrane marker PI
of the total endosomal and intralumenal vesicle pools of PI(3)P. (4,5)P2 through a specific binding site in its N terminus. PI(4,5)P2, cholesterol,
The concept of an ESCRT-microdomain link is speculative. In and certain other raft lipids are enriched in the viral membrane compared to the
the absence of other explanations for the unusual coatless plasma membrane. Through its C terminus, Gag recruits the ESCRT proteins
to the budding site. Gag can bind both ESCRT-I and ALIX, which both recruit
budding by the ESCRTs, these issues call for further investiga- ESCRT-III to the budding site.
tion.
Animal Cells Have More than One Kind of MVB
Animal cells have additional pathways of MVB formation not cles) with pre-existing phase separation (Trajkovic et al., 2008).
found in yeast. The mammalian late endosomal and lysosomal Sphingomyelinase cleavage of the phosphodiester bond
lipidome contains up to 20% of the unusual lipid lysobisphos- between ceramide and the SM headgroup provides a potential
phatidic acid (LBPA), which is not found in other organelles or mechanism to put energy into this budding pathway and make
in yeast. Mammalian cells have a late endosomal MVB pathway it thermodynamically irreversible. Ceramide-induced intralume-
that seems to depend on LBPA microdomains that are probably nal vesicles bud exclusively from the Lo phase (Trajkovic et al.,
induced on the lumenal leaflet by acidic pH (Matsuo et al., 2004). 2008). Ceramide has several special properties, including a small
The ultimate thermodynamic basis for membrane curvature in headgroup that would favor its presence in the inner leaflet of the
the LBPA pathway would presumably come from the energy intralumenal vesicle and an ability to self-associate through
expended in the pumping of protons into the lumen of the endo- headgroup hydrogen bonding. It is not clear which properties
some. This late endosomal pathway also involves ESCRT of ceramide are most important for the formation of intralumenal
proteins (Falguières et al., 2008). The late endosomal MVB vesicles. Exosomes produced by the sphingomyelinase
pathway should not, however, be confused with the canonical pathway are highly enriched in the tetraspanin CD63, suggestive
early endosomal ESCRT pathway described above, which of a coupling between TEMs and ceramide domains. Of the three
does not involve LBPA. MVB formation is involved in the biogen- pathways described above, the latter two are, based on current
esis of lysosome-related organelles, of which melanosomes are knowledge, ESCRT independent. It will be interesting to see if
the most intensively studied (Raposo and Marks, 2007). In there are ever circumstances under which the ESCRTs coop-
melanosome biogenesis, the glycoprotein Pmel17 is sorted erate with the melanosome or ceramide pathways.
into intralumenal vesicles in an ESCRT-independent reaction
(Theos et al., 2006). Pmel17 is a special cargo in that its lumenal Viral Budding
domain forms fibers and may be an example of the lumenal Enveloped Virus Budding: With ESCRTs and without
assembly of a cargo helping to drive its own inward budding Membrane budding is an essential part of the life cycle of envel-
into the endosome. oped viruses. Most, but not all, enveloped viruses bud from cells
Exosomes are 50–100 nm vesicles released from cells by the by co-opting the host ESCRT machinery (Bieniasz, 2006; Morita
fusion of MVBs with the plasma membrane (Simons and Raposo, and Sundquist, 2004; Welsch et al., 2007), whose role in budding
2009). At least one population of exosomes is produced by an of vesicles in MVBs is described above (Figure 6). Virus budding,
ESCRT-independent pathway in which neutral sphingomyeli- like MVB formation, involves budding away from cytosol. In the
nase, acting from the cytosolic face of the membrane, hydro- well-studied example of HIV-1, formation of the initial plasma-
lyzes sphingomyelin to ceramide (Trajkovic et al., 2008). The membrane attached bud is driven by the energetically favorable
formation of intralumenal vesicles by sphingomyelinase has self-assembly of the capsid (CA) domain of Gag into hexamers
been reconstituted in vitro using GUVs (giant unilamellar vesi- (Briggs et al., 2009; Wright et al., 2007). CA does not bind directly

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 881


to membranes, so the energy of CA self-assembly is transduced The in vitro buds are not cleaved by the matrix protein, indicating
to the membrane through the membrane-binding matrix (MA) the requirement for additional scission factors. Indeed, VSV
domain, part of the same polypeptide chain at this stage in budding from cells requires an ESCRT-I-binding late domain
HIV-1 assembly (Hill et al., 1996). Recombinant HIV-1 Gag (Irie et al., 2004). Why does the matrix protein of one putatively
constructs lacking a part of the membrane-binding MA domain ESCRT-dependent virus, NDV, support both budding and
and all of the ESCRT-binding p6 domain are able to assemble scission on its own, whereas that of another, VSV, supports
with RNA to form spherical shells in vitro, in the absence of only formation of attached buds? It is too soon to say whether
membranes (Campbell et al., 2001). These lipid-free shells are these are intrinsic differences between these viruses or relate
slightly smaller than authentic immature HIV-1 virions, with the merely to experimental differences.
differences accounted for by the absence of membrane and Even for HIV-1, the archetypal ESCRT-dependent virus, there
the MA domain (Briggs et al., 2009). The shells assemble via seem to be circumstances in which ESCRT dependence can be
CA domain hexamers, which cannot pack into a sphere, and circumvented. The effect of mutating its two ESCRT-interacting
therefore a few gaps remain in an otherwise almost complete late domains depends on the cell type, with primary monocyte-
lattice (Briggs et al., 2009). However, in authentic released HIV-1 derived macrophages and the Jurkat T cell line retaining >20%
particles, the Gag shells are only 60% complete on average particle release even when both domains were inactivated (Fujii
(Carlson et al., 2008). Could such an incomplete shell scaffold et al., 2009). Further, replacing the C-terminal part of Gag,
bud formation? Below, we describe the role of membrane micro- including the RNA-binding nucleocapsid domain and the late
domains as another key contributor to HIV-1 bud formation. domain-containing p6 domain, with a leucine zipper motif
In contrast to the self-encoded ability of HIV-1 to form preserves efficient particle release despite absence of ESCRT-
attached buds, the release of these buds from the host cell interacting motifs (Zhang et al., 1998). Deleting part of the
requires the co-option of the host cell ESCRT machinery. nucleocapsid domain and the flanking p1 sequence has the
ESCRT-recruiting motifs known as ‘‘late domains’’ for their func- same effect of making HIV-1 release independent of a functional
tion in late stages of virus assembly and release have been ESCRT machinery (Popova et al., 2010). All in all, these findings
identified in most genera of enveloped viruses (Bieniasz, 2006; show that there is a baseline level of ESCRT-independent HIV-1
Chen and Lamb, 2008; Freed, 2002; Morita and Sundquist, release, which can be elevated by subtle alterations in the Gag
2004). The prototypical example of an ESCRT-dependent virus protein.
is HIV-1, which engages the ESCRT-I complex though a PTAP The studies mentioned above quantified the amount of virus
motif in the p6 region of its Gag protein (Huang et al., 1995), released on a timescale of 16–72 hr, and it is still possible that
and interference with this interaction dramatically reduces the microscopic kinetics of the budding process, which takes
HIV-1 release (Demirov et al., 2002a, 2002b; Garrus et al., place on the timescale of 5–25 min (Ivanchenko et al., 2009;
2001; Martin-Serrano et al., 2001; VerPlank et al., 2001). To Jouvenet et al., 2008), may have been more severely compro-
make matters more complicated, efficient release can be mised. The ESCRT-independent scission observed for NDV
rescued by overexpressing the ESCRT-associated protein in vitro and for certain HIV-1 variants in vivo suggests that in
ALIX, which binds to another motif in Gag p6, YPXnL (Fisher some cases the role of ESCRTs is merely to speed up the final
et al., 2007; Usami et al., 2007). Defects in both of these interac- stage of release. In other cases, such as wild-type HIV-1, the
tions can be rescued by overexpression of HECT domain ubiqui- ESCRTs appear to have a deeper role in viral morphogenesis
tin ligases (Chung et al., 2008; Jadwin et al., 2010; Usami et al., (Carlson et al., 2008).
2008). All of these interactions serve the same ultimate purpose Membrane Microdomains and Influenza Budding
of recruiting ESCRT-III to the nascent viral bud for scission, The influenza virus is the best characterized example of an envel-
which is thought to be carried out by the same process as for oped virus that buds without an ESCRT. Influenza does not have
cleavage of intralumenal vesicles in MVBs (Hurley and Hanson, a typical late domain sequence, nor is its budding inhibited by
2010). overexpressing a dominant-negative Vps4 (Bruce et al., 2009;
If HIV-1 is the archetype of a virus dependent on the host cell Chen et al., 2007).
machinery for membrane scission, other viruses appear to carry Influenza virus associates with lipid rafts via the transmem-
out both budding and scission entirely with virally encoded brane domains of hemagglutinin and neuraminidase (Barman
proteins. The membrane-associated matrix protein of Newcastle et al., 2004), and the membrane of released influenza virions
disease virus (NDV, a paramyxovirus) induces both bud forma- has a pronounced raft character with higher order than the
tion and scission when assembled on model membranes (Shnyr- membranes of non-raft-associated enveloped virus (Polozov
ova et al., 2007). The release of virus-like particles is stimulated et al., 2008; Scheiffele et al., 1999). This raft association serves
by negatively charged lipids and cholesterol. NDV contains a late to cluster hemagglutinin on the plasma membrane, thus
domain motif identical to that of the closely related ESCRT- increasing its concentration on the released particles (Barman
dependent paramyxovirus SV5 (Schmitt et al., 2005). The func- et al., 2004; Takeda et al., 2003), and it is further involved in
tion, if any, of ESCRTs in NDV release might be to accelerate the sorting of hemagglutinin and neuraminidase to the apical
vesiculation, which the virus already is capable of performing. face of polarized cells (Barman et al., 2004).
The matrix protein of vesicular stomatitis virus (VSV) is capable Influenza’s M2 ion channel has recently been implicated in its
of inducing membrane buds in vitro (Solon et al., 2005). In vitro budding mechanism, reconciling its ESCRT independence and
VSV budding occurs in a simple mixture of acidic phospholipids raft association (Rossman et al., 2010). M2 has a conserved
and appears to be driven by self-assembly of the matrix protein. amphipathic helix that is sufficient for vesicle scission in

882 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


a minimal in vitro system, where it predominantly acts at the The binding of PI(4,5)P2 to Gag triggers the myristoyl switch,
border between Ld and cholesterol-enriched Lo domains. M2 leading to exposure of the buried myristoyl group (Saad et al.,
was further localized to the necks of budding influenza particles 2006). In the solution structure of the myristoylated matrix
by immunoelectron microscopy, and mutations disrupting its domain complex bound to a short-chain PI(4,5)P2, the myristoyl
amphipathic helix appear to increase the number of virus buds and the 10 fatty acid tail of PI(4,5)P2 extend into the lipid bilayer,
that remain associated with the cell. This is the first detailed whereas the 20 fatty acid tail of PI(4,5)P2 becomes buried in
description of an ESCRT-independent viral budding mechanism, a pocket in the matrix domain vacated by ejection of the myris-
and it will be interesting to see if it is paralleled in other systems. tate (Saad et al., 2006). In the current view of this mechanism,
How HIV-1 Uses Raft and Non-Raft Lipids to Bud the 10 tail is preferentially saturated and the 20 preferentially
from Cells unsaturated. Thus the matrix domain-PI(4,5)P2 complex would
The HIV-1 membrane is highly ordered (Aloia et al., 1993; in this scheme expose two saturated chains, transforming it
Lorizate et al., 2009), with elevated levels of cholesterol and into a raftophile. It will be interesting to see whether any cellular
certain other raft lipids (GM3 and ceramide) compared to the budding proteins—perhaps including the myristoylated ESCRT-
plasma membrane from which they bud (Brügger et al., 2006; III protein Vps20—use similar mechanisms to bridge raft and
Chan et al., 2008). Cholesterol depletion blocks HIV-1 particle non-raft lipids. HIV-1 release, with its exploitation of so many
release by inhibiting membrane binding and multimerization of of the known physiological budding paradigms in a single event,
Gag (Ono et al., 2007). Thus, the lipid segregation at HIV-1 is one of the most remarkable illustrations of how the dance
budding sites clearly has a functional role in the formation and between proteins and lipids leads to membrane buds.
release of HIV-1 particles. What, precisely, is this role? It is
tempting to speculate that microdomain formation not only Concluding Remarks
contributes to the normal HIV-1 budding pathway but facilitates We hope to have provided a few examples of how the geometry,
the ESCRT-independent budding noted above for unusual HIV-1 topology, and energetics of some selected membrane-budding
Gag constructs (for instance lacking the PTAP motif). However, events in cells are adapted to their biological functions. Trans-
note that cholesterol depletion actually promotes HIV-1 budding port through cytosolic vesicular carriers of membrane proteins
in the case of the PTAP-defective virus that buds independent of that have cytosolic tails is carried out most often through vesicles
the ESCRTs (Ono and Freed, 2001). This suggests that as with coated by the clathrin, COP I, and COP II complexes, which we
the ESCRT-independent budding of influenza, cholesterol has now know to have structural similarities to one another (Lee and
multiple roles. Goldberg, 2010). The cytosolic tails provide the signal for
HIV-1 and other retroviruses use protein-lipid interactions to assembly, coat proteins scaffold the membrane, amphipathic
target their assembly to the plasma membrane. The N-terminal helix and BAR domain factors help bend the membrane, and
matrix domain of HIV-1 Gag has a basic surface (Hill et al., uncoating-coupled hydrolysis of ATP or GTP provides the ther-
1996) and a covalently bound myristoyl fatty acid chain that is modynamic driving force. In the evolution of coats, the tradeoff
necessary for virus release (Ono and Freed, 1999). The ‘‘myristoyl has been between the benefits of flexibility and scaffolding
switch’’ model describes how this myristoyl moiety is in a buried power, with clathrin apparently optimized for flexibility, whereas
conformation in the monomeric cytosolic protein and becomes COP II is optimized as a more potent membrane-curving
exposed upon Gag oligomerization (Saad et al., 2006, 2008; scaffold.
Tang et al., 2004). Thus, the membrane binding of the Gag protein Viruses and toxins often enter cells by engaging with host
is linked to its multimerization and assembly into a lattice. The transmembrane proteins and co-opting coat-dependent
weak membrane affinity of the MA myristate and nonspecific budding mechanisms, but the defensive evolution of host organ-
interactions between the basic face of the matrix domain and isms combats this. Lipid-based entry through the induction of
bulk acidic phospholipids are not sufficient for efficient HIV-1 membrane microdomains, as exemplified by SV40, Shiga toxin,
particle release. For release to occur, the particle assembly and cholera toxin, illustrates one way that pathogens avoid
must be targeted either to the plasma membrane or to membra- having to rely on mutable surface proteins of the host. The phys-
nous compartments that can fuse with the plasma membrane, ical basis of this entry mechanism uses completely different prin-
leading to virion release. PI(4,5)P2, described above as a key ciples to the same functional end. Caveolae present a fascinating
factor in the formation of clathrin-coated vesicles, is the defining hybrid of protein scaffolding and membrane microdomain mech-
lipid marker of the plasma membrane (McLaughlin et al., 2002). anisms. The real cellular function(s) of caveolae are enigmatic,
The matrix domain of HIV-1 Gag targets specifically to the plasma leaving us for now in the dark as to the evolutionary drive for
membrane by binding tightly to the phosphoinositide PI(4,5)P2, such unusual structures.
and this interaction is required for Gag assembly and HIV-1 The ESCRT system, the main interest of our laboratory, is
budding (Ono et al., 2004). adapted for budding away from the cytosol in the opposite
How can the raft dependence of HIV-1 Gag assembly be topology of conventional coated vesicles. The ESCRT system
reconciled with its dependence on PI(4,5)P2? PI(4,5)P2 is gener- evolved to avoid the use of a protein coat because of this unusual
ally considered a non-raft lipid, although the microscopic anal- topology. The unique mechanism by which ESCRTs stabilize
ysis of the tail composition of different pools of PI(4,5)P2 is not and sever membrane buds has become much clearer over the
elaborated to the point where this can be said with certainty past year. However, the pathway of early bud formation, before
for all PI(4,5)P2. The apparent answer to this question highlights the bud neck has contracted enough for the ESCRT proteins to
the frightening ingenuity of HIV-1 in co-opting cellular systems. bridge across it, is still obscure. This led us to ask whether

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 883


membrane microdomains might have a role in ESCRT-mediated Barman, S., Adhikary, L., Chakrabarti, A.K., Bernas, C., Kawaoka, Y., and
bud formation. If this were the case, membrane microdomains Nayak, D.P. (2004). Role of transmembrane domain and cytoplasmic tail amino
acid sequences of influenza a virus neuraminidase in raft association and virus
might serve as a unifying principle connecting the diverse types
budding. J. Virol. 78, 5258–5269.
of ESCRT-dependent and microdomain-dependent MVBs in
animal cells. The various ESCRT- and microdomain-dependent Bassereau, P. (2010). Division of labour in ESCRT complexes. Nat. Cell Biol.
12, 422–423.
flavors of enveloped virus budding mirror the distinct varieties
of animal cell MVBs. This is not surprising given that these two Bauer, M., and Pelkmans, L. (2006). A new paradigm for membrane-organizing
and -shaping scaffolds. FEBS Lett. 580, 5559–5564.
processes share the same unusual property of budding away
from cytosol. Microdomains and ESCRTs have the same advan- Baumgart, T., Hess, S.T., and Webb, W.W. (2003). Imaging coexisting fluid
domains in biomembrane models coupling curvature and line tension. Nature
tage for budding away from cytosol, in that cytosolic coat
425, 821–824.
proteins need not be irreversibly consumed in the process.
Membrane budding and the related topic of membrane tubu- Baumgart, T., Hammond, A.T., Sengupta, P., Hess, S.T., Holowka, D.A., Baird,
B.A., and Webb, W.W. (2007). Large-scale fluid/fluid phase separation of
lation have become exceptionally vibrant fields, driven by
proteins and lipids in giant plasma membrane vesicles. Proc. Natl. Acad.
advances in technology. Computational resources now allow Sci. USA 104, 3165–3170.
sophisticated simulations of budding (Reynwar et al., 2007).
Bi, X., Corpina, R.A., and Goldberg, J. (2002). Structure of the Sec23/24-Sar1
Reconstitution of budding events from completely defined pre-budding complex of the COPII vesicle coat. Nature 419, 271–277.
systems (Bremser et al., 1999; Matsuoka et al., 1998; Römer
Bieniasz, P.D. (2006). Late budding domains and host proteins in enveloped
et al., 2007; Wollert and Hurley, 2010) has established molecular
virus release. Virology 344, 55–63.
mechanisms in several cases and opened the door to more
Bishop, N., and Woodman, P. (2000). ATPase-defective mammalian VPS4
sophisticated biophysical analysis (Bassereau, 2010). Electron
localizes to aberrant endosomes and impairs cholesterol trafficking. Mol.
microscopy has been the foundation of our understanding of Biol. Cell 11, 227–239.
membrane budding in cells since the beginning. Looking
Bloom, M., Evans, E., and Mouritsen, O.G. (1991). Physical properties of the
forward, advanced electron tomography will undoubtedly shape fluid lipid-bilayer component of cell membranes: a perspective. Q. Rev. Bio-
our future views of how membranes bud, as classical electron phys. 24, 293–397.
microscopy has in the past and present. As in other areas of
Bonifacino, J.S., and Lippincott-Schwartz, J. (2003). Coat proteins: shaping
cell biology, rapid advances in live-cell imaging are making membrane transport. Nat. Rev. Mol. Cell Biol. 4, 409–414.
powerful and ever-increasing contributions. Membrane budding
Bremser, M., Nickel, W., Schweikert, M., Ravazzola, M., Amherdt, M., Hughes,
is a required part of the life cycle of two of the most dangerous C.A., Söllner, T.H., Rothman, J.E., and Wieland, F.T. (1999). Coupling of coat
human pathogens, HIV and influenza, and insight into the funda- assembly and vesicle budding to packaging of putative cargo receptors.
mental nature of these budding events is perhaps the most Cell 96, 495–506.
urgently needed of all. Briggs, J.A.G., Riches, J.D., Glass, B., Bartonova, V., Zanetti, G., and
Kräusslich, H.G. (2009). Structure and assembly of immature HIV. Proc. Natl.
ACKNOWLEDGMENTS Acad. Sci. USA 106, 11090–11095.
Brodsky, F.M., Chen, C.Y., Knuehl, C., Towler, M.C., and Wakeham, D.E.
We thank E. Freed, G. Raposo, W. Prinz, M. Marks, J. Gruenberg, and (2001). Biological basket weaving: formation and function of clathrin-coated
J. Bonifacino for comments on drafts of the manuscript, J. Heuser for providing vesicles. Annu. Rev. Cell Dev. Biol. 17, 517–568.
the image used in Figure 1F, and many colleagues for stimulating discussions.
Bruce, E.A., Medcalf, L., Crump, C.M., Noton, S.L., Stuart, A.D., Wise, H.M.,
Research in the Hurley laboratory is supported the Intramural program of the
Elton, D., Bowers, K., and Digard, P. (2009). Budding of filamentous and
National Institutes of Health, NIDDK, and IATAP. B.R. was supported by
non-filamentous influenza A virus occurs via a VPS4 and VPS28-independent
a Marie Curie International Outgoing Fellowship within the 7th European
pathway. Virology 390, 268–278.
Community Framework Programme.
Brügger, B., Glass, B., Haberkant, P., Leibrecht, I., Wieland, F.T., and Kräus-
slich, H.G. (2006). The HIV lipidome: a raft with an unusual composition.
REFERENCES Proc. Natl. Acad. Sci. USA 103, 2641–2646.

Aloia, R.C., Tian, H.R., and Jensen, F.C. (1993). Lipid composition and fluidity Campbell, S., Fisher, R.J., Towler, E.M., Fox, S., Issaq, H.J., Wolfe, T., Phillips,
of the human immunodeficiency virus envelope and host cell plasma L.R., and Rein, A. (2001). Modulation of HIV-like particle assembly in vitro by
membranes. Proc. Natl. Acad. Sci. USA 90, 5181–5185. inositol phosphates. Proc. Natl. Acad. Sci. USA 98, 10875–10879.

Babst, M., Wendland, B., Estepa, E.J., and Emr, S.D. (1998). The Vps4p AAA Carlson, L.A., Briggs, J.A., Glass, B., Riches, J.D., Simon, M.N., Johnson,
ATPase regulates membrane association of a Vps protein complex required M.C., Müller, B., Grünewald, K., and Kräusslich, H.G. (2008). Three-dimen-
for normal endosome function. EMBO J. 17, 2982–2993. sional analysis of budding sites and released virus suggests a revised model
for HIV-1 morphogenesis. Cell Host Microbe 4, 592–599.
Babst, M., Katzmann, D.J., Estepa-Sabal, E.J., Meerloo, T., and Emr, S.D.
(2002). Escrt-III: An endosome-associated heterooligomeric protein complex Chan, R., Uchil, P.D., Jin, J., Shui, G.H., Ott, D.E., Mothes, W., and Wenk, M.R.
required for mvb sorting. Dev. Cell 3, 271–282. (2008). Retroviruses human immunodeficiency virus and murine leukemia virus
are enriched in phosphoinositides. J. Virol. 82, 11228–11238.
Bacia, K., Schwille, P., and Kurzchalia, T. (2005). Sterol structure determines
the separation of phases and the curvature of the liquid-ordered phase in Chen, B.J., and Lamb, R.A. (2008). Mechanisms for enveloped virus budding:
model membranes. Proc. Natl. Acad. Sci. USA 102, 3272–3277. can some viruses do without an ESCRT? Virology 372, 221–232.
Bajorek, M., Schubert, H.L., McCullough, J., Langelier, C., Eckert, D.M., Chen, B.J., Leser, G.P., Morita, E., and Lamb, R.A. (2007). Influenza virus
Stubblefield, W.M.B., Uter, N.T., Myszka, D.G., Hill, C.P., and Sundquist, hemagglutinin and neuraminidase, but not the matrix protein, are required
W.I. (2009). Structural basis for ESCRT-III protein autoinhibition. Nat. Struct. for assembly and budding of plasmid-derived virus-like particles. J. Virol. 81,
Mol. Biol. 16, 754–762. 7111–7123.

884 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


Chiantia, S., Kahya, N., Ries, J., and Schwille, P. (2006). Effects of ceramide ordered-liquid disordered phase separation in model plasma membranes.
on liquid-ordered domains investigated by simultaneous AFM and FCS. Proc. Natl. Acad. Sci. USA 102, 6320–6325.
Biophys. J. 90, 4500–4508. Hancock, J.F. (2006). Lipid rafts: contentious only from simplistic standpoints.
Chung, H.Y., Morita, E., von Schwedler, U., Müller, B., Kräusslich, H.G., and Nat. Rev. Mol. Cell Biol. 7, 456–462.
Sundquist, W.I. (2008). NEDD4L overexpression rescues the release and infec- Hansen, C.G., and Nichols, B.J. (2010). Exploring the caves: cavins, caveolins
tivity of human immunodeficiency virus type 1 constructs lacking PTAP and and caveolae. Trends Cell Biol. 20, 177–186.
YPXL late domains. J. Virol. 82, 4884–4897.
Hanson, P.I., Roth, R., Lin, Y., and Heuser, J.E. (2008). Plasma membrane
Damm, E.M., Pelkmans, L., Kartenbeck, J., Mezzacasa, A., Kurzchalia, T., and
deformation by circular arrays of ESCRT-III protein filaments. J. Cell Biol.
Helenius, A. (2005). Clathrin- and caveolin-1-independent endocytosis: entry
180, 389–402.
of simian virus 40 into cells devoid of caveolae. J. Cell Biol. 168, 477–488.
Helfrich, W. (1973). Elastic properties of lipid bilayers: theory and possible
Demirov, D.G., Ono, A., Orenstein, J.M., and Freed, E.O. (2002a). Overexpres-
experiments. Z. Naturforsch. C 28, 693–703.
sion of the N-terminal domain of TSG101 inhibits HIV-1 budding by blocking
late domain function. Proc. Natl. Acad. Sci. USA 99, 955–960. Heller, H., Schaefer, M., and Schulten, K. (1993). Molecular dynamics simula-
tion of a bilayer of 200 lipids in the gel and in the liquid crystal phases. J. Phys.
Demirov, D.G., Orenstein, J.M., and Freed, E.O. (2002b). The late domain of
Chem. 97, 8343–8360.
human immunodeficiency virus type 1 p6 promotes virus release in a cell
type-dependent manner. J. Virol. 76, 105–117. Hemler, M.E. (2005). Tetraspanin functions and associated microdomains.
Nat. Rev. Mol. Cell Biol. 6, 801–811.
Ehrlich, M., Boll, W., Van Oijen, A., Hariharan, R., Chandran, K., Nibert, M.L.,
and Kirchhausen, T. (2004). Endocytosis by random initiation and stabilization Henne, W.M., Boucrot, E., Meinecke, M., Evergren, E., Vallis, Y., Mittal, R., and
of clathrin-coated pits. Cell 118, 591–605. McMahon, H.T. (2010). FCHo proteins are nucleators of clathrin-mediated
endocytosis. Science 328, 1281–1284.
Eisenberg, E., and Greene, L.E. (2007). Multiple roles of auxilin and hsc70 in
clathrin-mediated endocytosis. Traffic 8, 640–646. Hill, C.P., Worthylake, D., Bancroft, D.P., Christensen, A.M., and Sundquist,
W.I. (1996). Crystal structures of the trimeric human immunodeficiency virus
Elson, E.L., Fried, E., Dolbow, J.E., and Genin, G.M. (2010). Phase separation
type 1 matrix protein: implications for membrane association and assembly.
in biological membranes: Integration of theory and experiment. Annu. Rev.
Proc. Natl. Acad. Sci. USA 93, 3099–3104.
Biophys. 39, 207–226.
Ewers, H., Römer, W., Smith, A.E., Bacia, K., Dmitrieff, S., Chai, W.G., Mancini, Hinrichsen, L., Meyerholz, A., Groos, S., and Ungewickell, E.J. (2006). Bending
R., Kartenbeck, J., Chambon, V., Berland, L., et al. (2010). GM1 structure a membrane: how clathrin affects budding. Proc. Natl. Acad. Sci. USA 103,
determines SV40-induced membrane invagination and infection. Nat. Cell 8715–8720.
Biol. 12, 11–18, 1–12. Hu, J.J., Shibata, Y., Voss, C., Shemesh, T., Li, Z.L., Coughlin, M., Kozlov,
Fabrikant, G., Lata, S., Riches, J.D., Briggs, J.A.G., Weissenhorn, W., and M.M., Rapoport, T.A., and Prinz, W.A. (2008). Membrane proteins of the endo-
Kozlov, M.M. (2009). Computational model of membrane fission catalyzed plasmic reticulum induce high-curvature tubules. Science 319, 1247–1250.
by ESCRT-III. PLoS Comput. Biol. 5, e1000575. Huang, M.J., Orenstein, J.M., Martin, M.A., and Freed, E.O. (1995). p6Gag is
Falguières, T., Luyet, P.P., Bissig, C., Scott, C.C., Velluz, M.C., and required for particle production from full-length human immunodeficiency
Gruenberg, J. (2008). In vitro budding of intralumenal vesicles into late endo- virus type 1 molecular clones expressing protease. J. Virol. 69, 6810–6818.
somes is regulated by Alix and Tsg101. Mol. Biol. Cell 19, 4942–4955. Hurley, J.H., and Hanson, P.I. (2010). Membrane budding and scission by the
Farsad, K., and De Camilli, P. (2003). Mechanisms of membrane deformation. ESCRT machinery: it’s all in the neck. Nat. Rev. Mol. Cell. Biol. 8, 556–566.
Curr. Opin. Cell Biol. 15, 372–381. Im, Y.J., and Hurley, J.H. (2008). Integrated structural model and membrane
Fisher, R.D., Chung, H.Y., Zhai, Q.T., Robinson, H., Sundquist, W.I., and targeting mechanism of the human ESCRT-II complex. Dev. Cell 14, 902–913.
Hill, C.P. (2007). Structural and biochemical studies of ALIX/AIP1 and its role Im, Y.J., Wollert, T., Boura, E., and Hurley, J.H. (2009). Structure and function
in retrovirus budding. Cell 128, 841–852. of the ESCRT-II-III interface in multivesicular body biogenesis. Dev. Cell 17,
Ford, M.G.J., Mills, I.G., Peter, B.J., Vallis, Y., Praefcke, G.J.K., Evans, P.R., 234–243.
and McMahon, H.T. (2002). Curvature of clathrin-coated pits driven by epsin. Irie, T., Licata, J.M., Jayakar, H.R., Whitt, M.A., Bell, P., and Harty, R.N. (2004).
Nature 419, 361–366. Functional analysis of late-budding domain activity associated with the PSAP
Fotin, A., Cheng, Y.F., Sliz, P., Grigorieff, N., Harrison, S.C., Kirchhausen, T., motif within the vesicular stomatitis virus M protein. J. Virol. 78, 7823–7827.
and Walz, T. (2004). Molecular model for a complete clathrin lattice from elec- Ivanchenko, S., Godinez, W.J., Lampe, M., Kräusslich, H.G., Eils, R., Rohr, K.,
tron cryomicroscopy. Nature 432, 573–579. Bräuchle, C., Müller, B., and Lamb, D.C. (2009). Dynamics of HIV-1 assembly
Freed, E.O. (2002). Viral late domains. J. Virol. 76, 4679–4687. and release. PLoS Pathog. 5, e1000652.
Fujii, K., Munshi, U.M., Ablan, S.D., Demirov, D.G., Soheilian, F., Nagashima, Jackson, L.P., Kelly, B.T., McCoy, A.J., Gaffry, T., James, L.C., Collins, B.M.,
K., Stephen, A.G., Fisher, R.J., and Freed, E.O. (2009). Functional role of Alix Höning, S., Evans, P.R., and Owen, D.J. (2010). A large-scale conformational
in HIV-1 replication. Virology 391, 284–292. change couples membrane recruitment to cargo binding in the AP2 clathrin
Garcı́a-Sáez, A.J., and Schwille, P. (2010). Stability of lipid domains. FEBS adaptor complex. Cell 141, 1220–1229.
Lett. 584, 1653–1658. Jadwin, J.A., Rudd, V., Sette, P., Challa, S., and Bouamr, F. (2010). Late
Garrus, J.E., von Schwedler, U.K., Pornillos, O.W., Morham, S.G., Zavitz, K.H., domain-independent rescue of a release-deficient Moloney murine leukemia
Wang, H.E., Wettstein, D.A., Stray, K.M., Côté, M., Rich, R.L., et al. (2001). virus by the ubiquitin ligase itch. J. Virol. 84, 704–715.
Tsg101 and the vacuolar protein sorting pathway are essential for HIV-1 Jouvenet, N., Bieniasz, P.D., and Simon, S.M. (2008). Imaging the biogenesis
budding. Cell 107, 55–65. of individual HIV-1 virions in live cells. Nature 454, 236–240.
Gillooly, D.J., Morrow, I.C., Lindsay, M., Gould, R., Bryant, N.J., Gaullier, J.M., Kirkham, M., Fujita, A., Chadda, R., Nixon, S.J., Kurzchalia, T.V., Sharma, D.K.,
Parton, R.G., and Stenmark, H. (2000). Localization of phosphatidylinositol Pagano, R.E., Hancock, J.F., Mayor, S., and Parton, R.G. (2005). Ultrastruc-
3-phosphate in yeast and mammalian cells. EMBO J. 19, 4577–4588. tural identification of uncoated caveolin-independent early endocytic vehicles.
Gruenberg, J., and Stenmark, H. (2004). The biogenesis of multivesicular en- J. Cell Biol. 168, 465–476.
dosomes. Nat. Rev. Mol. Cell Biol. 5, 317–323. Kitadokoro, K., Bordo, D., Galli, G., Petracca, R., Falugi, F., Abrignani, S.,
Hammond, A.T., Heberle, F.A., Baumgart, T., Holowka, D., Baird, B., and Fei- Grandi, G., and Bolognesi, M. (2001). CD81 extracellular domain 3D structure:
genson, G.W. (2005). Crosslinking a lipid raft component triggers liquid insight into the tetraspanin superfamily structural motifs. EMBO J. 20, 12–18.

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 885


Koga, Y., and Morii, H. (2005). Recent advances in structural research on ether Morita, E., and Sundquist, W.I. (2004). Retrovirus budding. Annu. Rev. Cell
lipids from archaea including comparative and physiological aspects. Biosci. Dev. Biol. 20, 395–425.
Biotechnol. Biochem. 69, 2019–2034. Murk, J.L.A.N., Humbel, B.M., Ziese, U., Griffith, J.M., Posthuma, G.,
Koivusalo, M., Jansen, M., Somerharju, P., and Ikonen, E. (2007). Endocytic Slot, J.W., Koster, A.J., Verkleij, A.J., Geuze, H.J., and Kleijmeer, M.J.
trafficking of sphingomyelin depends on its acyl chain length. Mol. Biol. Cell (2003). Endosomal compartmentalization in three dimensions: implications
18, 5113–5123. for membrane fusion. Proc. Natl. Acad. Sci. USA 100, 13332–13337.
Kostelansky, M.S., Schluter, C., Tam, Y.Y.C., Lee, S., Ghirlando, R., Beach, B., Neu, U., Woellner, K., Gauglitz, G., and Stehle, T. (2008). Structural basis of
Conibear, E., and Hurley, J.H. (2007). Molecular architecture and functional GM1 ganglioside recognition by simian virus 40. Proc. Natl. Acad. Sci. USA
model of the complete yeast ESCRT-I heterotetramer. Cell 129, 485–498. 105, 5219–5224.
Lata, S., Schoehn, G., Jain, A., Pires, R., Piehler, J., Gottlinger, H.G., and Weis- Nossal, R. (2001). Energetics of clathrin basket assembly. Traffic 2, 138–147.
senhorn, W. (2008). Helical structures of ESCRT-III are disassembled by VPS4. Ono, A., and Freed, E.O. (1999). Binding of human immunodeficiency virus
Science 321, 1354–1357. type 1 Gag to membrane: role of the matrix amino terminus. J. Virol. 73,
Lee, C., and Goldberg, J. (2010). Structure of coatomer cage proteins and the 4136–4144.
relationship among COPI, COPII, and clathrin vesicle coats. Cell 142, 123–132. Ono, A., and Freed, E.O. (2001). Plasma membrane rafts play a critical role in
Lee, M.C.S., Orci, L., Hamamoto, S., Futai, E., Ravazzola, M., and HIV-1 assembly and release. Proc. Natl. Acad. Sci. USA 98, 13925–13930.
Schekman, R. (2005). Sar1p N-terminal helix initiates membrane curvature Ono, A., Ablan, S.D., Lockett, S.J., Nagashima, K., and Freed, E.O. (2004).
and completes the fission of a COPII vesicle. Cell 122, 605–617. Phosphatidylinositol (4,5) bisphosphate regulates HIV-1 Gag targeting to the
Lenz, M., Crow, D.J.G., and Joanny, J.F. (2009). Membrane buckling induced plasma membrane. Proc. Natl. Acad. Sci. USA 101, 14889–14894.
by curved filaments. Phys. Rev. Lett. 103, 038101. Ono, A., Waheed, A.A., and Freed, E.O. (2007). Depletion of cellular cholesterol
Lindås, A.C., Karlsson, E.A., Lindgren, M.T., Ettema, T.J.G., and Bernander, R. inhibits membrane binding and higher-order multimerization of human immu-
(2008). A unique cell division machinery in the Archaea. Proc. Natl. Acad. Sci. nodeficiency virus type 1 Gag. Virology 360, 27–35.
USA 105, 18942–18946. Parton, R.G., and Simons, K. (2007). The multiple faces of caveolae. Nat. Rev.
Lineberry, N., Su, L., Soares, L., and Fathman, C.G. (2008). The single subunit Mol. Cell Biol. 8, 185–194.
transmembrane E3 ligase gene related to anergy in lymphocytes (GRAIL) Parton, R.G., Hanzal-Bayer, M., and Hancock, J.F. (2006). Biogenesis of cav-
captures and then ubiquitinates transmembrane proteins across the cell eolae: a structural model for caveolin-induced domain formation. J. Cell Sci.
membrane. J. Biol. Chem. 283, 28497–28505. 119, 787–796.
Lingwood, D., and Simons, K. (2010). Lipid rafts as a membrane-organizing Peck, J.W., Bowden, E.T., and Burbelo, P.D. (2004). Structure and function of
principle. Science 327, 46–50. human Vps20 and Snf7 proteins. Biochem. J. 377, 693–700.
Lipowsky, R. (1992). Budding of membranes induced by intramembrane Pelkmans, L., and Zerial, M. (2005). Kinase-regulated quantal assemblies and
domains. J. Phys. II France 2, 1825–1840. kiss-and-run recycling of caveolae. Nature 436, 128–133.
Liu, J., Kaksonen, M., Drubin, D.G., and Oster, G. (2006). Endocytic vesicle Pike, L.J. (2006). Rafts defined: a report on the Keystone Symposium on Lipid
scission by lipid phase boundary forces. Proc. Natl. Acad. Sci. USA 103, Rafts and Cell Function. J. Lipid Res. 47, 1597–1598.
10277–10282. Piper, R.C., and Katzmann, D.J. (2007). Biogenesis and function of multivesic-
Lorizate, M., Brügger, B., Akiyama, H., Glass, B., Müller, B., Anderluh, G., Wie- ular bodies. Annu. Rev. Cell Dev. Biol. 23, 519–547.
land, F.T., and Kräusslich, H.G. (2009). Probing HIV-1 membrane liquid order Polozov, I.V., Bezrukov, L., Gawrisch, K., and Zimmerberg, J. (2008). Progres-
by Laurdan staining reveals producer cell-dependent differences. J. Biol. sive ordering with decreasing temperature of the phospholipids of influenza
Chem. 284, 22238–22247. virus. Nat. Chem. Biol. 4, 248–255.
Martin-Serrano, J., Zang, T., and Bieniasz, P.D. (2001). HIV-1 and Ebola virus Pols, M.S., and Klumperman, J. (2009). Trafficking and function of the tetra-
encode small peptide motifs that recruit Tsg101 to sites of particle assembly to spanin CD63. Exp. Cell Res. 315, 1584–1592.
facilitate egress. Nat. Med. 7, 1313–1319. Popova, E., Popov, S., and Göttlinger, H.G. (2010). Human immunodeficiency
Matsuo, H., Chevallier, J., Mayran, N., Le Blanc, I., Ferguson, C., Fauré, J., virus type 1 nucleocapsid p1 confers ESCRT pathway dependence. J. Virol.
Blanc, N.S., Matile, S., Dubochet, J., Sadoul, R., et al. (2004). Role of LBPA 84, 6590–6597.
and Alix in multivesicular liposome formation and endosome organization. Pucadyil, T.J., and Schmid, S.L. (2009). Conserved functions of membrane
Science 303, 531–534. active GTPases in coated vesicle formation. Science 325, 1217–1220.
Matsuoka, K., Orci, L., Amherdt, M., Bednarek, S.Y., Hamamoto, S., Schek- Raiborg, C., and Stenmark, H. (2009). The ESCRT machinery in endosomal
man, R., and Yeung, T. (1998). COPII-coated vesicle formation reconstituted sorting of ubiquitylated membrane proteins. Nature 458, 445–452.
with purified coat proteins and chemically defined liposomes. Cell 93, Raposo, G., and Marks, M.S. (2007). Melanosomes—dark organelles
263–275. enlighten endosomal membrane transport. Nat. Rev. Mol. Cell Biol. 8,
McLaughlin, S., Wang, J.Y., Gambhir, A., and Murray, D. (2002). PIP(2) and 786–797.
proteins: interactions, organization, and information flow. Annu. Rev. Biophys. Reider, A., Barker, S.L., Mishra, S.K., Im, Y.J., Maldonado-Báez, L., Hurley,
Biomol. Struct. 31, 151–175. J.H., Traub, L.M., and Wendland, B. (2009). Syp1 is a conserved endocytic
McMahon, H.T., and Gallop, J.L. (2005). Membrane curvature and mecha- adaptor that contains domains involved in cargo selection and membrane
nisms of dynamic cell membrane remodelling. Nature 438, 590–596. tubulation. EMBO J. 28, 3103–3116.
Merritt, E.A., Sarfaty, S., van den Akker, F., L’Hoir, C., Martial, J.A., and Hol, Ren, X.F., and Hurley, J.H. (2010). VHS domains of ESCRT-0 cooperate in
W.G.J. (1994). Crystal structure of cholera toxin B-pentamer bound to receptor high-avidity binding to polyubiquitinated cargo. EMBO J. 29, 1045–1054.
GM1 pentasaccharide. Protein Sci. 3, 166–175. Resh, M.D. (2006). Trafficking and signaling by fatty-acylated and prenylated
Misra, S., Miller, G.J., and Hurley, J.H. (2001). Recognizing phosphatidylinosi- proteins. Nat. Chem. Biol. 2, 584–590.
tol 3-phosphate. Cell 107, 559–562. Reynwar, B.J., Illya, G., Harmandaris, V.A., Müller, M.M., Kremer, K., and
Möbius, W., van Donselaar, E., Ohno-Iwashita, Y., Shimada, Y., Heijnen, Deserno, M. (2007). Aggregation and vesiculation of membrane proteins by
H.F.G., Slot, J.W., and Geuze, H.J. (2003). Recycling compartments and the curvature-mediated interactions. Nature 447, 461–464.
internal vesicles of multivesicular bodies harbor most of the cholesterol found Robinson, M.S., and Bonifacino, J.S. (2001). Adaptor-related proteins. Curr.
in the endocytic pathway. Traffic 4, 222–231. Opin. Cell Biol. 13, 444–453.

886 Cell 143, December 10, 2010 ª2010 Elsevier Inc.


Rodal, S.K., Skretting, G., Garred, O., Vilhardt, F., van Deurs, B., and Subtil, A., Gaidarov, I., Kobylarz, K., Lampson, M.A., Keen, J.H., and McGraw,
Sandvig, K. (1999). Extraction of cholesterol with methyl-beta-cyclodextrin T.E. (1999). Acute cholesterol depletion inhibits clathrin-coated pit budding.
perturbs formation of clathrin-coated endocytic vesicles. Mol. Biol. Cell 10, Proc. Natl. Acad. Sci. USA 96, 6775–6780.
961–974. Takeda, M., Leser, G.P., Russell, C.J., and Lamb, R.A. (2003). Influenza virus
Römer, W., Berland, L., Chambon, V., Gaus, K., Windschiegl, B., Tenza, D., hemagglutinin concentrates in lipid raft microdomains for efficient viral fusion.
Aly, M.R.E., Fraisier, V., Florent, J.C., Perrais, D., et al. (2007). Shiga toxin Proc. Natl. Acad. Sci. USA 100, 14610–14617.
induces tubular membrane invaginations for its uptake into cells. Nature
Tang, C., Loeliger, E., Luncsford, P., Kinde, I., Beckett, D., and Summers, M.F.
450, 670–675.
(2004). Entropic switch regulates myristate exposure in the HIV-1 matrix
Rossman, J.S., Jing, X., Leser, G.P., and Lamb, R.A. (2010). Influenza virus M2 protein. Proc. Natl. Acad. Sci. USA 101, 517–522.
protein mediates ESCRT-independent membrane scission. Cell 142, 902–913.
Theos, A.C., Truschel, S.T., Tenza, D., Hurbain, I., Harper, D.C., Berson, J.F.,
Russell, C., and Stagg, S.M. (2010). New insights into the structural mecha- Thomas, P.C., Raposo, G., and Marks, M.S. (2006). A lumenal domain-depen-
nisms of the COPII coat. Traffic 11, 303–310. dent pathway for sorting to intralumenal vesicles of multivesicular endosomes
Saad, J.S., Miller, J., Tai, J., Kim, A., Ghanam, R.H., and Summers, M.F. involved in organelle morphogenesis. Dev. Cell 10, 343–354.
(2006). Structural basis for targeting HIV-1 Gag proteins to the plasma
Trajkovic, K., Hsu, C., Chiantia, S., Rajendran, L., Wenzel, D., Wieland, F.,
membrane for virus assembly. Proc. Natl. Acad. Sci. USA 103, 11364–11369.
Schwille, P., Brügger, B., and Simons, M. (2008). Ceramide triggers budding
Saad, J.S., Ablan, S.D., Ghanam, R.H., Kim, A., Andrews, K., Nagashima, K., of exosome vesicles into multivesicular endosomes. Science 319, 1244–1247.
Soheilian, F., Freed, E.O., and Summers, M.F. (2008). Structure of the myristy-
Traub, L.M., and Wendland, B. (2010). Cell biology: How to don a coat. Nature
lated human immunodeficiency virus type 2 matrix protein and the role of
465, 556–557.
phosphatidylinositol-(4,5)-bisphosphate in membrane targeting. J. Mol. Biol.
382, 434–447. Usami, Y., Popov, S., and Göttlinger, H.G. (2007). Potent rescue of human
immunodeficiency virus type 1 late domain mutants by ALIX/AIP1 depends
Saksena, S., Sun, J., Chu, T., and Emr, S.D. (2007). ESCRTing proteins in the
on its CHMP4 binding site. J. Virol. 81, 6614–6622.
endocytic pathway. Trends Biochem. Sci. 32, 561–573.
Samson, R.Y., Obita, T., Freund, S.M., Williams, R.L., and Bell, S.D. (2008). A Usami, Y., Popov, S., Popova, E., and Göttlinger, H.G. (2008). Efficient and
role for the ESCRT system in cell division in archaea. Science 322, 1710–1713. specific rescue of human immunodeficiency virus type 1 budding defects by
a Nedd4-like ubiquitin ligase. J. Virol. 82, 4898–4907.
Sandvig, K., Torgersen, M.L., Raa, H.A., and van Deurs, B. (2008). Clathrin-
independent endocytosis: from nonexisting to an extreme degree of van der Goot, F.G., and Gruenberg, J. (2006). Intra-endosomal membrane
complexity. Histochem. Cell Biol. 129, 267–276. traffic. Trends Cell Biol. 16, 514–521.
Scheiffele, P., Rietveld, A., Wilk, T., and Simons, K. (1999). Influenza viruses Veatch, S.L., Soubias, O., Keller, S.L., and Gawrisch, K. (2007). Critical fluctu-
select ordered lipid domains during budding from the plasma membrane. ations in domain-forming lipid mixtures. Proc. Natl. Acad. Sci. USA 104,
J. Biol. Chem. 274, 2038–2044. 17650–17655.
Schmitt, A.P., Leser, G.P., Morita, E., Sundquist, W.I., and Lamb, R.A. (2005). VerPlank, L., Bouamr, F., LaGrassa, T.J., Agresta, B., Kikonyogo, A., Leis, J.,
Evidence for a new viral late-domain core sequence, FPIV, necessary for and Carter, C.A. (2001). Tsg101, a homologue of ubiquitin-conjugating (E2)
budding of a paramyxovirus. J. Virol. 79, 2988–2997. enzymes, binds the L domain in HIV type 1 Pr55(Gag). Proc. Natl. Acad. Sci.
Shibata, Y., Hu, J.J., Kozlov, M.M., and Rapoport, T.A. (2009). Mechanisms USA 98, 7724–7729.
shaping the membranes of cellular organelles. Annu. Rev. Cell Dev. Biol. 25, Voeltz, G.K., and Prinz, W.A. (2007). Sheets, ribbons and tubules - how organ-
329–354. elles get their shape. Nat. Rev. Mol. Cell Biol. 8, 258–264.
Shnyrova, A.V., Ayllon, J., Mikhalyov, I.I., Villar, E., Zimmerberg, J., and Welsch, S., Müller, B., and Kräusslich, H.G. (2007). More than one door -
Frolov, V.A. (2007). Vesicle formation by self-assembly of membrane-bound Budding of enveloped viruses through cellular membranes. FEBS Lett. 581,
matrix proteins into a fluidlike budding domain. J. Cell Biol. 179, 627–633. 2089–2097.
Simons, M., and Raposo, G. (2009). Exosomes—vesicular carriers for intercel- Williams, R.L., and Urbé, S. (2007). The emerging shape of the ESCRT
lular communication. Curr. Opin. Cell Biol. 21, 575–581. machinery. Nat. Rev. Mol. Cell Biol. 8, 355–368.
Solon, J., Gareil, O., Bassereau, P., and Gaudin, Y. (2005). Membrane defor-
Wollert, T., and Hurley, J.H. (2010). Molecular mechanism of multivesicular
mations induced by the matrix protein of vesicular stomatitis virus in a minimal
body biogenesis by ESCRT complexes. Nature 464, 864–869.
system. J. Gen. Virol. 86, 3357–3363.
Wollert, T., Wunder, C., Lippincott-Schwartz, J., and Hurley, J.H. (2009).
Sorkin, A., and von Zastrow, M. (2009). Endocytosis and signalling: intertwin-
Membrane scission by the ESCRT-III complex. Nature 458, 172–177.
ing molecular networks. Nat. Rev. Mol. Cell Biol. 10, 609–622.
Stagg, S.M., Gürkan, C., Fowler, D.M., LaPointe, P., Foss, T.R., Potter, C.S., Wright, E.R., Schooler, J.B., Ding, H.J., Kieffer, C., Fillmore, C., Sundquist,
Carragher, B., and Balch, W.E. (2006). Structure of the Sec13/31 COPII coat W.I., and Jensen, G.J. (2007). Electron cryotomography of immature HIV-1
cage. Nature 439, 234–238. virions reveals the structure of the CA and SP1 Gag shells. EMBO J. 26,
2218–2226.
Stagg, S.M., LaPointe, P., Razvi, A., Gürkan, C., Potter, C.S., Carragher, B.,
and Balch, W.E. (2008). Structural basis for cargo regulation of COPII coat Yorikawa, C., Shibata, H., Waguri, S., Hatta, K., Horii, M., Katoh, K., Kobaya-
assembly. Cell 134, 474–484. shi, T., Uchiyama, Y., and Maki, M. (2005). Human CHMP6, a myristoylated
ESCRT-III protein, interacts directly with an ESCRT-II component EAP20
Stawiecka-Mirota, M., Pokrzywa, W., Morvan, J., Zoladek, T., Haguenauer-
and regulates endosomal cargo sorting. Biochem. J. 387, 17–26.
Tsapis, R., Urban-Grimal, D., and Morsomme, P. (2007). Targeting of Sna3p
to the endosomal pathway depends on its interaction with Rsp5p and multive- Zhang, Y.Q., Qian, H.Y., Love, Z., and Barklis, E. (1998). Analysis of the
sicular body sorting on its ubiquitylation. Traffic 8, 1280–1296. assembly function of the human immunodeficiency virus type 1 gag protein
Stimpson, H.E.M., Toret, C.P., Cheng, A.T., Pauly, B.S., and Drubin, D.G. nucleocapsid domain. J. Virol. 72, 1782–1789.
(2009). Early-arriving Syp1p and Ede1p function in endocytic site placement Zimmerberg, J., and Kozlov, M.M. (2006). How proteins produce cellular
and formation in budding yeast. Mol. Biol. Cell 20, 4640–4651. membrane curvature. Nat. Rev. Mol. Cell Biol. 7, 9–19.

Cell 143, December 10, 2010 ª2010 Elsevier Inc. 887

You might also like