10
Cell, Vol. 90, 661–670, August 22, 1997, Copyright 1997 by Cell Press The Crystal Structure of Plasma Gelsolin: Implications for Actin Severing, Capping, and Nucleation Leslie D. Burtnick, ² Edward K. Koepf, ² 1991). Plasma gelsolin plays a different and less compli- cated role, forming half of a two-protein actin scaveng- Jonathan Grimes,* E. Yvonne Jones,* ing system (Haddad et al., 1990; Vasconcellos and Lind, David I. Stuart,* Paul J. McLaughlin, 1993). On the death of cells, both G-actin (monomeric and Robert C. Robinson §k actin) and F-actin (filamentous actin) are released into * Laboratory of Molecular Biophysics and the extracellular space, much ending up in blood Oxford Centre for Molecular Sciences plasma, where the ionic strength, pH, and temperature Department of Biochemistry promote the formation of elongated F-actin filaments. University of Oxford Such filaments would raise blood viscosity and interfere Oxford OX1 3QU with microcirculation. Gelsolin in the plasma acts rapidly United Kingdom to block these effects by binding to, severing, and cap- ² Department of Chemistry ping plasma F-actin to yield short gelsolin-capped University of British Columbia oligomers that have little viscometric effect. In response Vancouver, British Columbia to the low concentration of free G-actin in plasma, the Canada V6T 1Z1 gelsolin-capped oligomers lose actin monomers from Department of Biochemistry their uncapped ends. These monomers are in turn se- Edinburgh University questered by the plasma vitamin D–binding protein Edinburgh EH8 9XD (DBP) and rapidly cleansed from the circulation (Herr- United Kingdom mannsdoerfer et al., 1993). Gelsolin is one member of a family of actin-binding proteins that contain 1–6 repeats of a 120–130 amino Summary acid domain (Vandekerckhove, 1990; Hartwig and Kwiat- kowski, 1991). Gelsolin contains 6 such repeats, S1 The structure of gelsolin has been determined by crys- through S6, that appear from sequence considerations tallography and comprises six structurally related do- (Kwiatkowski et al., 1986) to have arisen from gene tripli- mains that, in a Ca 21 -free environment, pack together cation of the prototypical domain, followed by gene du- to form a compact globular structure in which the plication, since S1 and S4, S2 and S5, and S3 and S6 putative actin-binding sequences are not sufficiently are pairwise most closely related. Plasma gelsolin is a exposed to enable binding to occur. We propose that 755 amino acid protein that differs from its cytoplasmic binding Ca 21 can release the connections that join the counterpart only in possessing a 25-residue amino-ter- N- and C-terminal halves of gelsolin, enabling each minal extension (Kwiatkowski et al., 1986; Koepf et al., half to bind actin relatively independently. Domain unpublished data and GenBank accession number shifts are proposed in response to Ca 21 as bases for U31699). By alternative transcription initiation and selec- models of how gelsolin acts to sever, cap, or nucleate tive RNA processing, a single gelsolin gene in each case F-actin filaments. The structure also invites discussion is able to produce two distinct mRNA messages that of polyphosphoinositide binding to segment 2 and code both forms of gelsolin. suggests how mutation at Asp-187 could initiate a Studies of the products of limited proteolytic digestion series of events that lead to deposition of amyloid of gelsolin (Kwiatkowski et al, 1985; Bryan and Hwo, plaques, as observed in victims of familial amyloidosis 1986; Chaponnier et al, 1986) and subsequent investiga- (Finnish type). tions of recombinant individual segments of gelsolin (Way et al., 1989; Pope et al., 1991) have ascribed dis- crete functions to individual domains. S1 binds actin monomers in the absence of Ca 21 , and the resulting Introduction complex is able to cap actin filaments. S4–S6 constitute a second, Ca 21 -dependent, actin monomer–binding Gelsolin is a ubiquitous actin filament–severing, -cap- fragment that competes for the same binding site on ping, and -nucleating protein in eukaryotes. Through actin as S1. S2 contains a Ca 21 -independent polymer- these actions, intracellular gelsolin regulates the archi- binding site and is able to decorate actin filaments. Nu- tecture and motility of cells (Yin, 1987). Gene knockout cleation of actin filaments can be achieved with S2–S6, experiments have shown that while gelsolin is not essen- whereas the F-actin-severing capability is contained in tial for survival, it is required for the rapid movement of S1–S3. Dissociation of actin/gelsolin complexes can be dynamic cells such as fibroblasts, in wound healing, and induced by polyphosphoinositides binding to S2. platelets, for short blood clotting times (Witke et al., The crystal structure of human gelsolin segment S1 1995). In addition, overexpression of gelsolin in fibro- bound to monomeric actin (McLaughlin et al., 1993) re- blasts results in increased motility (Cunningham et al., veals the folding motif that is repeated in the remaining segments and details the interaction of S1 with actin. S1 is a globular domain with a central b sheet flanked § Present Address: Structural Biology Laboratory, The Salk Institute by two a helices. The longer a helix binds in a cleft in for Biological Studies, P.O. Box 85800, San Diego, California 92186- the fast-growing end of an F-actin filament. Two Ca 21 5800. k To whom correspondence should be addressed. ions are associated with this structure, one exclusively

The Crystal Structure of Plasma Gelsolin: Implications for Actin Severing, Capping, and Nucleation

Embed Size (px)

Citation preview

Cell, Vol. 90, 661–670, August 22, 1997, Copyright 1997 by Cell Press

The Crystal Structure of Plasma Gelsolin:Implications for Actin Severing, Capping,and Nucleation

Leslie D. Burtnick,† Edward K. Koepf,† 1991). Plasma gelsolin plays a different and less compli-cated role, forming half of a two-protein actin scaveng-Jonathan Grimes,* E. Yvonne Jones,*ing system (Haddad et al., 1990; Vasconcellos and Lind,David I. Stuart,* Paul J. McLaughlin,‡1993). On the death of cells, both G-actin (monomericand Robert C. Robinson§‖actin) and F-actin (filamentous actin) are released into*Laboratory of Molecular Biophysics andthe extracellular space, much ending up in bloodOxford Centre for Molecular Sciencesplasma, where the ionic strength, pH, and temperatureDepartment of Biochemistrypromote the formation of elongated F-actin filaments.University of OxfordSuch filaments would raise blood viscosity and interfereOxford OX1 3QUwith microcirculation. Gelsolin in the plasma acts rapidlyUnited Kingdomto block these effects by binding to, severing, and cap-†Department of Chemistryping plasma F-actin to yield short gelsolin-cappedUniversity of British Columbiaoligomers that have little viscometric effect. In responseVancouver, British Columbiato the low concentration of free G-actin in plasma, theCanada V6T 1Z1gelsolin-capped oligomers lose actin monomers from‡Department of Biochemistrytheir uncapped ends. These monomers are in turn se-Edinburgh Universityquestered by the plasma vitamin D–binding proteinEdinburgh EH8 9XD(DBP) and rapidly cleansed from the circulation (Herr-United Kingdommannsdoerfer et al., 1993).

Gelsolin is one member of a family of actin-bindingproteins that contain 1–6 repeats of a 120–130 amino

Summary acid domain (Vandekerckhove, 1990; Hartwig and Kwiat-kowski, 1991). Gelsolin contains 6 such repeats, S1

The structure of gelsolin has been determined by crys- through S6, that appear from sequence considerationstallography and comprises six structurally related do- (Kwiatkowski et al., 1986) to have arisen from gene tripli-mains that, in a Ca21-free environment, pack together cation of the prototypical domain, followed by gene du-to form a compact globular structure in which the plication, since S1 and S4, S2 and S5, and S3 and S6putative actin-binding sequences are not sufficiently are pairwise most closely related. Plasma gelsolin is aexposed to enable binding to occur. We propose that 755 amino acid protein that differs from its cytoplasmicbinding Ca21 can release the connections that join the counterpart only in possessing a 25-residue amino-ter-N- and C-terminal halves of gelsolin, enabling each minal extension (Kwiatkowski et al., 1986; Koepf et al.,half to bind actin relatively independently. Domain unpublished data and GenBank accession numbershifts are proposed in response to Ca21 as bases for U31699). Byalternative transcription initiation and selec-models of how gelsolin acts to sever, cap, or nucleate tive RNA processing, a single gelsolin gene in each caseF-actin filaments. The structure also invites discussion is able to produce two distinct mRNA messages thatof polyphosphoinositide binding to segment 2 and code both forms of gelsolin.suggests how mutation at Asp-187 could initiate a Studies of theproducts of limited proteolytic digestionseries of events that lead to deposition of amyloid of gelsolin (Kwiatkowski et al, 1985; Bryan and Hwo,plaques, as observed in victims of familial amyloidosis 1986; Chaponnier et al, 1986) and subsequent investiga-(Finnish type). tions of recombinant individual segments of gelsolin

(Way et al., 1989; Pope et al., 1991) have ascribed dis-crete functions to individual domains. S1 binds actinmonomers in the absence of Ca21, and the resultingIntroductioncomplex is able to cap actin filaments. S4–S6 constitutea second, Ca21-dependent, actin monomer–bindingGelsolin is a ubiquitous actin filament–severing, -cap-fragment that competes for the same binding site onping, and -nucleating protein in eukaryotes. Throughactin as S1. S2 contains a Ca21-independent polymer-these actions, intracellular gelsolin regulates the archi-binding site and is able to decorate actin filaments. Nu-tecture and motility of cells (Yin, 1987). Gene knockoutcleation of actin filaments can be achieved with S2–S6,experiments have shown that while gelsolin is not essen-whereas the F-actin-severing capability is contained intial for survival, it is required for the rapid movement ofS1–S3. Dissociation of actin/gelsolin complexes can bedynamic cells such as fibroblasts, in wound healing, andinduced by polyphosphoinositides binding to S2.platelets, for short blood clotting times (Witke et al.,

The crystal structure of human gelsolin segment S11995). In addition, overexpression of gelsolin in fibro-bound to monomeric actin (McLaughlin et al., 1993) re-blasts results in increased motility (Cunningham et al.,veals the folding motif that is repeated in the remainingsegments and details the interaction of S1 with actin.S1 is a globular domain with a central b sheet flanked§Present Address: Structural Biology Laboratory, The Salk Instituteby two a helices. The longer a helix binds in a cleft infor Biological Studies, P.O. Box 85800, San Diego, California 92186-the fast-growing end of an F-actin filament. Two Ca215800.

‖To whom correspondence should be addressed. ions are associated with this structure, one exclusively

Cell662

bound to gelsolin and another bound at the actin/gel- A2. The model consists of residues 27–755; there wasno interpretable electron density for the N-terminal 26solin interface.

Despite extensive functional analysis of gelsolin and amino acids. Electron density is good for the mainchainthroughout the protein. Mainchain temperature factorsstructural analysis of single domains of gelsolin-like pro-

teins, very little is knownof the structural basis for modu- rise above 60 A2 in a number of regions: residues 27–28,94–97, 152–158, 256–268, 276–285, 290–291, 323–326,lation of F-actin architectures by whole gelsolin, or of

how this process is regulated. Here, we present the 367–368, 370–377, 455–458, 521–547, 554–634, 646–652, 686–695 and 751–755. The only Ramachandran plotcrystal structure of Ca21-free, whole horse plasma gel-

solin. This structure reveals the organization of the six outliers are ten residues from these ill-defined regions(28, 266, 280, 531, 532, 535, 547, 631, 632, and 677).related domains and allows discussion of the functional

implications of the arrangement in relation to the actin-severing, -capping, and -nucleating activities, as well Structural Description

The six domains of gelsolin share a similar folding topol-as for other reported behaviors of gelsolin and gelsolinfragments. ogy, consisting of a central five- or six-stranded b sheet

sandwiched between a 3.5-to-4.5 turn a helix (helix 1)running roughly parallel to the strands in the sheet, andResults and Discussiona 1-to-2 turn a helix (helix 2) running approximately per-pendicular to the strands (Figures 1A and 2). S1 is clearlyCrystallographic Structure Determination

Details of the structure determination are presented most closely related to S4 (90 Ca equivalences with rmsdistance 0.97 A), as is S3 to S6 (86 Ca equivalencesin the Experimental Procedures. The present gelsolin

model consists of two noncrystallographically related with rms distance 0.80 A). S5 is the outlier, and while itis most closely related to S2 (88 Ca equivalences withmolecules and includes 11,296 nonhydrogen atoms and

267 ordered water molecules. The model has an R value rms distance 1.82 A), the similarity is much less pro-nounced. Such structural homology was somewhat an-of 18.7%, and the free R factor is 23.0% for all 71,395

reflections measured in the range 20.0–2.5 A, with an ticipated from the sequence similarity. The definition ofthe domains based on the structure is S1 (Pro-39 torms Dbonds of 0.011 A and an rms Dangles of 1.628 (symbols

as defined by X-PLOR [Brunger, 1992]). No significant Tyr-133), S2 (Gly-137 to Leu-247), S3 (Ala-271 to Gln-364), S4 (Gln-419 to Leu-511), S5 (Pro-516 to Leu-618),structural differences were observed between the two

noncrystallographically related molecules. The mean and S6 (Pro-640 to Gly-731). The basic topology of thecommon five-stranded sheet from each domain is fourtemperature factor is 45.0 A2 for molecule 1 and 43.4 A2

for molecule 2, with isotropic B factors restrained such antiparallel strands (labeled A–D, Figure 2), with strandE being parallel to D. The C and D strands in S1 and S4that the rms DBbonds 5 5.36 A2 and the rms DBangles 5 8.25

Figure 1. The Structure of Gelsolin

(A) Schematic representations of the six domains of gelsolin in similar orientations. A standard coloring of the segments is used here and insubsequent figures: S1, red; S2, light green; S3, yellow; S4, pink; S5, dark green; and S6, orange.(B) The relative organization of S1–S3 and S4–S6. The schematic representation shows S1–S3 (upper panel) and S4–S6 (lower panel) inapproximately the same orientation. Linker regions are striped with the colors of the linked segments.(C) Schematic representations of the structure of whole gelsolin. The upper panel shows a view rotated by 908 clockwise around the horizontalwith respect to the lower panel. The C-terminal extension is striped orange and black.

The Crystal Structure of Plasma Gelsolin663

of S2 in comparison to S1 and S3 is noticeably differentfrom that of S5 relative to S4 and S6 (Figure 1B).

In whole gelsolin (Figure 1C), there are N-terminal andC-terminal extensions (although the portion of theN-terminal region characteristic of plasma gelsolin isnot visible), and a highly convoluted 53 amino acid linkerconnects S1–S3 to S4–S6. The segments pack closelytogether to give a globular shape, overall dimensions85 A 3 55 A 3 36 A, with S6 constituting the centraldomain of the molecule. Indeed, S6 is the only domainfrom S4–S6 to form direct contacts to S1–S3. The longhelices (helix 1) from S1, S3, S4, and S6 are exposedon one side of the protein, while the corresponding S2and S5 helices are on the opposite side. In the viewpresented in the top panel of Figure 1C, it can be seenthat S1–S3 is pseudosymmetric with S4–S6 related bya rotation of 1808 around an axis through the centerof the molecule and perpendicular to the paper. TheC-terminal 24 amino acids contain a stretch of helix,residues 745–754, that interacts with S2, lying approxi-mately aligned with the S2 long helix.

Structural and Sequence ComparisonsFigure 2. The Topology of Gelsolin The structure of horse plasma gelsolin S1, determinedThe topology of gelsolin. b strands are depicted as arrows and here within the context of whole gelsolin, when com-helices as rectangles. pared to that of isolated, recombinant human gelsolin

S1, complexedto actin and Ca21, shows extremely closesimilarity (108 Ca equivalences, rms distance of 0.64are extended in comparison to the other segments, withA). The domain fold is apparently rigid and structurallythe C-terminal end of the C strand and the N-terminalinsensitive to the binding of actin, Ca21, or the other fiveend of the D strand displaying a distinctive kink in S4domains of gelsolin. This confirms reports from NMRrelative to S1. Both S1 and S4 contain an extra smallstudies of isolated domains of villin (Markus et al., 1994)strand, labeled C9 (Figures 1B and 2), which is separateand severin (Schnuchel et al., 1995), that the binding offrom the main b sheet. S2 and S5 have a six-strand bCa21, and possibly actin, do not cause rearrangementsheet topology, with the extra N-terminal strand (labeledof the segment fold of these proteins.A9 in Figure 2) in a parallel configuration with strand B.

An overlay of the structure of gelsolin S1 onto theAccordingly, these two segments are slightly larger thanstructure of the analogous, isolated domain of villinthe other segments (111 and 103 residues, respectively,shows their considerable similarity (97 Ca equivalences,as opposed to 92–95 residues). The long helix (helix 1)rms distance of 1.85 A). Comparison of gelsolin S2 withof S5 is shorter than in the other segments, being 3.5the equivalent domain from severin (81 Ca equivalences,turns as opposed to 4.5 turns in length, and the longrms distance of 1.91 A) again confirms the sequence-helices in S3 and S6 display a distinctive kink approxi-based predictions for structural resemblance. In eachmately one third of the way along from their N-terminalcase, the domains share similar dimensions and an iden-ends. S2 contains a disulfide bond between the C andtical topology. The sheet and helical elements overlayD strands (Cys-188 and Cys-201, Figure 4B), typical ofreasonably well, with observed variations occurring inplasma but not cytoplasmic gelsolins (Wen et al., 1996).loop conformations.The domains cluster into two sets, S1–S3 and S4–S6,

The high degree of sequence identity between gel-with each triplet of domains showing broadly similarsolin, villin, and severin, particularly in regions shownpseudoquaternary structure (Figure 1B). Superpositionto be buried or to form secondary structure in the caseof S1–S3 onto S4–S6 reveals 223 Ca equivalences withof gelsolin, suggests that the domain folds in gelsolinan rms distance of 1.91 A. S1 and S3, and similarly, S4may be used to predict the geometries for the as yetand S6, form continuous ten-stranded b sheets, the twounsolved structures of the corresponding domains infive-strand sheets in each case interacting in an antipar-these other members of the gelsolin family of proteins.allel fashion though their E strands. The kinks in the longIn addition, these proteins display a significant amounthelices (helix 1) of S3 and S6 avoid collision of thoseof sequence identity in their interdomain regions (Figurehelices with the corresponding ones from S1 and S4,3). They are likely to display similar pseudoquaternaryrespectively. S1 and S4 contribute a single strand (C9)structures in line with their similarities in function.parallel to A9 in the sheets of S2 and S5, respectively,

forming continuous seven-stranded b sheets. The longloop between S5 and S6 (22 amino acids) is more convo- Actin- and Polyphosphoinositide-Binding Sites

The G-actin-binding determinant of S1 has been re-luted than the S2–S3 loop. When S1–S3 and S4–S6 arecompared, the organization of S1 relative to S3, and S4 vealed by the X-ray structure of the gelsolin-S1/G-actin

complex (McLaughlin et al., 1993). In contrast, little isrelative to S6 is very similar. However, the relationship

Cell664

Figure 3. Sequence Comparisons of Gelsolin

A sequence alignment of gelsolin, villin, and severin. The sequences of horse gelsolin (e_gel; Koepf et al., unpublished data and GenBankaccession number U31699), human gelsolin (h_gel; Kwiatkowski et al., 1986), human villin (h_vil; Arpin et al., 1988), and Dictyostelium discoideumseverin (d_sev; Andre et al., 1988) are aligned with respect to sequence and structural considerations. Conserved residues are colored yellow.Secondary structure is shown as arrows (strands) and cylinders (helices) below the alignment, colored according to segment (Figure 1A) andlabeled according to topology (Figure 1B). Black secondary structure denotes structure that falls outside the domains. ($) and (#) denoteresidues in the horse gelsolin structure with less than 10 A2 or 20 A2 solvent-accessible surface area, respectively. The numbering is that ofhorse plasma gelsolin.

known of the residues of S4 that bind G-actin (Pope PPI-binding sites have been identified (Yu et al., 1992)to lie within sections 132–149 (yellow) and 161–172 (or-et al., 1995). It has been suggested that due to the

sequence homology between S1 and S4, these two ange, Figure 4A) of the gelsolin sequence. These resi-dues map to the flat, solvent-exposed surface of S2, indomains will adopt similar G-actin-binding modes.

Structural superposition of S4 onto S1 reveals that of this case centered on the A9 and A strands, respectively.Within these two regions, arginine and lysine residuesthe 11 S1 residues forming the S1/G-actin interface

(McLaughlin et al., 1993), 6 are structurally conserved cluster to form potential phosphate-binding sites (Figure4A). The simplest interpretation of these observations(Q-95 is equivalent to Q-473; I-103 to I-481; Q-107 to

Q-485; D-110 to D-488; Q-118 to Q-496; and R-120 to is that PPI binding dislodges actin by binding to anoverlapping region of the surface of S2. A more radicalR-498), one retains similar character (V-106 to A-484),

another can provide a hydrogen-bonding oxygen (D-50 suggestion, based on experiments involving a peptide(residues 154–166), has come from Xian et al. (1995),to N-429), while two positions overlay residues of differ-

ent character (F-49 to S-428 and D-84 to I-462) and the who suggest that PPI binding may induce these residues(which include strand A of S2) to adopt a helical configu-final residue, 149, lies in a structurally different region

in whole gelsolin when compared to S1 complexed to ration. Such a significant disruption of the structure ofthe core of S2 would also be consistent with loss ofactin. Hence, S1 and S4 display a similar surface charac-

ter in the region of S1 demonstrated to bind actin, F-actin-binding capability. While consideration of thelikely structural origins of Familial Amyloidosis (Finnishstrengthening the notion that S4 will bind actin in an

analogous manner to S1. type) provides support that strand A of S2 may be dis-placed relatively easily from the b sheet (see below), theThe F-actin-binding residues on plasma gelsolin lie

within residues 161–172 (orange, Figure 4A; Sun et al., peptide studied by Xian et al. (1995) contains only partof the A9-A loop and the A strand in isolation. Devoid of1994), which comprise strand A and the AB loop in S2,

and within residues 197–226 (the long helix of S2, green, the stabilizing hydrogen bonds to the other strands ofthe b sheet, this peptide would be unlikely to adopt anFigure 4A; Van Troys et al., 1996). Furthermore, residues

within residues 150–160 in the C-terminal portion of appropriate conformation. Hence, competitive bindingmay provide a simpler mechanism for PPI disruption ofthe A9A loop of S2 have been implicated in severing

(Kwiatkowski et al., 1989; shown in light blue in Figure gelsolin/actin complexes.4A). The structure reveals that these regions in S2 consti-tute a single, solvent-exposed entity that is central to Implications for the Molecular Basis of Familial

Amyloidosis (Finnish Type)the interactions of gelsolin with F-actin.The severing and capping properties of gelsolin are A single mutation in gelsolin (Maury et al., 1990) leads

to the deposition of gelsolin fragments as amyloidinhibited by binding polyphosphoinositides (PPIs). Two

The Crystal Structure of Plasma Gelsolin665

native protein, the destabilizing effect of such a changemay be expected to be focused on strand A and the BCloop, allowing access of a trypsin-like protease to thepotential cleavage site between Arg-172 and Ala-173 atthe beginning of strand B. The disruption of the nativeconformation of strand A is evinced by the loss of actin-binding, -severing, and -nucleating activities in the mu-tant protein (Weeds et al, 1993), a dysfunction in linewith the putative role of this region (specifically residues161–172) in F-actin binding (Sun et al., 1994).

The regions in Figure 4B colored red and purple (Ala-173 - Met-243; four strands and one helix) comprise theamyloid-forming fragment, while residues colored onlypurple (Ala-173–Gly-202, three strands) comprise a syn-thetic gelsolin fragment with a greater propensity to formamyloid fibrils than the familial amyloidosis fragment(Weeds et al., 1993). The structure of the synthetic gel-solin fragment includes the structurally robust, antipar-allel b sheet formed from the C and D strands that isstabilized by a disulfide bond, from Cys-188 to Cys-201,found in the plasma gelsolin structure. Such a scenariois consistent with the hypothesis that there is a genericstructure for amyloid fibrils that are formed through poly-merization of a stable b sheet subunit to produce anextended b sheet helix whose axis is parallel to theprotofilament of the amyloid fibril (Blake and Serpell,1996).

Ca21-Free Gelsolin Does Not Bind ActinModels of an F-actin filament, based on the Heidelbergmodel (Holmes et al., 1990), were generated in which theactin/S1 structure (McLaughlin et al., 1993) was pastedFigure 4. Regions of Gelsolin Important in F-Actin and PIP Bindingonto F-actin to superimpose actins. The whole gelsolinand Implicated in FAFstructure then was overlaid onto the model such that(A) The figure shows the S1–S2 fragment of gelsolin (residues 39–

248). S1 is colored red. The yellow region denotes residues 132–149, the S1 subunits were superimposed (Figure 5). This pro-the cyan region residues 150–160, and the orange region residues duces a picture of how gelsolin may cap the fast-grow-161–172. The remainder of S2 is colored green. The cyan and orange ing end of an actin filament. However, the structure ofregions as well as the long S2 helix (green) have been implicated gelsolin presented here is devoid of bound Ca21, a statein F-actin binding. The yellow and orange regions have been impli-

in which it is unable to bind actin. Further inspection ofcated in PIP binding. Potential phosphate coordinating residues inthe gelsolin/F-actin model suggests a number of rea-these regions are also drawn.

(B) Features of the S2 segment of gelsolin important in familial sons why this Ca21-free form of gelsolin does not bindamyloidosis. Residues Asn-184, Lys-166, and Gln-164 are within actin. Firstly, there is significant steric conflict betweenbonding distance to the amyloidosis mutation site Asp-187. The S3 (yellow) and actin, which would prevent S1 reachingamyloid-forming fragment (173–243) is colored both red and purple, its actin-binding site. Secondly, S2 (light green), whichwhile the remainder of S2 is colored cyan. A synthetic peptide con-

contains theF-actin-binding capability (Sun et al., 1994),sisting of residues 173–202 (purple) is also capable of forming amy-is not in contact with actin. Finally, while S4, S5, andloid fibrils. The disulphide bond from Cys-188 and Cys-201 falls in

this region and may help to stabilize the amyloid fragment. S6 are in striking distance of the second actin in thefilament, S4 (pink), which contains the third actin-bind-ing site (Pope et al., 1995), does not actually contact

plaques in Finnish type familial amyloidosis (FAF), anactin. It seems that there must be large shifts in the

inherited disease typified by hyperelastic skin and amy-relative positions of the segments to overcome these

loid deposits in the cornea and cranial nerve, which leadsteric problems and to permit tight interaction with the

to facial muscle weakness. The amyloidosis is knownactin filament.

to arise as a direct result of the mutation of Asp-187 toeither asparagine or tyrosine. This change renders anotherwise masked cleavage site (between residues Arg- Ca21 Activation of Gelsolin

Crystallization of G-actin/S1 identifies two Ca21-binding172 and Ala-173) sensitive to a trypsin-like protease.In the structure of S2, Asp-187 (the final residue of sites on S1 (McLaughlin et al., 1993). The lower affinity

site, which may not be physiological (Weeds et al.,1995),the BC loop, Figure 4B) is within hydrogen bonding dis-tance of three residues, Gln-164 (3.26 A), Lys-166 (2.97 comprises residues 109, 114, and 116, with actin provid-

ing one ligand, Glu-167, to complete the coordination.A), and Asn-184 (2.96 A). These interactions contributeto the stability of the S2 b sheet and the BC loop. When This Ca21 site, when examined in whole gelsolin, issand-

wiched between the common b sheet and the two longAsp-187 is replaced by asparagine or tyrosine, this net-work of interactions is lost. From the structure of the helices of S1 and S3 (Figure 6A). In whole gelsolin,

Cell666

Figure 5. A Model for Actin CappingFigure 6. Ca21 Control of Gelsolin FunctionWhole gelsolin is shown positioned on F-actin (Holmes et al., 1990)

according to the structure of segment 1/G-actin (McLaughlin et al., (A) The lower affinity human S1-actin Ca21-binding site. S1 is shownin red and actin in gray (McLaughlin et al., 1993). The calcium-1993), such that the S1 subunits are superimposed. The segments

of gelsolin in schematic representation are colored as in Figure 1. binding residues 109, 114, and 116 from S1 (blue and green) andGlu-167 (gray) from actin are detailed. The position of S3 relative toFour protomers of an actin filament are shown in different shades

of gray and are shown in a space-filling representation. The residues S1 in whole gelsolin is shown in yellow. In this arrangement, S3would sterically clash with actin, and in particular, Lys-319 from S3on F-actin that bind gelsolin S2 are shown in dark blue and those

that bind a-actinin in light blue. occupies a position similar to that of Glu-167 from actin.(B) The high affinity human S1 Ca21-binding site. S1 is shown in red,the C-terminal portion of the S1 (1Ca21; McLaughlin et al., 1993)structure is shown in pink wrapping around the Ca21 ion (purple).Lys-319 from S3 occupies the same space that wouldThe structure of this region, which lies in S2, (2Ca21) from wholebe filled by residue Glu-167 from actin in the G-actin/gelsolin (light green) shows the path of these amino acids to be

S1 structure (Figure 6A). While lysines tend to be flexible, quite different, with the divergence being close to the Ca21-bindingit is tempting to propose that if this Ca21 is physiologi- site. The mainchain carbonyl group of Val-145 displays 1808 reorien-cally relevant, then the incompatibility of lysine with Ca21 tation, allowing it to coordinate the calcium, setting the path for the

continuation of the chain.results, not in the reorientation of the lysine sidechain,but in sufficient destablization of the orientation of S3with respect to S1 to induce S3 out of steric conflictwith actin (Figure 5) and allow interaction between actin site could be created in the whole gelsolin molecule by

a rather modest local rearrangement of the structureand S1.The second, higher affinity, Ca21 site (residues 35, 65, resulting in the rerouting of the mainchain. While the

differences in the conformation of residues 145–148,66, 97, and 145) observed in the G-actin/S1 complexwould be located between S1 and S2 in whole gelsolin. between Ca21-bound isolated S1 and Ca21-deficient

whole gelsolin, may simply be due to the fragment S1There is considerable evidence that this site modulatesgelsolin/actin interactions, resulting in this Ca21 becom- being out of its proper molecular context, it is, however,

intriguing that the adjacent residues, 150–160, haveing trapped on binding whole gelsolin to F-actin (Weedset al., 1993, 1995). Residues 145–148 at the C terminus been implicated in severing (Way et al., 1992). Binding of

this calcium ion may therefore facilitate a rearrangementof S1 take a very different path in isolated S1 (McLaugh-lin et al., 1993) than in whole gelsolin (Figure 6B), where, within S1–S3 required for F-actin severing, consistent

with observations that, while S1–S3 severing isnot underby our structural definition of the domains, they in factform part of domain S2. The major effect on the Ca21- absolute Ca21 control, the efficiency of severing is in-

creased in its presence.binding site is that in whole gelsolin the mainchain car-bonyl oxygen of Val-145 (a calcium ligand in the G-actin/ While the isolated half-gelsolin fragment, S1–S3, is

able to sever actin filaments in the absence of Ca21,S1 complex) is reoriented away from the Ca21 positionby a 1808 flip of the peptide. Thus the calcium-binding severing by whole gelsolin isabsolutely Ca21 dependent,

The Crystal Structure of Plasma Gelsolin667

implying an additional mechanism for the calcium con- Weeds and Maciver, 1993). Presumably, the first eventin severing must be the binding of S2 to the side of thetrol of gelsolin quite distinct from that discussed above,filament, followed by a second binding event, involvingresiding somewhere in S4–S6,or at the interface of theseS1 and actin, which destabilizes the adjacent actin–actinsegments with S1–S3 (Pope et al., 1991). Furthermore,contact and leads to filament severing.the conformational change in gelsolin induced by Ca21

The F-actin-binding residues, 161–172 and 197–226has been localized to the C-terminal half of the moleculeon plasma gelsolin (Sun et al., 1994) and the residuesby light scattering experiments (Hellweg et al., 1993).that have been implicated in severing (residues 150–160,S5 and S6 have been shown to contribute to twoWay et al., 1992), are all located within S2 (Figure 4A).separate Ca21-binding sites (Pope et al., 1995), with theActin residues 1–10 and 18–28 comprise the known por-Ca21 site of S6 being a high affinity site. S6 is centrallytion of the S2-binding site on F-actin (Feinberg et al.,placed in the structure of gelsolin, having multiple inter-1995). In addition, competition between isolated gelsolinactions with the first three segments, including ap-fragment S2–S3 and an a-actinin fragment, aA12, forproaches within 4 A of residues Arg-168 and Arg-169 inbinding to F-actin (Way et al., 1992; reviewed in Sheterlinthe presumed F-actin-binding site of S2 (Sun et al.,and Sparrow, 1994) suggests a role for actin residues1994). Ca21 control of F-actin severing by gelsolin is lostwithin 108–189 in binding S2. When the gelsolin struc-on deletion of the 23 C-terminal amino acids of gelsolinture is positioned on the actin filament in accordance(Kwiatkowski et al., 1989). These residues lie within thewith the G-actin/S1 structure, S2 does not contactC-terminal tail of the molecule and are unlikely to alterF-actin at all, being well removed from its proposedthe fold of S6. This C-terminal region contains a helixbinding sites (Figure 5). A large reorientation of S1 withthat is in intimate contact with the long helix of S2. Werespect to S2 would be necessary to allow S1 and S2suggest that, on binding Ca21 in the second half of theto interact simultaneously with actin through these ex-molecule, most probably in S6, there is a major re-pected binding sites. The C-terminal portion of S1 con-arrangement of the C-terminal helix relative to S2,tinues into the N-terminal of S2 without a significantexposing the F-actin-binding site on S2 and releasinglinker sequence, such as that observed between S2 andan important structural constraint on the relative orienta-S3 (Figure 1B). Barring considerable unfolding of thetion of the two halves of the gelsolin molecule.termini of these two segments, it would not be possiblefor S2 to reach along the filament axis to contact the

A Model for Actin Filament Nucleation outward-facing sites on an adjacent actin to which itby Whole Gelsolin is expected to bind in the model for filament severingFragment S2–S6 of gelsolin can function, like whole gel- proposed by McGough and Way (1995). Rather, dimen-solin, to nucleate actin polymerization through binding sions of the domains would permit reorientation, withtwo actin monomers in an appropriate orientation (Hes- the S1–S2 interface as a pivot, so that S2 could contact

both of the putative binding sites on the same actin unitterkamp et al., 1993). As discussed previously, S4 hasto which S1 is joined.been predicted to interact with actin in a manner similar

The structure of gelsolin does show that although S2to S1. For an interaction mode based on that of theis linked to S1 by a b sheet, it is relatively isolated fromcrystalline G-actin/S1 complex (Figure 5), S4 is close toS1 and S3, which themselves are in intimate contacta second actin but distant from the proposed interactionthrough a continuous b sheet (Figure 1B).The C terminussite. If the second half of the gelsolin molecule is reposi-of S1 immediately enters the N terminus of S2, whiletioned, docking it onto an actin across the filament axis,there is a long and convoluted loop between S2 andin an analogous mode to that of the first half, then aS3. This provides the possibility of a pivotal movementcap can be constructed that would present two actinbetween S1 and S2 that would be able to retain themonomers in proper alignment to act as a nucleus forrelationship between S1 and S3 by taking up the slackpolymerization. This action would not require thein the S2–S3 linker. While there is no direct evidencestretching out of gelsolin domains S1 through S6 to wrapfor such a movement, comparison of S1–S3 with thearound the filament; instead, the repositioning of thehomologous S4–S6 reveals that the relationship be-two halves of the molecule, as rigid bodies, may between S1 and S3 is very similar to that between S4accommodated by the polypeptide linker, of some 50and S6 (Figure 1B). The positioning of S2 relative to S1residues, between S3 and S4.and S3 is noticeably different to that of S5 relative toS4 and S6, which suggests that some mobility of S2

A Model for Actin Severing and Capping by S1–S3 relative to S1 and S3 may be possible, and indeed ofIsolated S1 can bind an actin monomer and cap but not S5 relative to S4 and S6. Furthermore, the high affinitysever an actin filament (Pope et al., 1991). The structure Ca21-binding site in S1–S2, discussed above, producesof G-actin/S1, when overlaid on the Heidelberg model a 1808 reorientation of the N-terminal portion of S2 con-of the actin filament (Holmes et al., 1990), shows steric sistent with this mechanism.clashes with the actin below which (toward the fast- S2–S3, which does not possess actin-severing activ-growing end) S1 is bound, consistent with the ability of ity, can both bind to the side of actin filaments and capS1 to prevent actin polymerization when bound to the the ends of actin filaments (Sun et al., 1994). This pointsend of a filament, but not to bind directly to and sever toward S3, in the S2–S3-capped filament, occupying anF-actin filaments (McLaughlin et al., 1993). Gelsolin frag- orientation that would cause steric conflict with an actinments containing only S1–S2 or S1–S3 are capable of below the one to which S2 is bound (at the fast-growingsevering and capping actin filaments in the absence end of the filament). Furthermore, while S1–S2 is suffi-

cient for severing, S3 is additionally required for strongof Ca21, but display Ca21-dependent rates (reviewed in

Cell668

Table 1. Summary of Data Collection and Phasing Statistics for Gelsolin Native and Heavy Atom Derivative Data

Data Set Resolution (A) Percentage of data Rmerge (%) MFID (%) Number of Sites Phasing Power Nonzero Anomalous (%)

Native 20.0–2.5 92 0.077 — — — —EMP 15.0–3.5 99 0.074 20.0 6 2.6 —EMP* 15.0–3.1 83 0.092 16.6 6 3.5 76EMP 1 DTT 15.0–4.0 96 0.132 27.8 10 1.6 —EMP 1 DTT* 15.0–3.5 84 0.104 20.6 10 1.7 59Baker’s 15.0–4.0 95 0.088 26.6 6 1.4 —

MFID, mean fractional isomorphous difference; EMP, ethylmercuric phosphate; DTT, dithiothreitol; Baker’s, Baker’s dimercurial, mercurymeso(2,3 dimethoxytetramethylene)bis(acetate), Rmerge (S|I 2 ,I.|/S,I.).*Denotes data collected at a synchrotron source.

F-actin capping (Sun et al., 1994). This is consistent with that is capable of binding to one actin at the end of afilament. The binding of S2 to the side of F-actin facili-maintenance of the close relationship between S1 and

S3 in the structure of gelsolin throughout the severing tates insertion of S1 between the S2-binding unit andan adjacent one to achieve S1 docking and effect sev-process, albeit with a slight rearrangement of the two

domains relative to one another. ering. The 50-residue tether from S3 to S4 permits thesecondhalf of gelsolin toorient appropriately for interac-Although details of the association of gelsolin with

actin remain obscure, the present results do allow us tion with the second actin unit at the newly exposedfilament end, consistent with the two actin monomersto suggest a schematic model for the severing of F-actin

by the whole gelsolin in the presence of calcium. The being flexible with respect to each other (Hesterkampet al., 1993). The final image represents the gelsolin-suggested sequence of events is presented in Figure 7.

We first indicate the approximate arrangement of S1 capped end of the severed filament.through S6 in the Ca21-free protein and how the binding

Experimental Proceduresof Ca21 might open up that structure to expose theF-actin-binding site on S2. From that point on, as little

Crystallizationis known about the effect of Ca21 on the structure of theGelsolin was purified from horse plasma according to previously

second half of gelsolin, beyond that its hydrodynamic reported protocols (Reid et al 1993). Crystals (1.0 3 0.5 3 0.5 mm3)volume increases significantly (Hellweg et al, 1993), we were grown by vapor diffusion from drops containing a 1:1 ratio ofrepresent that half of gelsolin as a singleellipsoidal entity 34% saturated ammonium sulfate solution, 100 mM Tris–HCl (pH

8.0) reservoir solution, and 10 mg/ml protein solution. The crystalsproved to be unstable, often dissolving in the capillary tube duringX-ray diffraction data collection. Presoaking the crystals in 60%saturated ammonium sulfate, 100 mM Tris–HCl (pH 8.0), prior tocrystal mounting, stabilized the crystals with no apparent loss indiffraction quality. The space group for the gelsolin crystals is P4212with cell dimensions a 5 b 5 169.4 A, c 5 154.2 A, a 5 b 5 g 5

908. The asymmetric unit contains two molecules and approximately55% solvent.

Diffraction Measurements and PhasingNative data (to a resolution limit of 2.5 A resolution) from threecrystals, and definitive heavy atom derivative sets were collectedat 188C using synchrotron radiation at the Synchrotron RadiationSource, Daresbury, UK (Station 9.6, 0.882 A wavelength beam lim-ited by 0.3 mm slits) using a 30 cm MAR-research imaging plate.Potential heavy atom derivatives were screened on an 18 cm MAR-research imaging plate mounted on a Rigaku RU200 rotating anode(CuKa X-rays) with a graphite monochromator. The data were au-toindexed, integrated, and corrected for Lorentz and polarizationeffects using DENZO (Otwinowski, 1993) and scaled and mergedwith SCALEPACK (Otwinowski, 1993).

Figure 7. A SchematicModel of Ca21-Controlled Severing of F-Actin Chemical modification studies of horse gelsolin had shown threeby Gelsolin of the five cysteine thiols to be reactive (Reid et al., 1993). Accord-

ingly, the thiol-specific reagents ethylmercuric phosphate (EMP) andThe top panel shows the activation of gelsolin by Ca21. Ca21 bindingto S4–S6 disrupts the contact of S2 with S6 and the C terminus, Baker’s mercurial were tested as heavy atom reagents to modify

the crystalline gelsolin. Reaction with EMP was carried out both withexposing the F-actin-binding site on S2, while S4–S6 undergoesstructural rearrangement. The central panel depicts S2 from Ca21- and without pretreatment of the gelsolin crystals with dithiothreitol

(DTT) in an attempt to expose the previously unreactive thiols. Usefulactivated gelsolin bound to F-actin. Secondly, S1 and S3 swingdown to allow S1 to disrupt the actin–actin contact, resulting in S1 derivatives resulted from soaking crystals (24 hr) in the following

conditions: EMP, 5 mM; EMP, 5 mM after pretreating the crystaland S3 being bound below the same actin to which S2 is bound.Lastly, S4–S6 swing over to bind a second actin to allow S4–S6 to with 10 mM DTT; and 5 mM Baker’s mercurial. Difference Patterson

maps were interpreted using SHELX (Sheldrick et al., 1993). Theinteract with the second actin in an analogous fashion to S1–S3.The bottom panel provides the key: S1 and S3, and similarly S4 and EMP product of the reaction without pretreatment with DTT was

the first to be identified, with six binding sites, three per moleculeS6 are shown as half ellipses to emphasize the common b sheetbetween these domains. (C) denotes the C terminus; (S4–S6) de- in the asymmetric unit (Cys-93, Cys-304, and Cys-645). Difference

Fourier analysis revealed a further two sites per molecule for thenotes Ca21-activated S4–S6; and (A) denotes an actin protomer.

The Crystal Structure of Plasma Gelsolin669

DTT-pretreated EMP derivative (Cys-188 and His-638). All the Laboratory, USA. Prerelease coordinates are available from R. C. R.(e-mail address: [email protected]).data detailed in Table 1 were included in the phase calculation

(MLPHARE; Otwinowski, 1991). The mean figures of merit of thephases prior to solvent flattening, 0.609 for centric reflections and Received April 28, 1997; revised June 30, 1997.0.818 for acentric, were inflated (especially for the acentric reflec-tions) due to the common heavy atom sites.

ReferencesThe program GAP (J. Grimes and D.I. Stuart, unpublished) wasemployed to improve phases by solvent flattening and molecular

Andre, E., Lottspeich, F., Schleicher, M., and Noegel, A. (1988).averaging. After solvent flattening, using an estimated solvent con-Severin, gelsolin, and villin share a homologous sequence in regionstent of 55%, separate envelopes for the two molecules were deter-presumed to contain F-actin severing domains. J. Biol. Chem. 263,mined by combining the solvent envelope with those of 12 sets of722–728.human gelsolin S1 coordinates manually positioned in the solvent

flattened electron density map. A noncrystallographic symmetry re- Arpin, M., Pringault, E., Finidori, J., Garcia, A., Jeltsch, J.,and Vande-lationship was calculated from the 10 heavy atom binding sites (5 kerckhove, J. (1988). Sequence of human villin: a large duplicatedper molecule) and cycles of molecular averaging, solvent flattening domain homologous with other actin-severing proteins and a uniqueand phase combination performed. The identities of the six domains small carboxy-terminal domain related to villin specificity. J. Cellof gelsolin, in each of the two molecules, were distinguishable in Biol. 107, 1759–1766.the averaged, solvent flattened, 20.0–3.1 A MIRAS (multiple isomor- Barton, G.J. (1993). ALSCRIPT: a tool to format multiple sequencephous replacement with anomalous scattering) map. Each segment alignments. Prot. Eng. 6, 37–40.of molecule 1 was rebuilt, inserting the correct amino acids, starting

Blake, C., and Serpell, L. (1996). Synchrotron X-ray studies suggestfrom the structure of gelsolin segment 1. Molecule 2 was generatedthat the core of the transthyretin amyloid fibril is a continuous bby application of the noncrystallographic symmetry operator. 4%sheet helix. Structure 4, 989–998.of the structure factors were reserved for the free R-factor calcula-Bryan, J., and Hwo, S. (1986). Definition of an amino-terminal actin-tion that was used to judge that domains 1–2, 3–4, 5 and 6 frombinding domain and a carboxyl-terminal Ca21-regulatory domain ineach molecule should be treated as separate rigid bodies in rigidhuman brevin. J. Cell Biol. 102, 1439–1446.body refinement. Noncrystallographic symmetry operators to relate

independently these portions of the two molecules were then calcu- Brunger, A.T. (1992). X-PLOR. Version 3.1. A system for X-ray crys-tallography and NMR. (New Haven, CT: Yale University Press).lated and fed back into the original MIRAS map molecular averaging

process. This resulted in an improved electron density map from Chaponnier, C., Janmey, P.A., and Yin, H.L. (1986). The actin fila-which the structures of the interdomain linker regions could be ment–severing domain of gelsolin. J. Cell Biol. 103, 1473–1481.established. Cunningham, C.C., Stossel, T.P., and Kwiatkowski, D.J. (1991). En-

hanced mobility in NIH 3T3 fibroblasts that overexpress gelsolin.Science 251, 1233–1236.RefinementFeinberg, J., Benyamin, Y., and Roustan, C. (1995). Definition of anThe resulting model (consisting of residues 20–556, 558–620, 629–interface implicated ingelsolin binding to the sidesof actin filaments.713 and 715–755) was refined (positional, simulated annealing, andBiochem. Biophys. Res. Commun. 209, 426–432.individual restrained B-factor refinement, XPLOR; Brunger, 1992) in

the range 10.0–2.5 A, keeping noncrystallographic symmetry con- Haddad, J.G., Harper, K.D., Guoth, M., Pietra, G.G., and Sanger,straints between the two copies of domains 1–2, 3–4, 5, and 6. After S.W. (1990). Angiopathic consequences of saturating the plasmaseveral further rounds of refinement and manual rebuilding, and scavenger system for actin. Proc. Natl. Acad. Sci. USA 87, 1381–inclusion of a bulk solvent correction, the R factor and free R factor 1385.were 21.2% and 26.1%, respectively, in the range 20.0–2.5 A. In the Hartwig, J.H., and Kwiatkowski, D.J. (1991). Actin binding proteins.final round of refinement, the noncrystallographic symmetry con- Curr. Opin. Cell. Biol. 3, 87–97.straints were replaced by noncrystallographic symmetry restraints,

Hellweg, T., Hinssen, H., and Eimer, W. (1993). The Ca21-inducedand 267 ordered waters were added.conformational change of gelsolin is located in the carboxyl-terminalhalf of the molecule. Biophys. J. 65, 799–805.

Herrmannsdoerfer, A.J., Heeb, P.J., Feustel, P.J., Estes, J.E., Kee-AnalysisStructural superpositions were performed using the program SHP nan, C.J., Minnear, F.L., Selden, L., Giunta, C., Flor, J.R., and Blu-

menstock, F.A. (1993). Vascular clearance and organ uptake of(Stuart et al., 1979), and secondary structure assignments werecalculated using the program DSSP (Kabsch and Sander, 1983). G- and F-actin in the rat. Amer. J. Physiol. 265, G1071–G1081.The quality of the final structure was adjudged using the program Hesterkamp, T., Weeds, A.G., and Mannherz, H.-G. (1993). The actinPROCHECK (Laskowski et al., 1993). The programs FRODO (Jones, monomers inthe ternary gelsolin: 2 actin complex are in an antiparal-1985), MOLSCRIPT (Kraulis, 1991; with modifications by R. Esnouf), lel orientation. Eur. J. Biochem. 218, 508–513.and RASTER3D (Merritt and Murphy, 1994) were used to view coor- Holmes, K.C., Popp, D., Gebhard, W., andKabsch, W. (1990). Atomicdinates and ALSCRIPT (Barton, 1993) to detail the sequence model of the actin filament. Nature 347, 44–49.alignment.

Jones, T.A. (1985). Interactive computer graphics: FRODO. Meth.Enzymol. 115, 157–171.

Acknowledgments Kabsch, W., and Sander, C. (1983). Dictionary of protein secondarystructure: pattern recognition of hydrogen-bonded and geometrical

We thank K. Harlos, E. Mitchell (ESRF), and the staff at the Synchro- features. Biopolymers 22, 2577–2637.tron Radiation Source, Daresbury, UK for help with X-ray data collec- Kraulis, P. (1991). MOLSCRIPT: a program to produce both detailedtion, and R. Bryan and R. Esnouf for computing facilities. We are and schematic plots of protein structures. J. Appl. Cryst. 24,grateful to J. Tatefor help with the preparation of figures. The Oxford 946–950.Centre for Molecular Sciences is supported by the Biotechnology

Kwiatkowski, D.J., Janmey, P.A., Mole, J.E., and Yin, H.L. (1985).and Biological Sciences Research Council and the Medical Re-Isolation and properties of two actin-binding domains in gelsolin.search Council (MRC). R. C. R. has been supported by a MRCJ. Biol. Chem. 260, 15232–15238.Research Studentship and by the Oxford Centre for Molecular Sci-Kwiatkowski, D.J., Janmey, P.A., and Yin, H.L. (1989). Identificationences; L. D. B. by the Natural Sciences and Engineering Council ofof critical functional and regulatory domains in gelsolin. J. Cell Biol.Canada and by the Royal Society Canada-UK Exchange Program;108, 1717–1726.P. J. M. L. by the Wellcome Trust; E. Y. J. by the Royal Society; and

D. I. S. by the MRC. Atomic coordinates for horse gelsolin have Kwiatkowski, D.J., Stossel, T.P., Orkin, S.H., Mole, J.E., Colten, H.,and Yin, H.L. (1986). Plasma and cytoplasmic gelsolins are encodedbeen deposited with the Protein Data Bank, Brookhaven National

Cell670

by a single gene and contain a duplicated actin binding domain. Weeds, A.G., Gooch, J., McLaughlin, P., Pope, B., Bengtsdotter, M.,Nature 323, 455–458. and Karlsson, R. (1995). Identification of the trapped calcium in

the gelsolin segment 1-actin complex: implications for the role ofLaskowski, R.J., Macarthur, M.W., Moss, D.S., and Thornton, J.M.calcium in the control of gelsolin activity. FEBS Letts. 360, 227–230.(1993). PROCHECK: a program to check the stereochemical quality

of protein structures. J. Appl. Cryst. 26, 283–290. Weeds, A.G., and Maciver, S. (1993). F-actin capping proteins. Curr.Opin. Cell Biol. 5, 63–69.Markus, M.A., Nakayama, T., Matsudaira, P., and Wagner, G. (1994).

Solution structure of villin 14T, a domain conserved among actin- Wen, D., Corina, K., Chow, E., Miller, S., Janmey, P., and Pepinsky,severing proteins. Prot. Sci. 3, 70–81. R. (1996). The plasma and cytoplasmic forms of human gelsolin

differ in disulfide structure. Biochemistry 35, 9700–9709.Maury, C.P. J., Alli, K., and Baumann, M. (1990). Finnish hereditaryamyloidosis. Amino acid sequence homology between the amyloid Witke, W., Sharpe, A.H., Hartwig, J.H., Azuma, T., Stossel, T.P., andfibril protein and human plasma gelsolin. FEBS Lett. 260, 85–87. Kwiatkowski, D.J. (1995). Hemostatic, inflammatory, and fibroblastMcGough, A., and Way, M. (1995). Molecular model of an actin responses are blunted in mice lacking gelsolin. Cell 81, 41–51.filament capped by a severing protein. J. Struct. Biol. 115, 144–150. Xian, W., Vegners, R., Janmey, P., and Braunlin, W. (1995). Spectro-McLaughlin, P., Gooch, J.T., Mannherz, M-G., and Weeds, A.G. scopic studies of a polyphosphoinositide-binding peptide from gel-(1993). Structure of gelsolin segment 1-actin complex and the mech- solin: behavior in solutions of mixed solvent and anionic micelles,anism of severing. Nature 364, 685–692. Biophys. J. 69, 2695–2702.Merritt, E.A., and Murphy, M.E.P. (1994). RASTER3D Version 2.0. Yin, H. (1987). Gelsolin: calcium and polyphosphoinositide-regu-A program for photorealistic molecular graphics. Acta Cryst. D50, lated actin-modulating protein. BioEssays 7, 176–179.869–873. Yu, F.-X., Sun, H.-W., Janmey, P.A., and Yin, H.L. (1992). Identifica-Otwinowski, Z. (1991). Oscillation data reduction program. In Iso- tion of a polyphosphoinositide-binding sequence in an actin mono-morphous Replacementand Anomalous Scattering (Warrington, UK: mer-binding domain of gelsolin. J. Biol. Chem. 267, 14616–14621.Daresbury Laboratory), pp. 80–86.

Otwinowski, Z. (1993). Oscillation data reduction program. In DataCollection and Processing, L. Sawyer, N. Issacs, and S. Bailey, eds.(Warrington, UK: SERC Daresbury Laboratory), pp. 56–62.

Pope, B., Way, M., and Weeds, A.G. (1991). Two of the three actinbinding domains of gelsolin bind to the same subdomain of actin.FEBS Lett. 280, 70–74.

Pope, B., Maciver, S., and Weeds, A. (1995). Localization of thecalcium-sensitive actin monomer binding site in gelsolin to segment4 and identification of calcium binding sites. Biochemistry 34, 1583–1588.

Reid, S.W., Koepf, E.K., and Burtnick, L.D. (1993). Fluorescent re-sponses of acrylodan-labeled plasma gelsolin. Arch. Biochem. Bio-phys. 302, 31–36.

Schnuchel, A., Wiltscheck, R., Eichinger, L, Schleicher, M., and Ho-lak, T.A. (1995). Structure of severin domain 2 in solution. J. Mol.Biol. 247, 21–27.

Sheldrick, G.M., Dauter, Z., Wilson, K.S., Hope, H., and Sieker, L.C.(1993). The application of direct methods and Patterson interpreta-tion to high-resolution native protein structure. Acta Cryst. D49,18–23.

Sheterlin, P., and Sparrow, J.C. (1994). Actin. Prot. Profile. 1, 1–121.

Stuart, D.I., Levine, M., Muirhead, M., and Stammers, D.K. (1979).The crystal structure of cat pyruvate kinase at a resolution of 2.6A. J. Mol. Biol. 134, 109–142.

Sun, H.-Q., Wooten, D.C., Janmey, P.A., and Yin, H.L. (1994). Theactin side-binding domain of gelsolin also caps actin filaments. Im-plications for actin filament severing. J. Biol. Chem. 269, 9473–9479.

Vandekerckhove, J. (1990). Actin binding proteins. Curr. Opin. CellBiol. 2, 41–50.

Van Troys, M., Dewitte, D., Goethals, M., Vanderkerckhove, J., andAmpe, C. (1996). Evidence for an actin binding helix in gelsolinsegment 2; have homologous sequences in segments 1 and 2 ofgelsolin evolved to divergent actin binding functions? FEBS Lett.397, 191–196.

Vasconcellos, C.A., and Lind, S.E. (1993). Coordinated inhibition ofactin-induced platelet aggregation by plasma gelsolin and vitaminD-binding protein. Blood 82, 3648–3657.

Way, M., Gooch, J., Pope, B., and Weeds, A.G. (1989). Expressionof human plasma gelsolin in Escherichia coli and dissection of actinbinding sites by segmental deletion mutagenesis. J. Cell Biol. 109,593–605.

Way, M., Pope, B., and Weeds, A.G. (1992). Evidence for functionalhomology in the F-actin binding domains of gelsolin and a-actinin:implications for the requirements of severing and capping. J. CellBiol. 119, 835–842.

Weeds, A.G., Gooch, J., McLaughlin, P., and Maury, C.P.J. (1993).Variant plasma gelsolin responsible for familial amyloidosis (Finnishtype) has defective actin severing activity. FEBS Lett. 335, 119–123.