SlideShare a Scribd company logo
Structure 
Article 
Biochemical Implications of a Three-Dimensional 
Model of Monomeric Actin Bound to 
Magnesium-Chelated ATP 
Keiji Takamoto,1,2,* J.K. Amisha Kamal,1,2 and Mark R. Chance1,2 
1Case Center for Proteomics, Case Western Reserve University, 10090 Euclid Avenue, Cleveland, OH 44106, USA 
2Lab address: http://casemed.case.edu/proteomics/ 
*Correspondence: keiji.takamoto@case.edu 
DOI 10.1016/j.str.2006.11.005 
SUMMARY 
Actin structure is of intense interest in biology 
due to its importance in cell function and 
motility mediated by the spatial and temporal 
regulation of actin monomer-filament intercon-versions 
in a wide range of developmental and 
disease states. Despite this interest, the struc-ture 
of many functionally important actin forms 
has eluded high-resolution analysis. Due to the 
propensity of actin monomers to assemble into 
filaments structural analysis of Mg-bound actin 
monomers has proven difficult, whereas high-resolution 
structures of actin with a diverse ar-ray 
of ligands that preclude polymerization have 
been quite successful. In this work, we provide 
a high-resolution structural model of the Mg- 
ATP-actin monomer using a combination of 
computational methods and experimental foot-printing 
data that we have previously published. 
The key conclusion of this study is that the 
structure of the nucleotide binding cleft defined 
by subdomains 2 and 4 is essentially closed, 
with specific contacts between two subdo-mains 
predicted by the data. 
INTRODUCTION 
Actin is a ubiquitous and important protein in eukaryotes 
and is extremely well conserved from yeast to man. Actin 
binding of nucleotides and its interactions with other pro-teins 
in the cell control the spatial and temporal assembly 
and disassembly of the cytoskeletal network; the careful 
regulation of this network has a profound influence on 
cell motility (Paavilainen et al., 2004; Schoenenberger 
et al., 2002; Wear et al., 2000; Winder, 2003). Even the 
nature of the metal ion bound to the nucleotide with the 
actin structure can profoundly alter the ability of actin 
monomers to assemble into filaments. In spite of its bio-logical 
importance, the structure of the Mg2+-ATP-bound 
form of the actin monomer (called G-actin) is not known. 
The determination of the high-resolution structure of 
Mg2+-G-actin is hampered by the propensity of actin to 
polymerize. Generation of crystals suitable for X-ray dif-fraction 
as well as NMR studies typically requires higher 
concentrations than the critical concentration for actin 
polymerization for the Mg2+-nucleotide-bound forms of 
the actin monomer. The high-resolution crystal structures 
of G-actin, whose richness and depth are a nearly unique 
resource for this investigation, never the less have either 
Ca2+-nucleotide bound to actin or, if Mg2+-bound nucleo-tide 
is used, actin is cocrystallized with ligands such as 
DNase I (Kabsch et al., 1985, 1990; Suck et al., 1981), gel-solin 
segment 1 (Mannherz et al., 1992; Vorobiev et al., 
2003), vitamin D binding protein (Otterbein et al., 2002; 
Swamy et al., 2002; Verboven et al., 2003), or macrolides 
(Allingham et al., 2004; Klenchin et al., 2003; Morton et al., 
2000; Reutzel et al., 2004; Yarmola et al., 2000). These li-gands 
typically prevent polymerization. Although struc-tures 
of Ca2+- and Mg2+-G-actin bound to ligands are 
overall similar in many respects, significant differences in 
structure and function between these species in the 
absence of other bound ligands in solution have been 
consistently reported, including biochemical/biophysical 
analyses such as limited proteolysis (Chen et al., 1995; 
Strzelecka-Golaszewska et al., 1993), fluorescence stud-ies 
(Frieden and Patane, 1985; Moraczewska et al., 1999; 
Selden et al., 1989; Valentin-Ranc and Carlier, 1991; Zim-merle 
et al., 1987), and molecular dynamics simulations 
(Wriggers and Schulten, 1997). 
Recently, we have investigated the solution structure of 
Mg2+-ATP-actin using hydroxyl-radical footprinting and 
mass spectrometry (MS); these experiments have in-cluded 
a comparison of Ca2+-ATP-actin in the presence 
and absence of gelsolin segment 1 (GS1) (Guan et al., 
2003). The measured side-chain solvent accessibilities 
of Ca2+-ATP-actin in the absence of GS1 are very similar 
to that in the presence of GS1 overall and consistent with 
the accessible surface area (ASA) calculated from solved 
crystal structures. In contrast, Mg2+-ATP-actin in the 
absence of GS1 is quite different in its side-chain surface 
accessibilities, particularly in subdomains (SD) 2 and 4. 
These differences are reversed in the presence of GS1, 
indicating the structure of the Mg2+-ATP-actin/GS1 com-plex 
in solution is consistent with crystal structure data. 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 39
Hydroxyl-radical footprinting and MS have been proven 
to be powerful tools for probing protein structure (Guan 
et al., 2003; Rashidzadeh et al., 2003) and conformational 
change (Kiselar et al., 2003a, 2003b), as well as for probing 
protein-ligand (Gupta et al., 2004) and protein-protein 
interactions (Guan et al., 2002, 2004; Liu et al., 2003). As 
the experimental details of the technique have matured 
and become reliable (Guan and Chance, 2005; Takamoto 
and Chance, 2006), we and others have attempted to use 
the data in conjunction with comparative modeling tech-niques 
to provide unique structural models (Gupta et al., 
2004; Sharp et al., 2005). In this study, we extend our 
previous work (Guan et al., 2003) with visualization of 
our descriptive prediction by generating an atomic model 
of the Mg2+-ATP-actin monomer structure using footprint-ing 
and crystallographic data. The computational strategy 
uses rigid-body rotations and translations of actin subdo-mains, 
primarily guided by known subdomain rearrange-ments 
from the range of actin crystallographic structures, 
in order to generate a structure consistent with surface 
accessibility data predicted by footprinting. In addition 
to an atomic model of the Mg2+-ATP-actin monomer, we 
also propose a specific mechanism for cleft closure 
mediated by changes in metal-ion coordination. 
RESULTS 
Rigid-Body Rearrangements of Actin 
Domain/Subdomains Derived from 
an Actin Crystallographic ‘‘Database’’ 
The wealth of actin structures available in the literature 
provides a nearly unique opportunity for examining the 
range of conformations accessible to actin in its various 
ligand-bound states. Actin is composed of two domains 
that are termed the large and small domains, and each do-main 
is composed of two subdomains (Figures 1A and 
1B). Table 1 provides a ‘‘database’’ of structures that 
encompass a number of relative arrangements of the var-ious 
actin subdomains. These are the major structures 
that are used in this paper to provide an analysis of the 
range of motions of the actin subdomains. Our approach 
in this section is to survey these structures and determine 
patterns of subdomain motions. 
Our previous footprinting data (Guan et al., 2003) 
indicated that the large cleft (nucleotide binding cleft) be-tween 
SD2 and SD4 is in a more closed configuration for 
monomeric actin bound to Mg2+-ATP compared to the 
Ca2+-ATP form. We surveyed the available crystal struc-tures 
to ascertain the range of relative motions of the large 
cleft. Most actin crystal structures (Table 1) are very 
similar, with an overall backbone rmsd of 0.8 A° 
for 13 
structures (including Protein Data Bank ID codes 1RFQ-A 
and 1RFQ-B, 1IJJ-A and 1IJJ-B, 1NM1, 1D4X, 1EQY-A, 
1NWK, 1QZ6, 1YAG, 1MA9, 1AQK, and 1NLV). However, 
there are some crystal structures that show significant de-viations 
from the majority of solved structures. One of the 
most prominent examples is crystal structure 1HLU (Chik 
et al., 1996), which has profilin bound to the small cleft 
Figure 1. Domain Rigid-Body Movements Observed in Actin 
Crystal Structures 
(A) The movement of the large domain relative to SD1 observed in crys-tal 
structure 1HLU. The models shown in tin and silver are 1YAG and 
1HLU, respectively. The movements of Ca atoms are indicated by 
arrows. The colors of arrows indicate the size of movements ranging 
from blue (small) to orange (large). The center of rotational movement 
and normal vector are shown by the yellow sphere and long arrow. 
(B) Schematic representation of large-domain movement. 
(C) The movements of SD2 relative to subdomain 1. The white arrows 
indicate deviations of the G63 Ca atom position from 1YAG. 
between SD1 and SD3; this structure exhibits a ‘‘super-open’’ 
nucleotide cleft, with a large movement of the 
ATP/cation complex within the cleft. This structure can be 
related to other more ‘‘typical’’ structures through domain 
movements by rotational/sheer transformation (Chik et al., 
1996; Page et al., 1998). A second example is 1RFQ 
40 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
Table 1. List of Crystal Structures Used for Analyses 
Protein Data Bank ID Code Species Cocrystal Cofactors Resolution Comments 
Crystal Structures of Interest 
1YAG Yeast GS1b Mg2+/ATP 1.90 ‘‘Template’’ molecule (with 
1RFQ-Ba Rabbit LARc Mg2+/ATP 3.00 Closed large cleft with altered 
1QZ5 Rabbit KABd Ca2+/ATP 1.45 Closed large cleft 
1HLU Bovine Profilin Ca2+/ATP 2.65 Large cleft in ‘‘super-open’’ 
Crystal Structures with Complete D Loop 
1J6Z Rabbit TMR Ca2+/ATP 1.54 TMR affects C-terminal structure? 
1YVN Yeast GS1 Mg2+/ATP 2.10 V159N mutant of 1YAG 
2BTF Bovine Profilin Sr2+/ATP 2.55 With Sr2+. b-actin 
1ATN Rabbit DNase I Ca2+/ATP 2.80 D loop interacts with DNase I 
1IJJ Rabbit LAR Mg2+/ATP 2.85 D loop that interacts with another 
1LCU Rabbit LAR Ca2+/ATP 3.50 
Crystal Structures with Partial D Loop 
1D4X C. elegans GS1b Ca2+/ATP 1.75 
1C0G Chimeric GS1b Ca2+/ATP 2.00 Chimeric, Q228K/T229A/A230Y/E360H 
1MDU Chicken GS1b Ca2+/ATP 2.20 Actin-trimer 
1C0F Chimeric GS1b Ca2+/ATP 2.40 Chimeric Dictyostelium/Tetrahymena 
1DEJ Chimeric GSe Mg2+/ATP 2.40 Chimeric, Q228K/T229A/A230Y/ 
1H1V Human GSe Ca2+/ATP 3.00 
Important crystal structures of monomeric actin. The first set is used for modeling process or strategy; the second set is actin struc-tures 
with complete D-loop backbone coordinates; and the last set is a partial but relatively better coverage of D-loop backbone 
coordinates. Except for 1RFQ, the interasymmetric unit interactions between their D loop and the other molecule are unknown. 
1RFQ does not have contact with other molecules. 
a 1RFQ-B, chain B of asymmetric unit in 1RFQ. 
b Gelsolin segment 1. 
c Latrunculin A. 
d Kabiramide C. 
e Gelsolin. 
(Reutzel et al., 2004), which has two chains in the asym-metric 
unit whose structures are quite different from 
each other. Chain A exhibits a ‘‘standard’’ cleft structure 
similar to those of the vast majority of solved structures. 
However, chain B (1RFQ-B) is interesting as it shows 
a ‘‘closed’’ form of the cleft, with very different geometries 
for the residues inside the cleft. The closed form in chain B 
of the crystal structure exemplifies a database entry that 
qualitatively satisfies an important aspect of our footprint-ing 
data, that is, it provides an example of how relative 
subdomain motions can result in a closed nucleotide cleft. 
Crystal structure 1QZ5 (Klenchin et al., 2003) shows a 
closed large cleft as well. In this structure, the geometry in-side 
the large cleft is almost identical to 1YAG (although it 
has a Ca2+ ion instead of an Mg2+ ion). To achieve cleft 
closure compared to canonical actin forms, SD2 moves 
complete D loop) 
side-chain geometries 
state, b-actin 
chain’s small cleft 
A231K/S232E/E360H 
toward SD4 without tilting (unlike 1RFQ-B), minimizing ste-ric 
conflicts between the turn in SD2 (residues 62–65) and 
the top of the helix in SD4 (residues 200–205). Although our 
hydroxyl-radical footprinting results indicate more signifi-cant 
closure of the cleft than observed in these structures, 
these structures can serve as ‘‘qualitative’’ templates for 
understanding the Mg2+ form of G-actin in solution. The 
B factors of the atoms in SD2 of 1QZ5 are relatively high, 
indicating the relative mobility of SD2. This suggests the 
likelihood of high relative mobility of SD2 (not only the D 
loop but also the SD2 core) in solution (see the Supple-mental 
Data available with this article online). 
Vector Analysis of Subdomain Relative Positions 
The structure 1YAG is used as our standard structural 
template having high-resolution (1.90 A° 
) and well-defined 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 41 
Structure 
Model of Mg2+-ATP-Actin Monomer
waters within the nucleotide binding pocket. The struc-tures 
of Protein Data Bank ID codes 1HLU (Chik et al., 
1996), 1QZ5 (Klenchin et al., 2003), 1RFQ-B (Reutzel 
et al., 2004), 1J6Z (Otterbein et al., 2001), 1ATN (Kabsch 
et al., 1990), 1S22 (Allingham et al., 2004), 1EQY-A 
(Mannherz et al., 1992), 1IJJ-A (Bubb et al., 2002), 1MA9 
(Verboven et al., 2003), 1KXP (Otterbein et al., 2002), 
and 2BTF (Schutt et al., 1993) were compared with 1YAG, 
and differences between Ca/backbone coordinates were 
visualized using Tcl scripts for visual molecular dynamics 
(VMD). The script also calculates the average center 
position and rotational normal vector of movement of 
Ca/backbone coordinates. 
Figure 1A shows the relative movements (vector) of the 
large domain with respect to SD1 in the comparison of 
1YAG (the standard) and 1HLU (most open nucleotide 
cleft structure). The arrows indicate the Ca movements 
(arrows from coordinates in 1YAG to 1HLU). As the move-ments 
are relative, we chose to show the large-domain 
movement relative to SD1 in this representation because 
SD2 also shows significant movement against SD1. As 
seen in Figure 1B, this movement is 8.9 rotational (hinge) 
movement that has a center position in the junction be-tween 
SD1 and SD3 (around Q137-A138). The axis of ro-tation 
goes from the front to the back side of the molecule 
(where the front side is defined as the surface of the actin 
molecule that exposes ATP). Figure 1B illustrates this 
movement (the rotational angle is exaggerated for clarity). 
A similar movement is reported in previous analyses and 
refinement of F-actin fiber X-ray data including alterations 
of the nucleotide binding pocket (Lorenz et al., 1993; Tirion 
et al., 1995). 
The SD2 subdomain is extremely mobile even in crystal 
structures. Figure 1C compares 11 SD2 structures with 
that of 1YAG; SD2 is observed in many orientations rela-tive 
to SD1. The directions of rotation of these crystal 
structures (relative to 1YAG) are shown in Figure 1C by 
arrows. All of these differences can be mediated by 
hinge movements having pivot points near P32-S33 and 
Y69-P70. 
Intracleft Interactions 
There are a number of interactions between the ATP/ 
cation complex and the protein within this cleft that are 
important for understanding large-cleft structure and the 
possible mechanisms of the structural changes. The sub-domains 
are not strongly connected by hydrogen-bond 
networks or other forms of interactions; the majority of 
the interactions between subdomains are mediated 
through a hydrogen-bond network with the ATP molecule 
and with waters in the pocket of the cleft. SD1 has interac-tions 
with phosphate oxygen atoms of ATP through resi-dues 
S14, G15, L/M16, and K18 (Figure 2A). Connections 
between SD1 and SD2 are mediated only by the hydrogen 
bonds between S14 and G74 (Figure 2B) (Chen et al., 
1995; Kabsch et al., 1990; Schu¨ ler, 2001). This link is 
severed in 1RFQ-B due to subdomain rearrangements. 
Although the metal ion is located near the interface of sub-domains 
1 and 2, it only interacts with SD1 by hydrogen 
Figure 2. Important Interactions Involving the ATP Molecule 
Brown and silver colors are 1YAG and 1FRQ, respectively. Hydrogen 
bonds and coordination bonds are shown by broken lines. 
(A) N-terminal b strand and ATP-metal interactions. 
(B) Hydrogen bonds that connect subdomains 1 and 2 and ATP. 
(C) Interactions with the Mg2+ ion. Water molecules observed in 1YAG 
are shown as blue balls. 
bonds through waters (no inner-sphere coordination) 
and no interaction is formed with SD2 residues. 
The interactions between SD3 and SD4 are also mainly 
mediated through ATP. In addition, there is a poorly 
packed hydrophobic cluster at the interface of the two 
42 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
Table 2. Probe Residue Solvent Accessibilities from Crystal Structures and Experimentally Determined 
Hydroxyl-Radical Reactivities 
Crystal Structures 
1ATN 1YAG 1RFQB 1RDW Footprinting Data 
Ca Mg Mg Mg Solution G-Actin 
DNase I GS1 LAR LAR Ca Mg 
subdomains. The adenosine moiety of ATP is located in 
the pocket between the two subdomains, with hydrogen 
bond and van der Waals contacts with surrounding resi-dues. 
Thus, SD3, SD4, and ATP form an interdependent 
set of interactions. 
Metal-Ion Coordination 
In most nonactin protein structures with ATP bound, mag-nesium 
ion coordinates to at least one residue from the 
protein (14 structures out of 16 nonredundant structures 
examined; data not shown; Dudev et al., 1999). In the 
case of actin, the magnesium ion (or calcium ion) forms in-ner- 
sphere coordination bonds with phosphate oxygens 
and water molecules, but not side-chain oxygen atoms, 
with a coordination number of 6 (or 7 for Ca2+). A notice-able 
difference between 1YAG (or other structures) and 
1RFQ-B is that residue atoms Q137:OE1 and D11:OD1, 
OD2 are closer to the Mg2+ ion in 1RFQ-B (3.0, 3.5, and 
4.0 A ° 
) compared to 1YAG (4.4, 4.3, and 4.2 A ° 
, respec-tively). 
Figure 2C shows the difference in the geometries 
surrounding the magnesium ion. Unfortunately for 1RFQ, 
at a resolution of 3.0 A ° 
, ordered water molecules are not 
observed within the nucleotide binding cleft. However, 
some water molecules observed in 1YAG must be radi-cally 
reorganized in 1RFQ-B, as these water molecules 
would clash with side-chain oxygen atoms (Figure 2C, 
pink side-chain oxygen atoms overlapping with blue water 
oxygen atoms). Based on these data, we used the geom-etry 
inside the large cleft from 1RFQ-B as a template for 
our structural modeling of the nucleotide binding cleft. 
Comparison of Crystallographic Data and 
Hydroxyl-Radical Footprinting Data 
Table 2 summarizes the calculated ASAs and rate con-stants 
of modification of actin peptides analyzed by 
hydroxyl-radical footprinting (Guan et al., 2003). 
The hydroxyl-radical footprinting data for Ca2+-ATP-G-actin 
are generally consistent with the calculated ASA, 
that is, solvent-accessible, reactive residues show oxida-tion 
and inaccessible residues are not appreciably oxi-dized. 
However, a number of probe residues in SD2 and 
SD4 and at the C terminus (SD1) show decreased rates 
of modification for Mg2+-ATP-G-actin where the solvent 
accessibilities for the structure are similar to those for 
Ca2+-ATP-G-actin. These sites that show protections 
from oxidation for Mg- versus Ca-actin are illustrated in 
Residue 
Number Residue 
Side-chain 
ASA (A° 2) 
Side-chain 
ASA (A° 2) 
Side-chain 
ASA (A° 2) 
Side-chain 
ASA (A° 2) 
Rate 
constant 
(s1) 
Rate 
constant 
(s1) 
Peptide 
(Residue 
Numbers) 
21 Phe 34.32 23.87 37.13 34.39 0.65/0.70 0.32/0.43 19–28 11–23 
44 Met 79.35 (16.31)a 137.63 (35.96)a N.D.b N.D.b 33 20 40–50 
47 Met 104.58 (30.94)a 113.37 (24.81)a N.D.b N.D.b 33 20 
53 Tyr 14.26 6.60 64.67c 55.88c 0.69 0.32 51–61 
67 Leu 28.70 22.16 25.46 40.25 0.40 0.08 63–68 
69 Tyr 80.46 98.57 68.54 66.81 1.1 0.18 58–72 
200 Phe 0.36 3.61 10.79 5.50 1.10/0.83 0.26/0.24 197–206 
196–207 
201 Val 82.01 65.45 108.42 101.42 1.10/0.83 0.26/0.24 
202 Thr 81.61 72.45 43.79 73.26 1.10/0.83 0.26/0.24 
243 Pro 43.26 63.97 72.23 68.46 0.69/0.75 0.26/0.30 239–254 
242–253 
362 Tyr 12.59 9.05 15.18 11.75 0.72 0.38 362–372 
367 Pro 45.75 40.74 39.49 41.88 0.72 0.38 
371 His 27.79 26.53 34.71 25.89 0.72 0.38 
374 Cys N.D. 3.53 49.40 24.29 8.5 3.6 363–375 
375 Phe N.D. 157.72 77.56 44.38 N.D.d N.D.d 
a Values in parentheses are ASAs for the sulfur atom in the side chain. 
b These crystals lacking observed D-loop structures. 
c These values are affected by the lack of a D loop (larger than expected). 
d The rate constant in the peptide is dominated by C374 and thus could not be determined. 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 43 
Structure 
Model of Mg2+-ATP-Actin Monomer
Figure 3. For example, probe residues in the D loop are 
modestly protected from oxidative modifications (40% 
reduction in rate constant), whereas residues 200–202 
and 243 also experience strong reductions in modification 
rate (75% and 60%, respectively). L67 and Y69 lo-cated 
inside the large cleft (Figure 3) experience an 80% 
reduction in modification rates. Consistent with these find-ings 
are limited proteolysis data, where cleavage between 
K68 and Y69 is almost completely suppressed in Mg2+- 
G-actin compared to Ca2+-G-actin (Chen et al., 1995; 
Strzelecka-Golaszewska et al., 1993). A closure of the 
large cleft is consistent with these data. 
Modeling Strategy: Overall Structural 
Considerations 
In the previous section, we analyzed possible relative 
movements of domains/subdomains. In light of these 
analyses, we can understand the cleft structures in each 
case. DNase I binding prevents large-cleft closure by 
prohibiting the movements of SD2 and the large domain 
by interacting with both of them (Kabsch et al., 1990). 
Latrunculin A binds between SD2 and SD4 (Bubb et al., 
2002; Reutzel et al., 2004); this prevents movements of 
those subdomains. Gelsolin segment 1 (Mannherz et al., 
1992; McLaughlin et al., 1993) and vitamin D binding pro-tein 
(Otterbein et al., 2002) prevent the domain movement 
between SD1 and SD3 blocking corresponding large-domain 
movements. Some macrolides (kabiramide C, 
jaspisamide A, and ulapualide A; Allingham et al., 2004; 
Klenchin et al., 2003) bind to the small cleft and block its 
movement as well. This is summarized in Figure 4A. It is 
important to note that macrolides are small enough to fit 
into small or large clefts and do not seem to prevent poly-merization 
by steric hindrance as opposed to actin binding 
proteins that result in steric blockage. It appears that 
these macrolides interfere with polymerization by prevent-ing 
domain/subdomain motions. 
The analysis of the various crystal structures and the 
footprinting data (Guan et al., 2003) provides significant 
clues about how the structure of Mg2+-G-actin differs 
from that of Ca2+-G-actin. First, the footprinting data indi-cate 
that the large cleft between SD2 and SD4 is almost 
completely closed. Second, the analyses of the various 
crystal structures indicate that movement of the SD2 core 
is a rigid-body movement. The movement of SD2 toward 
the large domain alone cannot explain the protections of 
Figure 3. Analyses of Hydroxyl-Radical Reactivity and Struc-tures 
The protection sites in Mg2+-G-actin (compared with Ca2+-G-actin) are 
displayed with residues. Colors are coded as blue (strong protection) 
to red (enhancement) by reactivity changes. 
Figure 4. Schematic Representation of Domain/Subdomain 
Movements 
(A) The effects on subdomain movements by ligand binding. 
Actin binding proteins (top) bind to actin and prevent polymerization, 
possibly by steric hindrance or interference of domain movements. 
Macrolides are small enough to fit into small or large clefts and interfere 
with the domain movements. They may prevent polymerization by this 
mechanism. 
(B) The proposed mechanism for large-cleft closure. SD2 movement 
alone cannot explain the complete closure of the large cleft (top right). 
The partial closure by large-domain movement brings SD4 and SD2 
close enough to allow interactions between the two subdomains, re-sulting 
in complete cleft closure. 
44 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
both L67 and Y69. Y69 is located deep within the large cleft 
and cannot be buried by SD2 rigid-body movement. In ad-dition, 
neither M44 nor M47 can reach SD4 to form pair-wise 
interactions to protect P243 and M44/47. This sug-gests 
that both the large domain and SD2 need to move 
as rigid bodies relative to SD1 in order to explain protec-tions 
observed in SD2 and SD4 (Figure 4B). 
The hydrogen-bond network among phosphate, S14, 
and G74 is the only one actually connecting SD2 to SD1 
and is observed in almost all structures (Schu¨ ler, 2001). 
This linkage is severed in the 1RFQ-B structure. We spec-ulate 
that latrunculin A prevents the movement of SD2 fur-ther 
from the observed structure although the geometry 
inside the large cleft is changed. As a result, the hydrogen 
bonds are broken. This hydrogen-bond linkage was main-tained 
in the model structure. We constructed a series of 
structures with combinations of different rotational angles 
and normal vectors for SD2 in order to determine the best-fit 
model for the hydroxyl-radical data. 
Regions Excluded from Modeling 
Although there is evidence for structural variation in the 
C-terminal region (Y362 to F375 changes solvent accessi-bility 
in hydroxyl-radical footprinting; Guan et al., 2003) 
and other data (Frieden and Patane, 1985; Strzelecka- 
Golaszewska et al., 1993; Valentin-Ranc and Carlier, 1991; 
Zimmerle et al., 1987), we do not have clear evidence to 
support the specific structural differences. Thus, we have 
not modeled the structural differences in this region. F21 
experiences an 50% reduction in modification rate in 
Mg2+-G-actin compared to Ca2+-G-actin, but this change 
cannot be explained with our modeling strategy. Our 
speculation is that the C-terminal (residues 336–375) 
and N-terminal (residues 1–33) regions interacting with 
SD1 are also affected by the movements of SD2 (con-nected 
to the N-terminal region) and the large domain 
(connected to the C-terminal region) and mediate these 
additional structural differences between the two forms. 
Model of the Mg2+-ATP-Actin Monomer 
In general, the hydroxyl-radical footprinting data are 
very consistent with known crystal structures. Solvent-accessible 
surface areas and rate constants of oxidation 
on side chains in known structures are in good correlation 
(Guan et al., 2004; Guan and Chance, 2005; Takamoto 
and Chance, 2006; Xu and Chance, 2005). This is not sur-prising, 
as the size of the hydroxyl radical is very close to 
that of a water molecule. Although it is not easy to com-pare 
rate constants among different side chains (such as 
Leu and Phe), it is very quantitative and reliable for the 
same residue in different conformational states. The site 
of oxidation is determined by MS/MS analyses that are 
well established and routine. With the use of different 
proteases, protein sequence coverage by MS analysis is 
usually 80%–90% and most of the probe residues can 
be detected. In our previous study, we covered 90% of 
the actin sequence with trypsin and Asp-N proteases. 
Thus, it is important to note that we have used high-quality 
data for our modeling. 
Figure 5. The Model 
(A) Overlaid structure of the model (silver) and template structure 1YAG 
(brown). The residues that experience protections are displayed as 
sticks and bubbles. Structures are aligned by subdomain 1. 
(B) The difference between the model and template structure 1YAG. 
The arrows indicate the differences of backbone atom positions. The 
sphere and long arrow indicate the center of rotation and its normal 
vectors for rigid-body movements (green for large domain and yellow 
for subdomain 2). 
Figure 5A shows the final model and the template struc-ture 
for Mg2+-G-actin. The residues indicated to change 
conformation in the footprinting experiments are shown 
in stick-and-bubble format for both structures. The struc-tures 
are aligned at SD1 in order to show both large-domain 
and SD2 movements. The large domain rotates 
toward the small domain and SD2 rotates to the front 
(as defined previously) and toward the large domain. 
Figure 5B show the backbone movement analyzed by 
VMD/Tcl scripts. Figures 6A and 6B show a magnified view 
of the SD2/SD4 interface in the model. In order to make 
contact between D-loop residue M44 and the hydropho-bic 
pocket in SD4 around protected residue P243, SD2 
was moved toward the front side along with its movement 
toward the large domain. As shown in Figure 6C, the M44 
side chain is inserted into the hydrophobic pocket and 
forms a contact with the P243 side chain. M47 cannot 
form this interaction with the hydrophobic pocket without 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 45 
Structure 
Model of Mg2+-ATP-Actin Monomer
forming improper bond angles in the context of these 
rigid-body movements. The two residues (M44 and 
P243) are simultaneously protected from solvent expo-sure 
through the ‘‘pairwise’’ interactions. On the other 
hand, M47 is exposed to solvent almost completely. 
This makes predictions that could be tested in future foot-printing 
experiments using high-resolution tandem MS. 
In systematic mutational analysis experiments on yeast 
actin, the double-mutation A204E/P243K abolishes poly-merization 
(Joel et al., 2004). These mutations introduce 
bulky and charged residues at the place where contacts 
are formed in our model, likely blocking this conforma-tional 
change. However, because these residues are also 
involved in the intermolecular contacts of the F-actin 
model, it is also possible that the mutations may hinder 
F-actin filament assembly. 
The closure of the large cleft in our model allows the 
formation of a hydrogen-bond network between SD2 
and SD4 (Figure 6D). In our model, seven new hydrogen 
Figure 6. Magnified Views of Interface 
between SD2 and SD4 
(A) Template structure 1YAG. 
(B) The model. The residues that experience 
protection are colored blue (in SD2) or red 
(SD4). 
(C) The interaction between the SD4 hydropho-bic 
pocket and D-loop M44 side chain. The 
side chain of P243 is colored green while the 
side chain of M44 is displayed as bubbles (sul-fur 
is colored yellow). 
(D) Newly formed hydrogen-bond network be-tween 
SD2 and SD4 in the model. Residues are 
silver in SD2 and brown in SD4. The possible 
p-stacking is indicated by surface models on 
residues Y69 and R183. 
bonds can be formed between SD2 and SD4, as indicated 
in the figure. Also, Y69 and R183 are in good locations for 
p-stacking interactions (this interaction may not be critical 
to form the structure, as the R183A/D184A mutant has no 
significant change in phenotype). In a comparison of ASA 
between 1YAG (rabbit sequence) and the model (Figure 7), 
M47’s side-chain ASA increased by almost 50%. On the 
other hand, M44 experiences almost complete burial of its 
side chain. If we take into account the extremely dynamic 
nature of the D loop, it is reasonable to assume that 
both M44 and M47 are much more accessible than the 
structure observed in 1YAG. Overall, the accessibility of 
M44/47 seen in footprinting is in good agreement with 
the model. The ASA changes in Y53 can be explained in 
the same way. The D-loop structure is highly mobile, and 
in the crystal structure the side chain of Y53 in 1YAG is 
covered by the D loop and appears buried, but this does 
not reflect the ensemble average experienced by the fluc-tuating 
structure. In the model, although the ASA of the 
Figure 7. Changes in Accessible Surface 
Areas of Probe Residues 
The red bars are ASAs of 1YAG (rabbit) and the 
blue bars are differences between 1YAG and 
the model. Negative values indicate less ex-posed 
in the model. 
46 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
Y53 side chain increases about 25A° 
2, the D-loop structure 
should be stiff through its formation of an interaction with 
SD4. Thus, if the dynamic nature of structural changes in 
SD2 is considered, this residue can experience a lower 
reactivity consistent with the structural changes seen in 
the model. 
DISCUSSION 
Possible Mechanism of Cleft Closure 
Our model is based on rigid-body movements of domains 
and subdomains observed in solved crystal structures. 
We applied changes strictly to the backbone angles that 
are involved in domain movements because we do not 
have any data that indicate there is large-scale reorgani-zation 
of structure other than domain movements. The 
large-domain movement relative to SD1 narrowed the 
large cleft including the nucleotide binding pocket as pro-posed 
in previous reports of F-actin structure (Belmont 
et al., 1999; Lorenz et al., 1993; Tirion et al., 1995). In the 
crystal structures except 1RFQ-B, the metal ion is coordi-nated 
only by water molecules and phosphate oxygen 
atoms. The narrowed cleft results in a closer proximity of 
side-chain oxygen atoms (D11:OD1, OD2, and Q137:OE1) 
to metal ions that coordinate phosphate oxygen atoms. In 
our model, the side chains of these residues form inner-sphere 
coordination to Mg2+. AsMg2+ ions strongly prefer 
to coordinate to oxygen atoms of side chains instead of 
those of water molecules with a likely gain of free energy 
(Dudev et al., 1999), it is reasonable to assume that ex-change 
of inner-sphere coordinations between water mol-ecules 
and oxygen atoms in D11 and Q137 occurs spon-taneously, 
as thermal fluctuations bring side-chain oxygen 
atoms close enough to the metal ions mediating the ex-change 
reaction. This can then be a driving force for the 
movement of the large domain toward SD1. Interestingly, 
there is no strong interaction between SD1 and SD3—no 
hydrogen-bond network nor hydrophobic interactions. 
Thus, there is no discernable preference (or penalty) for 
these subdomain motions compared to the structures ob-served 
in the crystals. Once the large domain is locked in 
the closed form by formation of inner-sphere coordination 
between side-chain oxygen atoms and the metal ion, SD2 
and SD4 are closer to each other and in the range where 
a detailed hydrogen-bond network (and possibly p-stack-ing) 
can form between them. This hydrogen-bond network 
and p-stacking stabilize and lock a ‘‘super-closed’’ form 
of the large cleft. The invariant residue Y69 appears to 
be a key residue for the interaction between SD2 and 
SD4. It possibly forms hydrogen bonds with the residues 
in SD4 at the bottom of the large cleft and acts as a latch. 
Mutations of residues forming interactions between SD2 
and SD4 (Figure 6D) are reportedly lethal in yeast (Sheter-line 
and Sparrow, 1994; Wertman et al., 1992) (R62, R206, 
E207), although these residues located within the cleft are 
unlikely to be involved in protein/protein interactions. 
The contribution of the D loop to cleft closure is 
documented in the literature (Khaitlina and Strzelecka- 
Golaszewska, 2002). The cleavages in the D loop by 
ECP32 protease (G42-V43) or subtilisin (M47-G48) greatly 
stimulate the tryptic susceptibility of residues within SD2 
(R62 and K68). The cleavage of the D loop apparently 
makes those accessibilities of tryptic cleavage sites in-crease. 
Thus, the D loop contributes to the stability of the 
closed structure although it is not clear how much contri-bution 
to the stability comes from the D loop. Our model 
predicts the interaction between M44 and the SD4 hydro-phobic 
pocket contributes to stabilization of the closed 
structure of the large cleft. The observations above sup-port 
our model, as the D-loop/SD4 interaction is a strong 
candidate for the change in dynamics of SD2. 
Catalytic Residue 
As a consequence of the large-cleft closure by large-domain 
movements relative to SD1, the geometry inside 
the nucleotide binding pocket is altered from those seen in 
typical crystal structures. The narrowed cleft brings D11, 
Q137, and D154 side chains closer to the metal ion and 
the b and g phosphates while maintaining the hydrogen-bond 
network already in place. 
D11 and Q137 are likely to be involved in the inner-sphere 
metal-ion coordination. On the other hand, D154 
seems to be too far from the metal ion to coordinate but 
is located in a good position to form an inline attack of the 
g phosphate. The double-mutant D154A/D157A is lethal in 
yeast (Wertman et al., 1992); however, the S14C/D157A 
double mutant does not show significant change in 
ATPase activity (Schu¨ ler et al., 1999). Thus, it is very likely 
that the D154A mutation is directly involved in lethality. As 
suggested in the literature (Schu¨ ler, 2001), D154 appears 
to be a prime candidate for the catalytic residue from both 
previous reports and our model. 
Actin Model Consistency with Nucleotide 
and Metal-Ion Exchange Rates 
The stability constants of the Mg2+- and Ca2+-ATP com-plexes 
are similar (4.2 and 4.0, respectively) (Williams, 
1970), and the exchange rate constants of the metal 
ions and ATP in water are quite rapid (3.3 3 105 and 2.0 3 
105 s1) (De La Cruz and Pollard, 1995; Pecoraro et al., 
1984). On the other hand, the exchange rates of Mg2+ 
and Ca2+ bound to ATP-G-actin are quite slow compared 
with the rates observed in water and are different from 
each other (association rates: 2.3 3 105 and 2 3 107 
M1 s1; dissociation rates: 0.0015 and 0.014 s1 for 
Mg2+- and Ca2+-ATP-G-actin, respectively, as summa-rized 
in Table 3) (Gershman et al., 1991; Selden et al., 
1989; Sheterline and Sparrow, 1994). These data indicate 
that the metal ions are significantly stabilized within the 
cleft by limited diffusion or coordination bond formation 
with protein side-chain oxygen atoms. The difference in 
association/dissociation with ATP between Mg2+ and 
Ca2+ is not significant in water but is an order of magnitude 
different in the case of actin-nucleotide-bound forms. Our 
model predicts that one factor explaining the difference in 
stability is coordination of Mg2+ with oxygen atoms of the 
protein. The other factor is the difference in accessibility of 
metal ions (steric effect). The exchange rates of ATP for 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 47 
Structure 
Model of Mg2+-ATP-Actin Monomer
Table 3. Exchange Rate Constants of Nucleotides/Metal Ions for Different Species of Actin 
Rate Constant Mg2+ Ca2+ Species 
Metal ion Exchange rate 3.3 3 105 s1 2.0 3 105 s1 ATP/metal ion 
Metal ion Association rate 2.3 3 105 M1 s1 2.0 3 107 M1 s1 Metal ion/ATP + actin 
Metal ion Dissociation rate 0.0015 s1 0.014 s1 Metal ion/ATP/actin 
ATP Exchange ratea 5.0 3 104 s1b 1.5 3 102 s1c Metal ion/ATP/actin 
Metal ion Affinity (Kd) 10 nM 2 nM ATP-G-actin 
a The exchange rate is highly dependent on metal-ion concentrations. 
bMg2+ concentration of 1 mM. 
c Ca2+ concentration of 0.1 mM. 
both Mg2+- and Ca2+-G-actin are even more different (5 3 
104 and 1.5 3 102, respectively) (Sheterline and Spar-row, 
1994). As ATP exchange rates are tightly connected 
to free metal concentrations, the metal-ion exchange 
rate would be a limiting factor for nucleotide exchange 
(Kinosian et al., 1993). Even if we attribute all of the metal 
exchange rate difference to coordination states, still there 
is a difference in ATP exchange rates. Thus, this ATP ex-change 
rate difference may be related to the steric effect 
of the cleft closure. These observed exchange rate differ-ences 
between Mg2+- and Ca2+-G-actin are quite consis-tent 
with our model (closed cleft and direct coordination of 
Mg2+ to the side-chain oxygen atoms). 
Implications for F-Actin Models 
The Holmes model of F-actin (Holmes et al., 1990; Lorenz 
et al., 1993; Tirion et al., 1995) has been widely accepted 
and is largely consistent with X-ray fiber diffraction (Bel-mont 
et al., 1999; Holmes et al., 1990; Lorenz et al., 
1993; Oda et al., 2001; Tirion et al., 1995; Wu and Ma, 
2004), cryoelectron microscopy (Schmid et al., 2004), 
and atomic force microscopy data (Shao et al., 2000; Shi 
et al., 2001). We have examined the F-actin structure by 
side-chain solvent accessibility utilizing hydroxyl-radical 
footprinting (Guan et al., 2005). Although our data were 
mainly consistent with the F-actin model, there are some 
disagreements with the Holmes model. The most appar-ent 
difference is the status of the large cleft. The protection 
sites in SD4 (residues 200–202 and 243) and the D loop 
(residues 40, 44, and 47) form contacts in the intermolec-ular 
interfaces; there is no disagreement with footprinting 
data for these residues. However, the hydroxyl-radical re-activity 
within the large cleft is significantly lower in F-actin 
compared with Ca2+-G-actin. A closed cleft is also ob-served 
in a reconstructed model from cryo-EM data 
(Schmid et al., 2004), although the degree of closure 
seems not to be enough to explain the almost 90% reduc-tion 
in hydroxyl-radical reactivity compared to the open 
form of Ca2+-G-actin. 
The ADP and ADP-BeF3 
 forms of fiber diffraction data 
show differences in the status of the large cleft (Belmont 
et al., 1999); the latter form indicates a more closed cleft. 
The replacement of ADP with ADP-BeF3 
 also diminishes 
susceptibility of F-actin to tryptic cleavage within SD2 
(Muhlrad et al., 1994). As ADP-BeF3 
 mimics ATP or 
ADP-Pi (Muhlrad et al., 1994), this form represents ATP 
or ADP-Pi-F-actin before it releasesPi. These observations 
indicate that in the F-actin filament, hydrolysis of ATP to 
ADPand subsequent release of Pi correlate with large-cleft 
opening. On the other hand, when Mg-ATP-G-actin is in-corporated 
into the filament, we suggest it is in the highly 
closed form seen here; this Mg-ATP form of actin has a 
lower critical concentration than Ca2+-ATP-G-actin. Catal-ysis 
mediated by the Mg2+ ion may drive conformational 
changes among the coordination bonds between side 
chains and the metal ion, leading to rearrangements of 
subdomains 2 and 4 and the generation of a more open 
cleft within the ADP-F-actin structure. This would provide 
a more flexible and dynamic SD2, which is thought to be 
a key factor in attracting binding proteins and driving fila-ment 
dynamics (Schmid et al., 2004). 
Conclusion 
Understanding monomeric G-actin structure is important, 
as it is relevant to many protein/protein interactions in 
which actin participates. Hydroxyl-radical footprinting is 
a powerful technique to probe surface accessibility of 
macromolecules. The technique is a local measure as it 
reports side-chain accessibility. Such information pro-vides 
strict restraints for estimating changes in structure 
and in defining domain interactions. In conjunction with 
a computational approach, a structural model of the Mg- 
ATP-actin monomer was built that is consistent with ex-perimental 
data. This model provides novel insights into 
the conformational rearrangements driving actin filament 
dynamics and provides a unique structural insight into ac-tin 
monomer structure. This combined computational and 
experimental approach is particularly useful for modeling 
the structural changes of proteins when we do not have 
access to crystal or solution structures for conformations 
of functional interest, but where we may have suitable 
starting templates for modeling the structure. Further 
developments in computational modeling will ease the 
integration of the experimental into the computational 
approach as a filtering/validation tool. 
EXPERIMENTAL PROCEDURES 
Rigid-Body Movement Analyses 
Tcl scripts for molecular viewer VMD (Humphrey et al., 1996) were de-veloped 
for the analysis and visualization of rigid-body rotational 
48 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
movements of domains/subdomains. The scripts accept the range of 
the target residues in two structures for rigid-body movement and then 
calculate vectors of atom movements and averaged normal for these 
vectors; they then search for the atoms within the reference structure 
that are nearest to the center of motion. The algorithm is based on the 
assumption that there are pivot points in bonds within the backbone 
allowing rigid-body rotational movements (the movements should be 
explained by rotations of some bond angles). The algorithm is based 
on the following step-by-step calculations. 
Define structures A and B, where A is the reference structure and B 
is the target structure for movement calculation. The total number of 
atoms(Ca orbackbone atoms) isn within the target range.The totalnum-ber 
of the atoms in reference structure A (Ca or backbone atoms) is N. 
v ! 
Atoms in structure A: ai (i = 1 to n), coordinates Pai. 
Atoms in structure B: bi (i = 1 to n), coordinates Pbi. 
Atoms for the center C: cj (j = 1 to N), coordinates Pcj. 
The midpoint of ai and bi: mi (i = 1 to n), coordinates Pmi. 
The vector between points ai and bi: i = Pbi  Pai . 
The vector between points mi and cj: u ! 
ij =Pmi  Pcj . 
The angle formed by the two vectors v ! 
i and u ! 
ij : 
qij = 
 
acos 
 
ð v ! 
i  u ! 
jiÞ 
jv ! 
 
i j 
u ! 
ji 
 
 
: 
The averaged angle for the point Pcj: 
q 
j = 
1 
n 
Xn 
i =1 
 
qij  
p 
2 
 
ð j =1toNÞ: 
The SD of the angles for atom Pcj: 
SDj = 
ffiPffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi n 
i =1ððqij  p 
2 
 
q 
j 
2 
n  1 
s 
ð j =1toNÞ 
The averaged angle between the real center of movement cr and the 
2; thus, q 
two atoms in structures A and B should be p 
j is 0. We need to 
find the atom that gives the averaged angle closest to 0. In order to 
avoid the cases where q 
j happens to be close to 0 but angles vary 
as qij can be positive or negative, the score is calculated by the follow-ing 
formula: 
Sj =q 
j 3SDj : 
Then, the script calculates the average positions of the top scored 
five atoms. This point is defined as gc. 
Now, we need to calculate the normal vector for the rotational move-ments. 
The normal vector for the rotational movements should be ap-proximately 
perpendicular to all vectors from the center of movement 
to the midpoint of atoms in the two structures A and B. The normal vec-tor 
is also perpendicular to the vector vi between ai and bi. In order to 
minimize the contribution from small vectors that tend to have larger 
deviation from ideal ones (the vectors are extended by unitization 
along with ‘‘error’’), ‘‘normalized’’ normal vectors are weighed by their 
norms. After the average normal vector is calculated, it is renormalized 
to the unit vector. 
The vector between points mi and gc: w ! 
i =Pmi  gc. 
The normal vector of rotation for points ai and bi: 
n ! 
i = 
v ! 
i 3 w ! 
i 
jv ! 
i jjw ! 
i j 
(crossproduct). 
Averaged (weighed) vector: 
n ! 
av = 
1 
n 
Xn 
i =1 
i jjw ! 
i j,n ! 
ðj v ! 
iÞ: 
Normal vector (unit vector): 
n ! 
unit = 
n ! 
av 
jn ! 
: 
av j 
The resultant averaged normal vector and center of rotation is 
visualized along with vectors vi within VMD by Tcl commands in the 
script. 
Accessible Surface Area Calculations 
The accessible surface areas of side chains are calculated using the 
GETAREA 1.1 (Fraczkiewicz and Braun, 1998) web server with 1.4 A ° 
probe radius (http://www.scsb.utmb.edu/cgi-bin/get_a_form.tcl) with 
additional atom type library and residue type entries. In our previous 
report, we used rabbit skeletal muscle actin. As ASA greatly depends 
on the side chain, the sequence of rabbit skeletal muscle actin was 
mapped on 1YAG using MODELLER 7v7(Sali and Blundell, 1993). 
This structure is designated as 1YAG (rabbit sequence) in the following 
sections. 
Computational Modeling 
Model construction started from a ‘‘template’’ structure of 1YAG (rab-bit 
sequence). The rotational movement toward the small domain was 
applied to the large domain (residues 138–332) at the pivot residue 138 
with the normal vector calculated in 1HLU (Chik et al., 1996). The rota-tional 
movement toward SD2 was applied as a combination of several 
movements calculated based on the analyses of several crystal struc-tures. 
The rotations were applied over four residues from specifically 
selected pivot points (residues 34–37 and 66–69) in order to prevent 
too much rotation to single bonds that can lead to improper bond an-gles. 
All rotational transformations of coordinates were performed us-ing 
Tcl scripts within VMD. Modeled structures were generated with 
combinations of different angles and normal vectors. ASAs of side 
chains were calculated for these structures in order to assess the pre-liminary 
models. The best model in the series of models by rigid-body 
movements was chosen according to the consistency with expected 
changes in ASAs of hydroxyl-radical probes. 
The model was then subjected to manual inspections/manipulations 
of side chains in the Swiss-PdbViewer (Guex and Peitsch, 1997) in or-der 
to better satisfy the hydroxyl-radical data. The rotamer of side 
chains was selected to maximize hydrogen-bond formations within 
the large cleft. As the large cleft between SD2 and SD4 was closed 
by moving the large domain relative to SD1, the nucleotide binding 
pocket was also affected and became significantly narrower than in 
the template structure 1YAG. The coordinates of the phosphates of 
ATP were manually adjusted within VMD using Tcl scripts by rotating 
the phosphate-oxygen bond and fit into the narrower cleft interactively 
in order to avoid steric clashes. The D loop was constructed manually 
in the Swiss-PdbViewer by rotating the phi and psi angles of the tem-plate 
structure. This process was guided mainly to allow M44 to reach 
the hydrophobic pocket in SD4 while at the same time avoiding im-proper 
angles. At this stage, the model structure was examined by 
PROCHECK (Laskowski et al., 1993) and Verify3D (Bowie et al., 1991) 
for model validity (Figure S1). The model coordinates were then passed 
to MODELLER 7v7 (Sali and Blundell, 1993) as a template for adjusting 
and correcting the stereochemical parameters. MODELLER was also 
used to adjust the geometry of the residues and metal ion in the nucle-otide 
binding pocket with manually assigned restraints according to the 
1RFQ-B (chain B) structure (Reutzel et al., 2004). After further manual 
inspections/adjustments in VMD and energy minimization within the 
Swiss-PdbViewer for certain side chains, the final structural model 
was subjected to PROCHECK and Verify3D for the assessment of 
structure validity. 
Visualization of Structures 
The structures were visualized using POV-Ray ray-tracer (Version 3.6) 
with output from VMD (Version 1.8.3) and manual editing of scene files. 
Some presentations were directly defined by POV-Ray scene. All op-erations 
were done on a Power Macintosh G5 with Mac OS X except 
for energy minimization within the Swiss-PdbViewer that was per-formed 
on a Windows 2000 workstation (OS X version is an a version 
and energy minimization is only partially implemented). 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 49 
Structure 
Model of Mg2+-ATP-Actin Monomer
Supplemental Data 
Supplemental Data include two figures and Supplemental Results 
and are available at http://www.structure.org/cgi/content/full/15/1/ 
39/DC1/. 
ACKNOWLEDGMENTS 
This work is supported in part by the Biomedical Technology Centers 
Program of the National Institutes for Biomedical Imaging and Bioen-gineering 
(P41-EB-01979). 
Received: July 26, 2006 
Revised: November 6, 2006 
Accepted: November 18, 2006 
Published: January 16, 2007 
REFERENCES 
Allingham, J.S., Tanaka, J., Marriott, G., and Rayment, I. (2004). Abso-lute 
stereochemistry of ulapualide A. Org. Lett. 6, 597–599. 
Belmont, L.D., Orlova, A., Drubin, D.G., and Egelman, E.H. (1999). A 
change in actin conformation associated with filament instability after 
Pi release. Proc. Natl. Acad. Sci. USA 96, 29–34. 
Bowie, J.U., Luthy, R., and Eisenberg, D. (1991). A method to identify 
protein sequences that fold into a known three-dimensional structure. 
Science 253, 164–170. 
Bubb, M.R., Govindasamy, L., Yarmola, E.G., Vorobiev, S.M., Almo, 
S.C., Somasundaram, T., Chapman, M.S., Agbandje-McKenna, M., 
and McKenna, R. (2002). Polylysine induces an antiparallel actin dimer 
that nucleates filament assembly: crystal structure at 3.5-A° resolution. 
J. Biol. Chem. 277, 20999–21006. 
Chen, X., Peng, J., Pedram, M., Swenson, C.A., and Rubenstein, P.A. 
(1995). The effect of the S14A mutation on the conformation and ther-mostability 
of Saccharomyces cerevisiae G-actin and its interaction 
° 
with adenine A nucleotides. J. Biol. Chem. 270, 11415–11423. 
Chik, J.K., Lindberg, U., and Schutt, C.E. (1996). The structure of an 
open state of b-actin at 2.65 resolution. J. Mol. Biol. 263, 607–623. 
De La Cruz, E.M., and Pollard, T.D. (1995). Nucleotide-free actin: 
stabilization by sucrose and nucleotide binding kinetics. Biochemistry 
34, 5452–5461. 
Dudev, T., Cowan, J.A., and Lim, C. (1999). Competitive binding in 
magnesium coordination chemistry: water versus ligands of biological 
interest. J. Am. Chem. Soc. 121, 7665–7673. 
Fraczkiewicz, R., and Braun, W. (1998). Exact and efficient analytical 
calculation of the accessible surface areas and their gradients for mac-romolecules. 
J. Comput. Chem. 19, 319–333. 
Frieden, C., and Patane, K. (1985). Differences in G-actin containing 
bound ATP or ADP: the Mg2+-induced conformational change requires 
ATP. Biochemistry 24, 4192–4196. 
Gershman, L.C., Selden, L.A., and Estes, J.E. (1991). High affinity diva-lent 
cation exchange on actin. Association rate measurements support 
the simple competitive model. J. Biol. Chem. 266, 76–82. 
Guan, J.Q., and Chance, M.R. (2005). Structural proteomics of macro-molecular 
assemblies using oxidative footprinting and mass spec-trometry. 
Trends Biochem. Sci. 30, 583–592. 
Guan, J.Q., Vorobiev, S., Almo, S.C., and Chance, M.R. (2002). Map-ping 
the G-actin binding surface of cofilin using synchrotron protein 
footprinting. Biochemistry 41, 5765–5775. 
Guan, J.Q., Almo, S.C., Reisler, E., and Chance, M.R. (2003). Structural 
reorganization of proteins revealed by radiolysis and mass spectrom-etry: 
G-actin solution structure is divalent cation dependent. Biochem-istry 
42, 11992–12000. 
Guan, J.Q., Almo, S.C., and Chance, M.R. (2004). Synchrotron radiol-ysis 
and mass spectrometry: a new approach to research on the actin 
cytoskeleton. Acc. Chem. Res. 37, 221–229. 
Guan, J.Q., Takamoto, K., Almo, S.C., Reisler, E., and Chance, M.R. 
(2005). Structure and dynamics of the actin filament. Biochemistry 
44, 3166–3175. 
Guex, N., and Peitsch, M.C. (1997). SWISS-MODEL and the Swiss- 
PdbViewer: an environment for comparative protein modeling. 
Electrophoresis 18, 2714–2723. 
Gupta, S., Mangel, W.F., McGrath, W.J., Perek, J.L., Lee, D.W., Taka-moto, 
K., and Chance, M.R. (2004). DNA binding provides a molecular 
strap activating the adenovirus proteinase. Mol. Cell. Proteomics 3, 
950–959. 
Holmes, K.C., Popp, D., Gebhard, W., and Kabsch, W. (1990). Atomic 
model of the actin filament. Nature 347, 44–49. 
Humphrey, W., Dalke, A., and Schulten, K. (1996). VMD: visual molec-ular 
dynamics. J. Mol. Graph. 14, 33–38. 
Joel, P.B., Fagnant, P.M., and Trybus, K.M. (2004). Expression of a 
nonpolymerizable actin mutant in Sf9 cells. Biochemistry 43, 11554– 
11559. 
Kabsch, W., Mannherz, H.G., and Suck, D. (1985). Three-dimensional 
structure of the complex of actin and DNase I at 4.5 A° 
resolution. 
EMBO J. 4, 2113–2118. 
Kabsch, W., Mannherz, H.G., Suck, D., Pai, E.F., and Holmes, K.C. 
(1990). Atomic structure of the actin:DNase I complex. Nature 347, 
34–44. 
Khaitlina, S.Y., and Strzelecka-Golaszewska, H. (2002). Role of the 
DNase-I-binding loop in dynamic properties of actin filament. Biophys. 
J. 82, 321–334. 
Kinosian, H.J., Selden, L.A., Estes, J.E., and Gershman, L.C. (1993). 
Nucleotide binding to actin. Cation dependence of nucleotide dissoci-ation 
and exchange rates. J. Biol. Chem. 268, 8683–8691. 
Kiselar, J.G., Janmey, P.A., Almo, S.C., and Chance, M.R. (2003a). 
Structural analysis of gelsolin using synchrotron protein footprinting. 
Mol. Cell. Proteomics 2, 1120–1132. 
Kiselar, J.G., Janmey, P.A., Almo, S.C., and Chance, M.R. (2003b). 
Visualizing the Ca2+-dependent activation of gelsolin by using syn-chrotron 
footprinting. Proc. Natl. Acad. Sci. USA 100, 3942–3947. 
Klenchin, V.A., Allingham, J.S., King, R., Tanaka, J., Marriott, G., and 
Rayment, I. (2003). Trisoxazole macrolide toxins mimic the binding of 
actin-capping proteins to actin. Nat. Struct. Biol. 10, 1058–1063. 
Laskowski, R.A., MacArthur, M.W., Moss, D.S., and Thornton, J.M. 
(1993). PROCHECK: a program to check the stereochemical quality 
of protein structures. J. Appl. Crystallogr. 26, 283–291. 
Liu, R., Guan, J.Q., Zak, O., Aisen, P., and Chance, M.R. (2003). 
Structural reorganization of the transferrin C-lobe and transferrin 
receptor upon complex formation: the C-lobe binds to the receptor 
helical domain. Biochemistry 42, 12447–12454. 
Lorenz, M., Popp, D., and Holmes, K.C. (1993). Refinement of the 
F-actin model against X-ray fiber diffraction data by the use of a di-rected 
mutation algorithm. J. Mol. Biol. 234, 826–836. 
Mannherz, H.G., Gooch, J., Way, M., Weeds, A.G., and McLaughlin, 
P.J. (1992). Crystallization of the complex of actin with gelsolin seg-ment 
1. J. Mol. Biol. 226, 899–901. 
McLaughlin, P.J., Gooch, J.T., Mannherz, H.G., and Weeds, A.G. 
(1993). Structure of gelsolin segment 1-actin complex and the mecha-nism 
of filament severing. Nature 364, 685–692. 
Moraczewska, J., Wawro, B., Seguro, K., and Strzelecka-Golaszew-ska, 
H. (1999). Divalent cation-, nucleotide-, and polymerization-dependent 
changes in the conformation of subdomain 2 of actin. 
Biophys. J. 77, 373–385. 
Morton, W.M., Ayscough, K.R., and McLaughlin, P.J. (2000). Latruncu-lin 
alters the actin-monomer subunit interface to prevent polymeriza-tion. 
Nat. Cell Biol. 2, 376–378. 
Muhlrad, A., Cheung, P., Phan, B.C., Miller, C., and Reisler, E. (1994). 
Dynamic properties of actin. Structural changes induced by beryllium 
fluoride. J. Biol. Chem. 269, 11852–11858. 
50 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 
Structure 
Model of Mg2+-ATP-Actin Monomer
Oda, T., Makino, K., Yamashita, I., Namba, K., and Maeda, Y. (2001). 
Distinct structural changes detected by X-ray fiber diffraction in stabi-lization 
of F-actin by lowering pH and increasing ionic strength. 
Biophys. J. 80, 841–851. 
Otterbein, L.R., Graceffa, P., and Dominguez, R. (2001). The crystal 
structure of uncomplexed actin in the ADP state. Science 293, 708– 
711. 
Otterbein, L.R., Cosio, C., Graceffa, P., and Dominguez, R. (2002). 
Crystal structures of the vitamin D-binding protein and its complex 
with actin: structural basis of the actin-scavenger system. Proc. Natl. 
Acad. Sci. USA 99, 8003–8008. 
Paavilainen, V.O., Bertling, E., Falck, S., and Lappalainen, P. (2004). 
Regulation of cytoskeletal dynamics by actin-monomer-binding pro-teins. 
Trends Cell Biol. 14, 386–394. 
Page, R., Lindberg, U., and Schutt, C.E. (1998). Domain motions in 
actin. J. Mol. Biol. 280, 463–474. 
Pecoraro, V.L., Hermes, J.D., and Cleland, W.W. (1984). Stability 
constants of Mg2+ and Cd2+ complexes of adenine nucleotides and 
thionucleotides and rate constants for formation and dissociation of 
MgATP and MgADP. Biochemistry 23, 5262–5271. 
Rashidzadeh, H., Khrapunov, S., Chance, M.R., and Brenowitz, M. 
(2003). Solution structure and interdomain interactions of the Saccha-romyces 
cerevisiae ‘‘TATA binding protein’’ (TBP) probed by radiolytic 
protein footprinting. Biochemistry 42, 3655–3665. 
Reutzel, R., Yoshioka, C., Govindasamy, L., Yarmola, E.G., Agbandje- 
McKenna, M., Bubb, M.R., and McKenna, R. (2004). Actin crystal dy-namics: 
structural implications for F-actin nucleation, polymerization, 
and branching mediated by the anti-parallel dimer. J. Struct. Biol. 
146, 291–301. 
Sali, A., and Blundell, T.L. (1993). Comparative protein modelling by 
satisfaction of spatial restraints. J. Mol. Biol. 234, 779–815. 
Schmid, M.F., Sherman, M.B., Matsudaira, P., and Chiu, W. (2004). 
Structure of the acrosomal bundle. Nature 431, 104–107. 
Schoenenberger, C.A., Bischler, N., Fahrenkrog, B., and Aebi, U. 
(2002). Actin’s propensity for dynamic filament patterning. FEBS 
Lett. 529, 27–33. 
Schu¨ ler, H. (2001). ATPase activity and conformational changes in the 
regulation of actin. Biochim. Biophys. Acta 1549, 137–147. 
Schu¨ ler, H., Korenbaum, E., Schutt, C.E., Lindberg, U., and Karlsson, 
R. (1999). Mutational analysis of Ser14 and Asp157 in the nucleotide-binding 
site of b-actin. Eur. J. Biochem. 265, 210–220. 
Schutt, C.E., Myslik, J.C., Rozycki, M.D., Goonesekere, N.C., and 
Lindberg, U. (1993). The structure of crystalline profilin-b-actin. Nature 
365, 810–816. 
Selden, L.A., Estes, J.E., and Gershman, L.C. (1989). High affinity diva-lent 
cation binding to actin. Effect of low affinity salt binding. J. Biol. 
Chem. 264, 9271–9277. 
Shao, Z., Shi, D., and Somlyo, A.V. (2000). Cryoatomic force micros-copy 
of filamentous actin. Biophys. J. 78, 950–958. 
Sharp, J.S., Guo, J.T., Uchiki, T., Xu, Y., Dealwis, C., and Hettich, R.L. 
(2005). Photochemical surface mapping of C14S-Sml1p for con-strained 
computational modeling of protein structure. Anal. Biochem. 
340, 201–212. 
Sheterline, P., and Sparrow, J.C. (1994). Actin. Protein Profile 1, 1–121. 
Shi, D., Somlyo, A.V., Somlyo, A.P., and Shao, Z. (2001). Visualizing fil-amentous 
actin on lipid bilayers by atomic force microscopy in solu-tion. 
J. Microsc. 201, 377–382. 
Strzelecka-Golaszewska, H., Moraczewska, J., Khaitlina, S.Y., and 
Mossakowska, M. (1993). Localization of the tightly bound divalent-cation- 
dependent and nucleotide-dependent conformation changes 
in G-actin using limited proteolytic digestion. Eur. J. Biochem. 211, 
731–742. 
Suck, D., Kabsch, W., and Mannherz, H.G. (1981). Three-dimensional 
structure of the complex of skeletal muscle actin and bovine pancre-atic 
DNAse I at 6-A° resolution. Proc. Natl. Acad. Sci. USA 78, 4319– 
4323. 
Swamy, N., Head, J.F., Weitz, D., and Ray, R. (2002). Biochemical and 
preliminary crystallographic characterization of the vitamin D sterol-and 
actin-binding by human vitamin D-binding protein. Arch. Biochem. 
Biophys. 402, 14–23. 
Takamoto, K., and Chance, M.R. (2006). Radiolytic protein footprinting 
with mass spectrometry to probe the structure of macromolecular 
complexes. Annu. Rev. Biophys. Biomol. Struct. 35, 251–276. 
Tirion, M.M., ben-Avraham, D., Lorenz, M., and Holmes, K.C. (1995). 
Normal modes as refinement parameters for the F-actin model. 
Biophys. J. 68, 5–12. 
Valentin-Ranc, C., and Carlier, M.F. (1991). Role of ATP-bound diva-lent 
metal ion in the conformation and function of actin. Comparison 
of Mg-ATP, Ca-ATP, and metal ion-free ATP-actin. J. Biol. Chem. 
266, 7668–7675. 
Verboven, C., Bogaerts, I., Waelkens, E., Rabijns, A., Van Baelen, H., 
Bouillon, R., and De Ranter, C. (2003). Actin-DBP: the perfect struc-tural 
fit? Acta Crystallogr. D Biol. Crystallogr. 59, 263–273. 
Vorobiev, S., Strokopytov, B., Drubin, D.G., Frieden, C., Ono, S., 
Condeelis, J., Rubenstein, P.A., and Almo, S.C. (2003). The structure 
of nonvertebrate actin: implications for the ATP hydrolytic mechanism. 
Proc. Natl. Acad. Sci. USA 100, 5760–5765. 
Wear, M.A., Schafer, D.A., and Cooper, J.A. (2000). Actin dynamics: 
assembly and disassembly of actin networks. Curr. Biol. 10, R891– 
R895. 
Wertman, K.F., Drubin, D.G., and Botstein, D. (1992). Systematic 
mutational analysis of the yeast ACT1 gene. Genetics 132, 337–350. 
Williams, R.P. (1970). Biochemistry of sodium, potassium, magnesium 
and calcium. Q. Rev. Chem. Soc. 24, 331–365. 
Winder, S.J. (2003). Structural insights into actin-binding, branching 
and bundling proteins. Curr. Opin. Cell Biol. 15, 14–22. 
Wriggers, W., and Schulten, K. (1997). Stability and dynamics of 
G-actin: back-door water diffusion and behavior of a subdomain 3/4 
loop. Biophys. J. 73, 624–639. 
Wu, Y., and Ma, J. (2004). Refinement of F-actin model against 
fiber diffraction data by long-range normal modes. Biophys. J. 86, 
116–124. 
Xu, G., and Chance, M.R. (2005). Radiolytic modification and reactivity 
of amino acid residues serving as structural probes for protein foot-printing. 
Anal. Chem. 77, 4549–4555. 
Yarmola, E.G., Somasundaram, T., Boring, T.A., Spector, I., and Bubb, 
M.R. (2000). Actin-latrunculin A structure and function. Differential 
modulation of actin-binding protein function by latrunculin A. J. Biol. 
Chem. 275, 28120–28127. 
Zimmerle, C.T., Patane, K., and Frieden, C. (1987). Divalent cation 
binding to the high- and low-affinity sites on G-actin. Biochemistry 
26, 6545–6552. 
Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 51 
Structure 
Model of Mg2+-ATP-Actin Monomer

More Related Content

Viewers also liked

Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
Keiji Takamoto
 
Viral Protein Structure Predictions - Consensus Strategy
Viral Protein Structure Predictions - Consensus StrategyViral Protein Structure Predictions - Consensus Strategy
Viral Protein Structure Predictions - Consensus Strategy
Keiji Takamoto
 
Eng 121 week 5 quiz
Eng 121 week 5 quizEng 121 week 5 quiz
Eng 121 week 5 quiz
opstalettin1981
 
Ant 101 entire course
Ant 101 entire courseAnt 101 entire course
Ant 101 entire course
opstalettin1981
 
Reume 2014Sales
Reume 2014SalesReume 2014Sales
Reume 2014Sales
Alexander Pender
 
Controlled Formation of Low-Volume Liquid Pillars between Plates with Lattic...
Controlled Formation of Low-Volume Liquid Pillars  between Plates with Lattic...Controlled Formation of Low-Volume Liquid Pillars  between Plates with Lattic...
Controlled Formation of Low-Volume Liquid Pillars between Plates with Lattic...
Keiji Takamoto
 
Mgt 426 week 5 discussion questions
Mgt 426 week 5 discussion questionsMgt 426 week 5 discussion questions
Mgt 426 week 5 discussion questions
opstalettin1981
 
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
Keiji Takamoto
 
Bono mercantil de valdez
Bono mercantil de valdezBono mercantil de valdez
Bono mercantil de valdez
oscaredg
 
Novel Strategy for Small Viral Protein Structure Predictions
Novel Strategy for Small Viral Protein Structure PredictionsNovel Strategy for Small Viral Protein Structure Predictions
Novel Strategy for Small Viral Protein Structure PredictionsKeiji Takamoto
 
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
Keiji Takamoto
 
Multipoint dilution and permeation gas calibrators
Multipoint dilution and permeation gas calibratorsMultipoint dilution and permeation gas calibrators
Multipoint dilution and permeation gas calibrators
Aymeric Beaupied
 
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
Keiji Takamoto
 
Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...
Keiji Takamoto
 

Viewers also liked (14)

Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
Carboxy-terminal Degradation of Peptides using Perfluoroacyl Anhydrides : C-T...
 
Viral Protein Structure Predictions - Consensus Strategy
Viral Protein Structure Predictions - Consensus StrategyViral Protein Structure Predictions - Consensus Strategy
Viral Protein Structure Predictions - Consensus Strategy
 
Eng 121 week 5 quiz
Eng 121 week 5 quizEng 121 week 5 quiz
Eng 121 week 5 quiz
 
Ant 101 entire course
Ant 101 entire courseAnt 101 entire course
Ant 101 entire course
 
Reume 2014Sales
Reume 2014SalesReume 2014Sales
Reume 2014Sales
 
Controlled Formation of Low-Volume Liquid Pillars between Plates with Lattic...
Controlled Formation of Low-Volume Liquid Pillars  between Plates with Lattic...Controlled Formation of Low-Volume Liquid Pillars  between Plates with Lattic...
Controlled Formation of Low-Volume Liquid Pillars between Plates with Lattic...
 
Mgt 426 week 5 discussion questions
Mgt 426 week 5 discussion questionsMgt 426 week 5 discussion questions
Mgt 426 week 5 discussion questions
 
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
Theoretical evaluation of shotgun proteomic analysis strategies; Peptide obse...
 
Bono mercantil de valdez
Bono mercantil de valdezBono mercantil de valdez
Bono mercantil de valdez
 
Novel Strategy for Small Viral Protein Structure Predictions
Novel Strategy for Small Viral Protein Structure PredictionsNovel Strategy for Small Viral Protein Structure Predictions
Novel Strategy for Small Viral Protein Structure Predictions
 
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
Bond-Specific Chemical Cleavages of Peptides & Proteins with Perfluoric Acid ...
 
Multipoint dilution and permeation gas calibrators
Multipoint dilution and permeation gas calibratorsMultipoint dilution and permeation gas calibrators
Multipoint dilution and permeation gas calibrators
 
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
Radiolytic Protein Footprinting with Mass Spectrometry to Probe the Structure...
 
Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...
 

Similar to Biochemical Implications of Three-Dimensional Model of Monomeric Actin Bound to Magnesium-Chelated ATP

Panushka UROP (1)
Panushka UROP (1)Panushka UROP (1)
Panushka UROP (1)
Joe Panushka
 
ScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
ScalingUptheShapeAnovelGrowthPatternOfGa49-70ClstrsScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
ScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
???? ?????
 
The effect of core destabilisation on the mechanical resistance of i27
The effect of core destabilisation on the mechanical resistance of i27The effect of core destabilisation on the mechanical resistance of i27
The effect of core destabilisation on the mechanical resistance of i27
John Clarkson
 
Investigation of interfacial properties at quartz alkane interfaces using mol...
Investigation of interfacial properties at quartz alkane interfaces using mol...Investigation of interfacial properties at quartz alkane interfaces using mol...
Investigation of interfacial properties at quartz alkane interfaces using mol...
IAEME Publication
 
State of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
State of the Art in the Characterization of Nano- and Atomic-Scale CatalystsState of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
State of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
Devika Laishram
 
Superconductivity in Al-substituted Ba8Si46 clathrates
Superconductivity in Al-substituted Ba8Si46 clathratesSuperconductivity in Al-substituted Ba8Si46 clathrates
Superconductivity in Al-substituted Ba8Si46 clathrates
Yang Li
 
First principles study on structural and electronic properties of re ag (re= ...
First principles study on structural and electronic properties of re ag (re= ...First principles study on structural and electronic properties of re ag (re= ...
First principles study on structural and electronic properties of re ag (re= ...
Alexander Decker
 
Main_Ms_JBNMR_Final_version
Main_Ms_JBNMR_Final_versionMain_Ms_JBNMR_Final_version
Main_Ms_JBNMR_Final_version
Abhilash Kannan
 
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
Stoyan Sarg Sargoytchev
 
structural-analysis-poly
structural-analysis-polystructural-analysis-poly
structural-analysis-poly
Edward Burt Driscoll
 
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
Deniz Karataş
 
Article2016
Article2016Article2016
Article2016
jabraoui
 
Oxygen Vacancy Conduction in Double Perovskites
Oxygen Vacancy Conduction in Double PerovskitesOxygen Vacancy Conduction in Double Perovskites
Oxygen Vacancy Conduction in Double Perovskites
Megha Patel
 
CBE_Symposium_Poster_Aparajita - sjp
CBE_Symposium_Poster_Aparajita - sjpCBE_Symposium_Poster_Aparajita - sjp
CBE_Symposium_Poster_Aparajita - sjp
Aparajita Dasgupta
 
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Keiji Takamoto
 
Limitations & lessons in the use of x ray structural information in drug design
Limitations & lessons in the use of x ray structural information in drug designLimitations & lessons in the use of x ray structural information in drug design
Limitations & lessons in the use of x ray structural information in drug design
Dilip Darade
 
Structural proteomics
Structural proteomicsStructural proteomics
Structural proteomics
Michiko Matsuo
 
Walton-2015-Nature
Walton-2015-NatureWalton-2015-Nature
Walton-2015-Nature
Joe Patterson
 
Poster presentat a les jornades doctorals de la UAB
Poster presentat a les jornades doctorals de la UABPoster presentat a les jornades doctorals de la UAB
Poster presentat a les jornades doctorals de la UAB
Elisabeth Ortega
 
Physical Models of LENR Processes Using the BSM-SG Atomic Models
Physical Models of LENR Processes Using the BSM-SG Atomic ModelsPhysical Models of LENR Processes Using the BSM-SG Atomic Models
Physical Models of LENR Processes Using the BSM-SG Atomic Models
Stoyan Sarg Sargoytchev
 

Similar to Biochemical Implications of Three-Dimensional Model of Monomeric Actin Bound to Magnesium-Chelated ATP (20)

Panushka UROP (1)
Panushka UROP (1)Panushka UROP (1)
Panushka UROP (1)
 
ScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
ScalingUptheShapeAnovelGrowthPatternOfGa49-70ClstrsScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
ScalingUptheShapeAnovelGrowthPatternOfGa49-70Clstrs
 
The effect of core destabilisation on the mechanical resistance of i27
The effect of core destabilisation on the mechanical resistance of i27The effect of core destabilisation on the mechanical resistance of i27
The effect of core destabilisation on the mechanical resistance of i27
 
Investigation of interfacial properties at quartz alkane interfaces using mol...
Investigation of interfacial properties at quartz alkane interfaces using mol...Investigation of interfacial properties at quartz alkane interfaces using mol...
Investigation of interfacial properties at quartz alkane interfaces using mol...
 
State of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
State of the Art in the Characterization of Nano- and Atomic-Scale CatalystsState of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
State of the Art in the Characterization of Nano- and Atomic-Scale Catalysts
 
Superconductivity in Al-substituted Ba8Si46 clathrates
Superconductivity in Al-substituted Ba8Si46 clathratesSuperconductivity in Al-substituted Ba8Si46 clathrates
Superconductivity in Al-substituted Ba8Si46 clathrates
 
First principles study on structural and electronic properties of re ag (re= ...
First principles study on structural and electronic properties of re ag (re= ...First principles study on structural and electronic properties of re ag (re= ...
First principles study on structural and electronic properties of re ag (re= ...
 
Main_Ms_JBNMR_Final_version
Main_Ms_JBNMR_Final_versionMain_Ms_JBNMR_Final_version
Main_Ms_JBNMR_Final_version
 
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
Graphical 3D Modeling of Molecules and Nanostructures in Sub-nanometer Scale ...
 
structural-analysis-poly
structural-analysis-polystructural-analysis-poly
structural-analysis-poly
 
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021
 
Article2016
Article2016Article2016
Article2016
 
Oxygen Vacancy Conduction in Double Perovskites
Oxygen Vacancy Conduction in Double PerovskitesOxygen Vacancy Conduction in Double Perovskites
Oxygen Vacancy Conduction in Double Perovskites
 
CBE_Symposium_Poster_Aparajita - sjp
CBE_Symposium_Poster_Aparajita - sjpCBE_Symposium_Poster_Aparajita - sjp
CBE_Symposium_Poster_Aparajita - sjp
 
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
 
Limitations & lessons in the use of x ray structural information in drug design
Limitations & lessons in the use of x ray structural information in drug designLimitations & lessons in the use of x ray structural information in drug design
Limitations & lessons in the use of x ray structural information in drug design
 
Structural proteomics
Structural proteomicsStructural proteomics
Structural proteomics
 
Walton-2015-Nature
Walton-2015-NatureWalton-2015-Nature
Walton-2015-Nature
 
Poster presentat a les jornades doctorals de la UAB
Poster presentat a les jornades doctorals de la UABPoster presentat a les jornades doctorals de la UAB
Poster presentat a les jornades doctorals de la UAB
 
Physical Models of LENR Processes Using the BSM-SG Atomic Models
Physical Models of LENR Processes Using the BSM-SG Atomic ModelsPhysical Models of LENR Processes Using the BSM-SG Atomic Models
Physical Models of LENR Processes Using the BSM-SG Atomic Models
 

More from Keiji Takamoto

Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
Keiji Takamoto
 
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
Keiji Takamoto
 
Payload Attachment Chemistry and Payload Design
Payload Attachment Chemistry and Payload DesignPayload Attachment Chemistry and Payload Design
Payload Attachment Chemistry and Payload Design
Keiji Takamoto
 
The derivation of ungapped global protein alignment score distributions - Part1
The derivation of ungapped global protein alignment score distributions - Part1The derivation of ungapped global protein alignment score distributions - Part1
The derivation of ungapped global protein alignment score distributions - Part1
Keiji Takamoto
 
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
Keiji Takamoto
 
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
Keiji Takamoto
 

More from Keiji Takamoto (6)

Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
Monovalent Cations Mediate Formation of Native Tertiary Structure of Tetrahym...
 
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
Statistical Equilibrium Wealth Distributions in an Exchange Economy with Stoc...
 
Payload Attachment Chemistry and Payload Design
Payload Attachment Chemistry and Payload DesignPayload Attachment Chemistry and Payload Design
Payload Attachment Chemistry and Payload Design
 
The derivation of ungapped global protein alignment score distributions - Part1
The derivation of ungapped global protein alignment score distributions - Part1The derivation of ungapped global protein alignment score distributions - Part1
The derivation of ungapped global protein alignment score distributions - Part1
 
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
C-terminal Sequencing of Protein : Novel Partial Acid Hydrolysis & Analysis b...
 
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
Semi-automated Single-band Peak-fitting Analysis of Hydroxyl Radical Nucleic ...
 

Biochemical Implications of Three-Dimensional Model of Monomeric Actin Bound to Magnesium-Chelated ATP

  • 1. Structure Article Biochemical Implications of a Three-Dimensional Model of Monomeric Actin Bound to Magnesium-Chelated ATP Keiji Takamoto,1,2,* J.K. Amisha Kamal,1,2 and Mark R. Chance1,2 1Case Center for Proteomics, Case Western Reserve University, 10090 Euclid Avenue, Cleveland, OH 44106, USA 2Lab address: http://casemed.case.edu/proteomics/ *Correspondence: keiji.takamoto@case.edu DOI 10.1016/j.str.2006.11.005 SUMMARY Actin structure is of intense interest in biology due to its importance in cell function and motility mediated by the spatial and temporal regulation of actin monomer-filament intercon-versions in a wide range of developmental and disease states. Despite this interest, the struc-ture of many functionally important actin forms has eluded high-resolution analysis. Due to the propensity of actin monomers to assemble into filaments structural analysis of Mg-bound actin monomers has proven difficult, whereas high-resolution structures of actin with a diverse ar-ray of ligands that preclude polymerization have been quite successful. In this work, we provide a high-resolution structural model of the Mg- ATP-actin monomer using a combination of computational methods and experimental foot-printing data that we have previously published. The key conclusion of this study is that the structure of the nucleotide binding cleft defined by subdomains 2 and 4 is essentially closed, with specific contacts between two subdo-mains predicted by the data. INTRODUCTION Actin is a ubiquitous and important protein in eukaryotes and is extremely well conserved from yeast to man. Actin binding of nucleotides and its interactions with other pro-teins in the cell control the spatial and temporal assembly and disassembly of the cytoskeletal network; the careful regulation of this network has a profound influence on cell motility (Paavilainen et al., 2004; Schoenenberger et al., 2002; Wear et al., 2000; Winder, 2003). Even the nature of the metal ion bound to the nucleotide with the actin structure can profoundly alter the ability of actin monomers to assemble into filaments. In spite of its bio-logical importance, the structure of the Mg2+-ATP-bound form of the actin monomer (called G-actin) is not known. The determination of the high-resolution structure of Mg2+-G-actin is hampered by the propensity of actin to polymerize. Generation of crystals suitable for X-ray dif-fraction as well as NMR studies typically requires higher concentrations than the critical concentration for actin polymerization for the Mg2+-nucleotide-bound forms of the actin monomer. The high-resolution crystal structures of G-actin, whose richness and depth are a nearly unique resource for this investigation, never the less have either Ca2+-nucleotide bound to actin or, if Mg2+-bound nucleo-tide is used, actin is cocrystallized with ligands such as DNase I (Kabsch et al., 1985, 1990; Suck et al., 1981), gel-solin segment 1 (Mannherz et al., 1992; Vorobiev et al., 2003), vitamin D binding protein (Otterbein et al., 2002; Swamy et al., 2002; Verboven et al., 2003), or macrolides (Allingham et al., 2004; Klenchin et al., 2003; Morton et al., 2000; Reutzel et al., 2004; Yarmola et al., 2000). These li-gands typically prevent polymerization. Although struc-tures of Ca2+- and Mg2+-G-actin bound to ligands are overall similar in many respects, significant differences in structure and function between these species in the absence of other bound ligands in solution have been consistently reported, including biochemical/biophysical analyses such as limited proteolysis (Chen et al., 1995; Strzelecka-Golaszewska et al., 1993), fluorescence stud-ies (Frieden and Patane, 1985; Moraczewska et al., 1999; Selden et al., 1989; Valentin-Ranc and Carlier, 1991; Zim-merle et al., 1987), and molecular dynamics simulations (Wriggers and Schulten, 1997). Recently, we have investigated the solution structure of Mg2+-ATP-actin using hydroxyl-radical footprinting and mass spectrometry (MS); these experiments have in-cluded a comparison of Ca2+-ATP-actin in the presence and absence of gelsolin segment 1 (GS1) (Guan et al., 2003). The measured side-chain solvent accessibilities of Ca2+-ATP-actin in the absence of GS1 are very similar to that in the presence of GS1 overall and consistent with the accessible surface area (ASA) calculated from solved crystal structures. In contrast, Mg2+-ATP-actin in the absence of GS1 is quite different in its side-chain surface accessibilities, particularly in subdomains (SD) 2 and 4. These differences are reversed in the presence of GS1, indicating the structure of the Mg2+-ATP-actin/GS1 com-plex in solution is consistent with crystal structure data. Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 39
  • 2. Hydroxyl-radical footprinting and MS have been proven to be powerful tools for probing protein structure (Guan et al., 2003; Rashidzadeh et al., 2003) and conformational change (Kiselar et al., 2003a, 2003b), as well as for probing protein-ligand (Gupta et al., 2004) and protein-protein interactions (Guan et al., 2002, 2004; Liu et al., 2003). As the experimental details of the technique have matured and become reliable (Guan and Chance, 2005; Takamoto and Chance, 2006), we and others have attempted to use the data in conjunction with comparative modeling tech-niques to provide unique structural models (Gupta et al., 2004; Sharp et al., 2005). In this study, we extend our previous work (Guan et al., 2003) with visualization of our descriptive prediction by generating an atomic model of the Mg2+-ATP-actin monomer structure using footprint-ing and crystallographic data. The computational strategy uses rigid-body rotations and translations of actin subdo-mains, primarily guided by known subdomain rearrange-ments from the range of actin crystallographic structures, in order to generate a structure consistent with surface accessibility data predicted by footprinting. In addition to an atomic model of the Mg2+-ATP-actin monomer, we also propose a specific mechanism for cleft closure mediated by changes in metal-ion coordination. RESULTS Rigid-Body Rearrangements of Actin Domain/Subdomains Derived from an Actin Crystallographic ‘‘Database’’ The wealth of actin structures available in the literature provides a nearly unique opportunity for examining the range of conformations accessible to actin in its various ligand-bound states. Actin is composed of two domains that are termed the large and small domains, and each do-main is composed of two subdomains (Figures 1A and 1B). Table 1 provides a ‘‘database’’ of structures that encompass a number of relative arrangements of the var-ious actin subdomains. These are the major structures that are used in this paper to provide an analysis of the range of motions of the actin subdomains. Our approach in this section is to survey these structures and determine patterns of subdomain motions. Our previous footprinting data (Guan et al., 2003) indicated that the large cleft (nucleotide binding cleft) be-tween SD2 and SD4 is in a more closed configuration for monomeric actin bound to Mg2+-ATP compared to the Ca2+-ATP form. We surveyed the available crystal struc-tures to ascertain the range of relative motions of the large cleft. Most actin crystal structures (Table 1) are very similar, with an overall backbone rmsd of 0.8 A° for 13 structures (including Protein Data Bank ID codes 1RFQ-A and 1RFQ-B, 1IJJ-A and 1IJJ-B, 1NM1, 1D4X, 1EQY-A, 1NWK, 1QZ6, 1YAG, 1MA9, 1AQK, and 1NLV). However, there are some crystal structures that show significant de-viations from the majority of solved structures. One of the most prominent examples is crystal structure 1HLU (Chik et al., 1996), which has profilin bound to the small cleft Figure 1. Domain Rigid-Body Movements Observed in Actin Crystal Structures (A) The movement of the large domain relative to SD1 observed in crys-tal structure 1HLU. The models shown in tin and silver are 1YAG and 1HLU, respectively. The movements of Ca atoms are indicated by arrows. The colors of arrows indicate the size of movements ranging from blue (small) to orange (large). The center of rotational movement and normal vector are shown by the yellow sphere and long arrow. (B) Schematic representation of large-domain movement. (C) The movements of SD2 relative to subdomain 1. The white arrows indicate deviations of the G63 Ca atom position from 1YAG. between SD1 and SD3; this structure exhibits a ‘‘super-open’’ nucleotide cleft, with a large movement of the ATP/cation complex within the cleft. This structure can be related to other more ‘‘typical’’ structures through domain movements by rotational/sheer transformation (Chik et al., 1996; Page et al., 1998). A second example is 1RFQ 40 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 3. Table 1. List of Crystal Structures Used for Analyses Protein Data Bank ID Code Species Cocrystal Cofactors Resolution Comments Crystal Structures of Interest 1YAG Yeast GS1b Mg2+/ATP 1.90 ‘‘Template’’ molecule (with 1RFQ-Ba Rabbit LARc Mg2+/ATP 3.00 Closed large cleft with altered 1QZ5 Rabbit KABd Ca2+/ATP 1.45 Closed large cleft 1HLU Bovine Profilin Ca2+/ATP 2.65 Large cleft in ‘‘super-open’’ Crystal Structures with Complete D Loop 1J6Z Rabbit TMR Ca2+/ATP 1.54 TMR affects C-terminal structure? 1YVN Yeast GS1 Mg2+/ATP 2.10 V159N mutant of 1YAG 2BTF Bovine Profilin Sr2+/ATP 2.55 With Sr2+. b-actin 1ATN Rabbit DNase I Ca2+/ATP 2.80 D loop interacts with DNase I 1IJJ Rabbit LAR Mg2+/ATP 2.85 D loop that interacts with another 1LCU Rabbit LAR Ca2+/ATP 3.50 Crystal Structures with Partial D Loop 1D4X C. elegans GS1b Ca2+/ATP 1.75 1C0G Chimeric GS1b Ca2+/ATP 2.00 Chimeric, Q228K/T229A/A230Y/E360H 1MDU Chicken GS1b Ca2+/ATP 2.20 Actin-trimer 1C0F Chimeric GS1b Ca2+/ATP 2.40 Chimeric Dictyostelium/Tetrahymena 1DEJ Chimeric GSe Mg2+/ATP 2.40 Chimeric, Q228K/T229A/A230Y/ 1H1V Human GSe Ca2+/ATP 3.00 Important crystal structures of monomeric actin. The first set is used for modeling process or strategy; the second set is actin struc-tures with complete D-loop backbone coordinates; and the last set is a partial but relatively better coverage of D-loop backbone coordinates. Except for 1RFQ, the interasymmetric unit interactions between their D loop and the other molecule are unknown. 1RFQ does not have contact with other molecules. a 1RFQ-B, chain B of asymmetric unit in 1RFQ. b Gelsolin segment 1. c Latrunculin A. d Kabiramide C. e Gelsolin. (Reutzel et al., 2004), which has two chains in the asym-metric unit whose structures are quite different from each other. Chain A exhibits a ‘‘standard’’ cleft structure similar to those of the vast majority of solved structures. However, chain B (1RFQ-B) is interesting as it shows a ‘‘closed’’ form of the cleft, with very different geometries for the residues inside the cleft. The closed form in chain B of the crystal structure exemplifies a database entry that qualitatively satisfies an important aspect of our footprint-ing data, that is, it provides an example of how relative subdomain motions can result in a closed nucleotide cleft. Crystal structure 1QZ5 (Klenchin et al., 2003) shows a closed large cleft as well. In this structure, the geometry in-side the large cleft is almost identical to 1YAG (although it has a Ca2+ ion instead of an Mg2+ ion). To achieve cleft closure compared to canonical actin forms, SD2 moves complete D loop) side-chain geometries state, b-actin chain’s small cleft A231K/S232E/E360H toward SD4 without tilting (unlike 1RFQ-B), minimizing ste-ric conflicts between the turn in SD2 (residues 62–65) and the top of the helix in SD4 (residues 200–205). Although our hydroxyl-radical footprinting results indicate more signifi-cant closure of the cleft than observed in these structures, these structures can serve as ‘‘qualitative’’ templates for understanding the Mg2+ form of G-actin in solution. The B factors of the atoms in SD2 of 1QZ5 are relatively high, indicating the relative mobility of SD2. This suggests the likelihood of high relative mobility of SD2 (not only the D loop but also the SD2 core) in solution (see the Supple-mental Data available with this article online). Vector Analysis of Subdomain Relative Positions The structure 1YAG is used as our standard structural template having high-resolution (1.90 A° ) and well-defined Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 41 Structure Model of Mg2+-ATP-Actin Monomer
  • 4. waters within the nucleotide binding pocket. The struc-tures of Protein Data Bank ID codes 1HLU (Chik et al., 1996), 1QZ5 (Klenchin et al., 2003), 1RFQ-B (Reutzel et al., 2004), 1J6Z (Otterbein et al., 2001), 1ATN (Kabsch et al., 1990), 1S22 (Allingham et al., 2004), 1EQY-A (Mannherz et al., 1992), 1IJJ-A (Bubb et al., 2002), 1MA9 (Verboven et al., 2003), 1KXP (Otterbein et al., 2002), and 2BTF (Schutt et al., 1993) were compared with 1YAG, and differences between Ca/backbone coordinates were visualized using Tcl scripts for visual molecular dynamics (VMD). The script also calculates the average center position and rotational normal vector of movement of Ca/backbone coordinates. Figure 1A shows the relative movements (vector) of the large domain with respect to SD1 in the comparison of 1YAG (the standard) and 1HLU (most open nucleotide cleft structure). The arrows indicate the Ca movements (arrows from coordinates in 1YAG to 1HLU). As the move-ments are relative, we chose to show the large-domain movement relative to SD1 in this representation because SD2 also shows significant movement against SD1. As seen in Figure 1B, this movement is 8.9 rotational (hinge) movement that has a center position in the junction be-tween SD1 and SD3 (around Q137-A138). The axis of ro-tation goes from the front to the back side of the molecule (where the front side is defined as the surface of the actin molecule that exposes ATP). Figure 1B illustrates this movement (the rotational angle is exaggerated for clarity). A similar movement is reported in previous analyses and refinement of F-actin fiber X-ray data including alterations of the nucleotide binding pocket (Lorenz et al., 1993; Tirion et al., 1995). The SD2 subdomain is extremely mobile even in crystal structures. Figure 1C compares 11 SD2 structures with that of 1YAG; SD2 is observed in many orientations rela-tive to SD1. The directions of rotation of these crystal structures (relative to 1YAG) are shown in Figure 1C by arrows. All of these differences can be mediated by hinge movements having pivot points near P32-S33 and Y69-P70. Intracleft Interactions There are a number of interactions between the ATP/ cation complex and the protein within this cleft that are important for understanding large-cleft structure and the possible mechanisms of the structural changes. The sub-domains are not strongly connected by hydrogen-bond networks or other forms of interactions; the majority of the interactions between subdomains are mediated through a hydrogen-bond network with the ATP molecule and with waters in the pocket of the cleft. SD1 has interac-tions with phosphate oxygen atoms of ATP through resi-dues S14, G15, L/M16, and K18 (Figure 2A). Connections between SD1 and SD2 are mediated only by the hydrogen bonds between S14 and G74 (Figure 2B) (Chen et al., 1995; Kabsch et al., 1990; Schu¨ ler, 2001). This link is severed in 1RFQ-B due to subdomain rearrangements. Although the metal ion is located near the interface of sub-domains 1 and 2, it only interacts with SD1 by hydrogen Figure 2. Important Interactions Involving the ATP Molecule Brown and silver colors are 1YAG and 1FRQ, respectively. Hydrogen bonds and coordination bonds are shown by broken lines. (A) N-terminal b strand and ATP-metal interactions. (B) Hydrogen bonds that connect subdomains 1 and 2 and ATP. (C) Interactions with the Mg2+ ion. Water molecules observed in 1YAG are shown as blue balls. bonds through waters (no inner-sphere coordination) and no interaction is formed with SD2 residues. The interactions between SD3 and SD4 are also mainly mediated through ATP. In addition, there is a poorly packed hydrophobic cluster at the interface of the two 42 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 5. Table 2. Probe Residue Solvent Accessibilities from Crystal Structures and Experimentally Determined Hydroxyl-Radical Reactivities Crystal Structures 1ATN 1YAG 1RFQB 1RDW Footprinting Data Ca Mg Mg Mg Solution G-Actin DNase I GS1 LAR LAR Ca Mg subdomains. The adenosine moiety of ATP is located in the pocket between the two subdomains, with hydrogen bond and van der Waals contacts with surrounding resi-dues. Thus, SD3, SD4, and ATP form an interdependent set of interactions. Metal-Ion Coordination In most nonactin protein structures with ATP bound, mag-nesium ion coordinates to at least one residue from the protein (14 structures out of 16 nonredundant structures examined; data not shown; Dudev et al., 1999). In the case of actin, the magnesium ion (or calcium ion) forms in-ner- sphere coordination bonds with phosphate oxygens and water molecules, but not side-chain oxygen atoms, with a coordination number of 6 (or 7 for Ca2+). A notice-able difference between 1YAG (or other structures) and 1RFQ-B is that residue atoms Q137:OE1 and D11:OD1, OD2 are closer to the Mg2+ ion in 1RFQ-B (3.0, 3.5, and 4.0 A ° ) compared to 1YAG (4.4, 4.3, and 4.2 A ° , respec-tively). Figure 2C shows the difference in the geometries surrounding the magnesium ion. Unfortunately for 1RFQ, at a resolution of 3.0 A ° , ordered water molecules are not observed within the nucleotide binding cleft. However, some water molecules observed in 1YAG must be radi-cally reorganized in 1RFQ-B, as these water molecules would clash with side-chain oxygen atoms (Figure 2C, pink side-chain oxygen atoms overlapping with blue water oxygen atoms). Based on these data, we used the geom-etry inside the large cleft from 1RFQ-B as a template for our structural modeling of the nucleotide binding cleft. Comparison of Crystallographic Data and Hydroxyl-Radical Footprinting Data Table 2 summarizes the calculated ASAs and rate con-stants of modification of actin peptides analyzed by hydroxyl-radical footprinting (Guan et al., 2003). The hydroxyl-radical footprinting data for Ca2+-ATP-G-actin are generally consistent with the calculated ASA, that is, solvent-accessible, reactive residues show oxida-tion and inaccessible residues are not appreciably oxi-dized. However, a number of probe residues in SD2 and SD4 and at the C terminus (SD1) show decreased rates of modification for Mg2+-ATP-G-actin where the solvent accessibilities for the structure are similar to those for Ca2+-ATP-G-actin. These sites that show protections from oxidation for Mg- versus Ca-actin are illustrated in Residue Number Residue Side-chain ASA (A° 2) Side-chain ASA (A° 2) Side-chain ASA (A° 2) Side-chain ASA (A° 2) Rate constant (s1) Rate constant (s1) Peptide (Residue Numbers) 21 Phe 34.32 23.87 37.13 34.39 0.65/0.70 0.32/0.43 19–28 11–23 44 Met 79.35 (16.31)a 137.63 (35.96)a N.D.b N.D.b 33 20 40–50 47 Met 104.58 (30.94)a 113.37 (24.81)a N.D.b N.D.b 33 20 53 Tyr 14.26 6.60 64.67c 55.88c 0.69 0.32 51–61 67 Leu 28.70 22.16 25.46 40.25 0.40 0.08 63–68 69 Tyr 80.46 98.57 68.54 66.81 1.1 0.18 58–72 200 Phe 0.36 3.61 10.79 5.50 1.10/0.83 0.26/0.24 197–206 196–207 201 Val 82.01 65.45 108.42 101.42 1.10/0.83 0.26/0.24 202 Thr 81.61 72.45 43.79 73.26 1.10/0.83 0.26/0.24 243 Pro 43.26 63.97 72.23 68.46 0.69/0.75 0.26/0.30 239–254 242–253 362 Tyr 12.59 9.05 15.18 11.75 0.72 0.38 362–372 367 Pro 45.75 40.74 39.49 41.88 0.72 0.38 371 His 27.79 26.53 34.71 25.89 0.72 0.38 374 Cys N.D. 3.53 49.40 24.29 8.5 3.6 363–375 375 Phe N.D. 157.72 77.56 44.38 N.D.d N.D.d a Values in parentheses are ASAs for the sulfur atom in the side chain. b These crystals lacking observed D-loop structures. c These values are affected by the lack of a D loop (larger than expected). d The rate constant in the peptide is dominated by C374 and thus could not be determined. Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 43 Structure Model of Mg2+-ATP-Actin Monomer
  • 6. Figure 3. For example, probe residues in the D loop are modestly protected from oxidative modifications (40% reduction in rate constant), whereas residues 200–202 and 243 also experience strong reductions in modification rate (75% and 60%, respectively). L67 and Y69 lo-cated inside the large cleft (Figure 3) experience an 80% reduction in modification rates. Consistent with these find-ings are limited proteolysis data, where cleavage between K68 and Y69 is almost completely suppressed in Mg2+- G-actin compared to Ca2+-G-actin (Chen et al., 1995; Strzelecka-Golaszewska et al., 1993). A closure of the large cleft is consistent with these data. Modeling Strategy: Overall Structural Considerations In the previous section, we analyzed possible relative movements of domains/subdomains. In light of these analyses, we can understand the cleft structures in each case. DNase I binding prevents large-cleft closure by prohibiting the movements of SD2 and the large domain by interacting with both of them (Kabsch et al., 1990). Latrunculin A binds between SD2 and SD4 (Bubb et al., 2002; Reutzel et al., 2004); this prevents movements of those subdomains. Gelsolin segment 1 (Mannherz et al., 1992; McLaughlin et al., 1993) and vitamin D binding pro-tein (Otterbein et al., 2002) prevent the domain movement between SD1 and SD3 blocking corresponding large-domain movements. Some macrolides (kabiramide C, jaspisamide A, and ulapualide A; Allingham et al., 2004; Klenchin et al., 2003) bind to the small cleft and block its movement as well. This is summarized in Figure 4A. It is important to note that macrolides are small enough to fit into small or large clefts and do not seem to prevent poly-merization by steric hindrance as opposed to actin binding proteins that result in steric blockage. It appears that these macrolides interfere with polymerization by prevent-ing domain/subdomain motions. The analysis of the various crystal structures and the footprinting data (Guan et al., 2003) provides significant clues about how the structure of Mg2+-G-actin differs from that of Ca2+-G-actin. First, the footprinting data indi-cate that the large cleft between SD2 and SD4 is almost completely closed. Second, the analyses of the various crystal structures indicate that movement of the SD2 core is a rigid-body movement. The movement of SD2 toward the large domain alone cannot explain the protections of Figure 3. Analyses of Hydroxyl-Radical Reactivity and Struc-tures The protection sites in Mg2+-G-actin (compared with Ca2+-G-actin) are displayed with residues. Colors are coded as blue (strong protection) to red (enhancement) by reactivity changes. Figure 4. Schematic Representation of Domain/Subdomain Movements (A) The effects on subdomain movements by ligand binding. Actin binding proteins (top) bind to actin and prevent polymerization, possibly by steric hindrance or interference of domain movements. Macrolides are small enough to fit into small or large clefts and interfere with the domain movements. They may prevent polymerization by this mechanism. (B) The proposed mechanism for large-cleft closure. SD2 movement alone cannot explain the complete closure of the large cleft (top right). The partial closure by large-domain movement brings SD4 and SD2 close enough to allow interactions between the two subdomains, re-sulting in complete cleft closure. 44 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 7. both L67 and Y69. Y69 is located deep within the large cleft and cannot be buried by SD2 rigid-body movement. In ad-dition, neither M44 nor M47 can reach SD4 to form pair-wise interactions to protect P243 and M44/47. This sug-gests that both the large domain and SD2 need to move as rigid bodies relative to SD1 in order to explain protec-tions observed in SD2 and SD4 (Figure 4B). The hydrogen-bond network among phosphate, S14, and G74 is the only one actually connecting SD2 to SD1 and is observed in almost all structures (Schu¨ ler, 2001). This linkage is severed in the 1RFQ-B structure. We spec-ulate that latrunculin A prevents the movement of SD2 fur-ther from the observed structure although the geometry inside the large cleft is changed. As a result, the hydrogen bonds are broken. This hydrogen-bond linkage was main-tained in the model structure. We constructed a series of structures with combinations of different rotational angles and normal vectors for SD2 in order to determine the best-fit model for the hydroxyl-radical data. Regions Excluded from Modeling Although there is evidence for structural variation in the C-terminal region (Y362 to F375 changes solvent accessi-bility in hydroxyl-radical footprinting; Guan et al., 2003) and other data (Frieden and Patane, 1985; Strzelecka- Golaszewska et al., 1993; Valentin-Ranc and Carlier, 1991; Zimmerle et al., 1987), we do not have clear evidence to support the specific structural differences. Thus, we have not modeled the structural differences in this region. F21 experiences an 50% reduction in modification rate in Mg2+-G-actin compared to Ca2+-G-actin, but this change cannot be explained with our modeling strategy. Our speculation is that the C-terminal (residues 336–375) and N-terminal (residues 1–33) regions interacting with SD1 are also affected by the movements of SD2 (con-nected to the N-terminal region) and the large domain (connected to the C-terminal region) and mediate these additional structural differences between the two forms. Model of the Mg2+-ATP-Actin Monomer In general, the hydroxyl-radical footprinting data are very consistent with known crystal structures. Solvent-accessible surface areas and rate constants of oxidation on side chains in known structures are in good correlation (Guan et al., 2004; Guan and Chance, 2005; Takamoto and Chance, 2006; Xu and Chance, 2005). This is not sur-prising, as the size of the hydroxyl radical is very close to that of a water molecule. Although it is not easy to com-pare rate constants among different side chains (such as Leu and Phe), it is very quantitative and reliable for the same residue in different conformational states. The site of oxidation is determined by MS/MS analyses that are well established and routine. With the use of different proteases, protein sequence coverage by MS analysis is usually 80%–90% and most of the probe residues can be detected. In our previous study, we covered 90% of the actin sequence with trypsin and Asp-N proteases. Thus, it is important to note that we have used high-quality data for our modeling. Figure 5. The Model (A) Overlaid structure of the model (silver) and template structure 1YAG (brown). The residues that experience protections are displayed as sticks and bubbles. Structures are aligned by subdomain 1. (B) The difference between the model and template structure 1YAG. The arrows indicate the differences of backbone atom positions. The sphere and long arrow indicate the center of rotation and its normal vectors for rigid-body movements (green for large domain and yellow for subdomain 2). Figure 5A shows the final model and the template struc-ture for Mg2+-G-actin. The residues indicated to change conformation in the footprinting experiments are shown in stick-and-bubble format for both structures. The struc-tures are aligned at SD1 in order to show both large-domain and SD2 movements. The large domain rotates toward the small domain and SD2 rotates to the front (as defined previously) and toward the large domain. Figure 5B show the backbone movement analyzed by VMD/Tcl scripts. Figures 6A and 6B show a magnified view of the SD2/SD4 interface in the model. In order to make contact between D-loop residue M44 and the hydropho-bic pocket in SD4 around protected residue P243, SD2 was moved toward the front side along with its movement toward the large domain. As shown in Figure 6C, the M44 side chain is inserted into the hydrophobic pocket and forms a contact with the P243 side chain. M47 cannot form this interaction with the hydrophobic pocket without Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 45 Structure Model of Mg2+-ATP-Actin Monomer
  • 8. forming improper bond angles in the context of these rigid-body movements. The two residues (M44 and P243) are simultaneously protected from solvent expo-sure through the ‘‘pairwise’’ interactions. On the other hand, M47 is exposed to solvent almost completely. This makes predictions that could be tested in future foot-printing experiments using high-resolution tandem MS. In systematic mutational analysis experiments on yeast actin, the double-mutation A204E/P243K abolishes poly-merization (Joel et al., 2004). These mutations introduce bulky and charged residues at the place where contacts are formed in our model, likely blocking this conforma-tional change. However, because these residues are also involved in the intermolecular contacts of the F-actin model, it is also possible that the mutations may hinder F-actin filament assembly. The closure of the large cleft in our model allows the formation of a hydrogen-bond network between SD2 and SD4 (Figure 6D). In our model, seven new hydrogen Figure 6. Magnified Views of Interface between SD2 and SD4 (A) Template structure 1YAG. (B) The model. The residues that experience protection are colored blue (in SD2) or red (SD4). (C) The interaction between the SD4 hydropho-bic pocket and D-loop M44 side chain. The side chain of P243 is colored green while the side chain of M44 is displayed as bubbles (sul-fur is colored yellow). (D) Newly formed hydrogen-bond network be-tween SD2 and SD4 in the model. Residues are silver in SD2 and brown in SD4. The possible p-stacking is indicated by surface models on residues Y69 and R183. bonds can be formed between SD2 and SD4, as indicated in the figure. Also, Y69 and R183 are in good locations for p-stacking interactions (this interaction may not be critical to form the structure, as the R183A/D184A mutant has no significant change in phenotype). In a comparison of ASA between 1YAG (rabbit sequence) and the model (Figure 7), M47’s side-chain ASA increased by almost 50%. On the other hand, M44 experiences almost complete burial of its side chain. If we take into account the extremely dynamic nature of the D loop, it is reasonable to assume that both M44 and M47 are much more accessible than the structure observed in 1YAG. Overall, the accessibility of M44/47 seen in footprinting is in good agreement with the model. The ASA changes in Y53 can be explained in the same way. The D-loop structure is highly mobile, and in the crystal structure the side chain of Y53 in 1YAG is covered by the D loop and appears buried, but this does not reflect the ensemble average experienced by the fluc-tuating structure. In the model, although the ASA of the Figure 7. Changes in Accessible Surface Areas of Probe Residues The red bars are ASAs of 1YAG (rabbit) and the blue bars are differences between 1YAG and the model. Negative values indicate less ex-posed in the model. 46 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 9. Y53 side chain increases about 25A° 2, the D-loop structure should be stiff through its formation of an interaction with SD4. Thus, if the dynamic nature of structural changes in SD2 is considered, this residue can experience a lower reactivity consistent with the structural changes seen in the model. DISCUSSION Possible Mechanism of Cleft Closure Our model is based on rigid-body movements of domains and subdomains observed in solved crystal structures. We applied changes strictly to the backbone angles that are involved in domain movements because we do not have any data that indicate there is large-scale reorgani-zation of structure other than domain movements. The large-domain movement relative to SD1 narrowed the large cleft including the nucleotide binding pocket as pro-posed in previous reports of F-actin structure (Belmont et al., 1999; Lorenz et al., 1993; Tirion et al., 1995). In the crystal structures except 1RFQ-B, the metal ion is coordi-nated only by water molecules and phosphate oxygen atoms. The narrowed cleft results in a closer proximity of side-chain oxygen atoms (D11:OD1, OD2, and Q137:OE1) to metal ions that coordinate phosphate oxygen atoms. In our model, the side chains of these residues form inner-sphere coordination to Mg2+. AsMg2+ ions strongly prefer to coordinate to oxygen atoms of side chains instead of those of water molecules with a likely gain of free energy (Dudev et al., 1999), it is reasonable to assume that ex-change of inner-sphere coordinations between water mol-ecules and oxygen atoms in D11 and Q137 occurs spon-taneously, as thermal fluctuations bring side-chain oxygen atoms close enough to the metal ions mediating the ex-change reaction. This can then be a driving force for the movement of the large domain toward SD1. Interestingly, there is no strong interaction between SD1 and SD3—no hydrogen-bond network nor hydrophobic interactions. Thus, there is no discernable preference (or penalty) for these subdomain motions compared to the structures ob-served in the crystals. Once the large domain is locked in the closed form by formation of inner-sphere coordination between side-chain oxygen atoms and the metal ion, SD2 and SD4 are closer to each other and in the range where a detailed hydrogen-bond network (and possibly p-stack-ing) can form between them. This hydrogen-bond network and p-stacking stabilize and lock a ‘‘super-closed’’ form of the large cleft. The invariant residue Y69 appears to be a key residue for the interaction between SD2 and SD4. It possibly forms hydrogen bonds with the residues in SD4 at the bottom of the large cleft and acts as a latch. Mutations of residues forming interactions between SD2 and SD4 (Figure 6D) are reportedly lethal in yeast (Sheter-line and Sparrow, 1994; Wertman et al., 1992) (R62, R206, E207), although these residues located within the cleft are unlikely to be involved in protein/protein interactions. The contribution of the D loop to cleft closure is documented in the literature (Khaitlina and Strzelecka- Golaszewska, 2002). The cleavages in the D loop by ECP32 protease (G42-V43) or subtilisin (M47-G48) greatly stimulate the tryptic susceptibility of residues within SD2 (R62 and K68). The cleavage of the D loop apparently makes those accessibilities of tryptic cleavage sites in-crease. Thus, the D loop contributes to the stability of the closed structure although it is not clear how much contri-bution to the stability comes from the D loop. Our model predicts the interaction between M44 and the SD4 hydro-phobic pocket contributes to stabilization of the closed structure of the large cleft. The observations above sup-port our model, as the D-loop/SD4 interaction is a strong candidate for the change in dynamics of SD2. Catalytic Residue As a consequence of the large-cleft closure by large-domain movements relative to SD1, the geometry inside the nucleotide binding pocket is altered from those seen in typical crystal structures. The narrowed cleft brings D11, Q137, and D154 side chains closer to the metal ion and the b and g phosphates while maintaining the hydrogen-bond network already in place. D11 and Q137 are likely to be involved in the inner-sphere metal-ion coordination. On the other hand, D154 seems to be too far from the metal ion to coordinate but is located in a good position to form an inline attack of the g phosphate. The double-mutant D154A/D157A is lethal in yeast (Wertman et al., 1992); however, the S14C/D157A double mutant does not show significant change in ATPase activity (Schu¨ ler et al., 1999). Thus, it is very likely that the D154A mutation is directly involved in lethality. As suggested in the literature (Schu¨ ler, 2001), D154 appears to be a prime candidate for the catalytic residue from both previous reports and our model. Actin Model Consistency with Nucleotide and Metal-Ion Exchange Rates The stability constants of the Mg2+- and Ca2+-ATP com-plexes are similar (4.2 and 4.0, respectively) (Williams, 1970), and the exchange rate constants of the metal ions and ATP in water are quite rapid (3.3 3 105 and 2.0 3 105 s1) (De La Cruz and Pollard, 1995; Pecoraro et al., 1984). On the other hand, the exchange rates of Mg2+ and Ca2+ bound to ATP-G-actin are quite slow compared with the rates observed in water and are different from each other (association rates: 2.3 3 105 and 2 3 107 M1 s1; dissociation rates: 0.0015 and 0.014 s1 for Mg2+- and Ca2+-ATP-G-actin, respectively, as summa-rized in Table 3) (Gershman et al., 1991; Selden et al., 1989; Sheterline and Sparrow, 1994). These data indicate that the metal ions are significantly stabilized within the cleft by limited diffusion or coordination bond formation with protein side-chain oxygen atoms. The difference in association/dissociation with ATP between Mg2+ and Ca2+ is not significant in water but is an order of magnitude different in the case of actin-nucleotide-bound forms. Our model predicts that one factor explaining the difference in stability is coordination of Mg2+ with oxygen atoms of the protein. The other factor is the difference in accessibility of metal ions (steric effect). The exchange rates of ATP for Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 47 Structure Model of Mg2+-ATP-Actin Monomer
  • 10. Table 3. Exchange Rate Constants of Nucleotides/Metal Ions for Different Species of Actin Rate Constant Mg2+ Ca2+ Species Metal ion Exchange rate 3.3 3 105 s1 2.0 3 105 s1 ATP/metal ion Metal ion Association rate 2.3 3 105 M1 s1 2.0 3 107 M1 s1 Metal ion/ATP + actin Metal ion Dissociation rate 0.0015 s1 0.014 s1 Metal ion/ATP/actin ATP Exchange ratea 5.0 3 104 s1b 1.5 3 102 s1c Metal ion/ATP/actin Metal ion Affinity (Kd) 10 nM 2 nM ATP-G-actin a The exchange rate is highly dependent on metal-ion concentrations. bMg2+ concentration of 1 mM. c Ca2+ concentration of 0.1 mM. both Mg2+- and Ca2+-G-actin are even more different (5 3 104 and 1.5 3 102, respectively) (Sheterline and Spar-row, 1994). As ATP exchange rates are tightly connected to free metal concentrations, the metal-ion exchange rate would be a limiting factor for nucleotide exchange (Kinosian et al., 1993). Even if we attribute all of the metal exchange rate difference to coordination states, still there is a difference in ATP exchange rates. Thus, this ATP ex-change rate difference may be related to the steric effect of the cleft closure. These observed exchange rate differ-ences between Mg2+- and Ca2+-G-actin are quite consis-tent with our model (closed cleft and direct coordination of Mg2+ to the side-chain oxygen atoms). Implications for F-Actin Models The Holmes model of F-actin (Holmes et al., 1990; Lorenz et al., 1993; Tirion et al., 1995) has been widely accepted and is largely consistent with X-ray fiber diffraction (Bel-mont et al., 1999; Holmes et al., 1990; Lorenz et al., 1993; Oda et al., 2001; Tirion et al., 1995; Wu and Ma, 2004), cryoelectron microscopy (Schmid et al., 2004), and atomic force microscopy data (Shao et al., 2000; Shi et al., 2001). We have examined the F-actin structure by side-chain solvent accessibility utilizing hydroxyl-radical footprinting (Guan et al., 2005). Although our data were mainly consistent with the F-actin model, there are some disagreements with the Holmes model. The most appar-ent difference is the status of the large cleft. The protection sites in SD4 (residues 200–202 and 243) and the D loop (residues 40, 44, and 47) form contacts in the intermolec-ular interfaces; there is no disagreement with footprinting data for these residues. However, the hydroxyl-radical re-activity within the large cleft is significantly lower in F-actin compared with Ca2+-G-actin. A closed cleft is also ob-served in a reconstructed model from cryo-EM data (Schmid et al., 2004), although the degree of closure seems not to be enough to explain the almost 90% reduc-tion in hydroxyl-radical reactivity compared to the open form of Ca2+-G-actin. The ADP and ADP-BeF3 forms of fiber diffraction data show differences in the status of the large cleft (Belmont et al., 1999); the latter form indicates a more closed cleft. The replacement of ADP with ADP-BeF3 also diminishes susceptibility of F-actin to tryptic cleavage within SD2 (Muhlrad et al., 1994). As ADP-BeF3 mimics ATP or ADP-Pi (Muhlrad et al., 1994), this form represents ATP or ADP-Pi-F-actin before it releasesPi. These observations indicate that in the F-actin filament, hydrolysis of ATP to ADPand subsequent release of Pi correlate with large-cleft opening. On the other hand, when Mg-ATP-G-actin is in-corporated into the filament, we suggest it is in the highly closed form seen here; this Mg-ATP form of actin has a lower critical concentration than Ca2+-ATP-G-actin. Catal-ysis mediated by the Mg2+ ion may drive conformational changes among the coordination bonds between side chains and the metal ion, leading to rearrangements of subdomains 2 and 4 and the generation of a more open cleft within the ADP-F-actin structure. This would provide a more flexible and dynamic SD2, which is thought to be a key factor in attracting binding proteins and driving fila-ment dynamics (Schmid et al., 2004). Conclusion Understanding monomeric G-actin structure is important, as it is relevant to many protein/protein interactions in which actin participates. Hydroxyl-radical footprinting is a powerful technique to probe surface accessibility of macromolecules. The technique is a local measure as it reports side-chain accessibility. Such information pro-vides strict restraints for estimating changes in structure and in defining domain interactions. In conjunction with a computational approach, a structural model of the Mg- ATP-actin monomer was built that is consistent with ex-perimental data. This model provides novel insights into the conformational rearrangements driving actin filament dynamics and provides a unique structural insight into ac-tin monomer structure. This combined computational and experimental approach is particularly useful for modeling the structural changes of proteins when we do not have access to crystal or solution structures for conformations of functional interest, but where we may have suitable starting templates for modeling the structure. Further developments in computational modeling will ease the integration of the experimental into the computational approach as a filtering/validation tool. EXPERIMENTAL PROCEDURES Rigid-Body Movement Analyses Tcl scripts for molecular viewer VMD (Humphrey et al., 1996) were de-veloped for the analysis and visualization of rigid-body rotational 48 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 11. movements of domains/subdomains. The scripts accept the range of the target residues in two structures for rigid-body movement and then calculate vectors of atom movements and averaged normal for these vectors; they then search for the atoms within the reference structure that are nearest to the center of motion. The algorithm is based on the assumption that there are pivot points in bonds within the backbone allowing rigid-body rotational movements (the movements should be explained by rotations of some bond angles). The algorithm is based on the following step-by-step calculations. Define structures A and B, where A is the reference structure and B is the target structure for movement calculation. The total number of atoms(Ca orbackbone atoms) isn within the target range.The totalnum-ber of the atoms in reference structure A (Ca or backbone atoms) is N. v ! Atoms in structure A: ai (i = 1 to n), coordinates Pai. Atoms in structure B: bi (i = 1 to n), coordinates Pbi. Atoms for the center C: cj (j = 1 to N), coordinates Pcj. The midpoint of ai and bi: mi (i = 1 to n), coordinates Pmi. The vector between points ai and bi: i = Pbi Pai . The vector between points mi and cj: u ! ij =Pmi Pcj . The angle formed by the two vectors v ! i and u ! ij : qij = acos ð v ! i u ! jiÞ jv ! i j u ! ji : The averaged angle for the point Pcj: q j = 1 n Xn i =1 qij p 2 ð j =1toNÞ: The SD of the angles for atom Pcj: SDj = ffiPffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi n i =1ððqij p 2 q j 2 n 1 s ð j =1toNÞ The averaged angle between the real center of movement cr and the 2; thus, q two atoms in structures A and B should be p j is 0. We need to find the atom that gives the averaged angle closest to 0. In order to avoid the cases where q j happens to be close to 0 but angles vary as qij can be positive or negative, the score is calculated by the follow-ing formula: Sj =q j 3SDj : Then, the script calculates the average positions of the top scored five atoms. This point is defined as gc. Now, we need to calculate the normal vector for the rotational move-ments. The normal vector for the rotational movements should be ap-proximately perpendicular to all vectors from the center of movement to the midpoint of atoms in the two structures A and B. The normal vec-tor is also perpendicular to the vector vi between ai and bi. In order to minimize the contribution from small vectors that tend to have larger deviation from ideal ones (the vectors are extended by unitization along with ‘‘error’’), ‘‘normalized’’ normal vectors are weighed by their norms. After the average normal vector is calculated, it is renormalized to the unit vector. The vector between points mi and gc: w ! i =Pmi gc. The normal vector of rotation for points ai and bi: n ! i = v ! i 3 w ! i jv ! i jjw ! i j (crossproduct). Averaged (weighed) vector: n ! av = 1 n Xn i =1 i jjw ! i j,n ! ðj v ! iÞ: Normal vector (unit vector): n ! unit = n ! av jn ! : av j The resultant averaged normal vector and center of rotation is visualized along with vectors vi within VMD by Tcl commands in the script. Accessible Surface Area Calculations The accessible surface areas of side chains are calculated using the GETAREA 1.1 (Fraczkiewicz and Braun, 1998) web server with 1.4 A ° probe radius (http://www.scsb.utmb.edu/cgi-bin/get_a_form.tcl) with additional atom type library and residue type entries. In our previous report, we used rabbit skeletal muscle actin. As ASA greatly depends on the side chain, the sequence of rabbit skeletal muscle actin was mapped on 1YAG using MODELLER 7v7(Sali and Blundell, 1993). This structure is designated as 1YAG (rabbit sequence) in the following sections. Computational Modeling Model construction started from a ‘‘template’’ structure of 1YAG (rab-bit sequence). The rotational movement toward the small domain was applied to the large domain (residues 138–332) at the pivot residue 138 with the normal vector calculated in 1HLU (Chik et al., 1996). The rota-tional movement toward SD2 was applied as a combination of several movements calculated based on the analyses of several crystal struc-tures. The rotations were applied over four residues from specifically selected pivot points (residues 34–37 and 66–69) in order to prevent too much rotation to single bonds that can lead to improper bond an-gles. All rotational transformations of coordinates were performed us-ing Tcl scripts within VMD. Modeled structures were generated with combinations of different angles and normal vectors. ASAs of side chains were calculated for these structures in order to assess the pre-liminary models. The best model in the series of models by rigid-body movements was chosen according to the consistency with expected changes in ASAs of hydroxyl-radical probes. The model was then subjected to manual inspections/manipulations of side chains in the Swiss-PdbViewer (Guex and Peitsch, 1997) in or-der to better satisfy the hydroxyl-radical data. The rotamer of side chains was selected to maximize hydrogen-bond formations within the large cleft. As the large cleft between SD2 and SD4 was closed by moving the large domain relative to SD1, the nucleotide binding pocket was also affected and became significantly narrower than in the template structure 1YAG. The coordinates of the phosphates of ATP were manually adjusted within VMD using Tcl scripts by rotating the phosphate-oxygen bond and fit into the narrower cleft interactively in order to avoid steric clashes. The D loop was constructed manually in the Swiss-PdbViewer by rotating the phi and psi angles of the tem-plate structure. This process was guided mainly to allow M44 to reach the hydrophobic pocket in SD4 while at the same time avoiding im-proper angles. At this stage, the model structure was examined by PROCHECK (Laskowski et al., 1993) and Verify3D (Bowie et al., 1991) for model validity (Figure S1). The model coordinates were then passed to MODELLER 7v7 (Sali and Blundell, 1993) as a template for adjusting and correcting the stereochemical parameters. MODELLER was also used to adjust the geometry of the residues and metal ion in the nucle-otide binding pocket with manually assigned restraints according to the 1RFQ-B (chain B) structure (Reutzel et al., 2004). After further manual inspections/adjustments in VMD and energy minimization within the Swiss-PdbViewer for certain side chains, the final structural model was subjected to PROCHECK and Verify3D for the assessment of structure validity. Visualization of Structures The structures were visualized using POV-Ray ray-tracer (Version 3.6) with output from VMD (Version 1.8.3) and manual editing of scene files. Some presentations were directly defined by POV-Ray scene. All op-erations were done on a Power Macintosh G5 with Mac OS X except for energy minimization within the Swiss-PdbViewer that was per-formed on a Windows 2000 workstation (OS X version is an a version and energy minimization is only partially implemented). Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 49 Structure Model of Mg2+-ATP-Actin Monomer
  • 12. Supplemental Data Supplemental Data include two figures and Supplemental Results and are available at http://www.structure.org/cgi/content/full/15/1/ 39/DC1/. ACKNOWLEDGMENTS This work is supported in part by the Biomedical Technology Centers Program of the National Institutes for Biomedical Imaging and Bioen-gineering (P41-EB-01979). Received: July 26, 2006 Revised: November 6, 2006 Accepted: November 18, 2006 Published: January 16, 2007 REFERENCES Allingham, J.S., Tanaka, J., Marriott, G., and Rayment, I. (2004). Abso-lute stereochemistry of ulapualide A. Org. Lett. 6, 597–599. Belmont, L.D., Orlova, A., Drubin, D.G., and Egelman, E.H. (1999). A change in actin conformation associated with filament instability after Pi release. Proc. Natl. Acad. Sci. USA 96, 29–34. Bowie, J.U., Luthy, R., and Eisenberg, D. (1991). A method to identify protein sequences that fold into a known three-dimensional structure. Science 253, 164–170. Bubb, M.R., Govindasamy, L., Yarmola, E.G., Vorobiev, S.M., Almo, S.C., Somasundaram, T., Chapman, M.S., Agbandje-McKenna, M., and McKenna, R. (2002). Polylysine induces an antiparallel actin dimer that nucleates filament assembly: crystal structure at 3.5-A° resolution. J. Biol. Chem. 277, 20999–21006. Chen, X., Peng, J., Pedram, M., Swenson, C.A., and Rubenstein, P.A. (1995). The effect of the S14A mutation on the conformation and ther-mostability of Saccharomyces cerevisiae G-actin and its interaction ° with adenine A nucleotides. J. Biol. Chem. 270, 11415–11423. Chik, J.K., Lindberg, U., and Schutt, C.E. (1996). The structure of an open state of b-actin at 2.65 resolution. J. Mol. Biol. 263, 607–623. De La Cruz, E.M., and Pollard, T.D. (1995). Nucleotide-free actin: stabilization by sucrose and nucleotide binding kinetics. Biochemistry 34, 5452–5461. Dudev, T., Cowan, J.A., and Lim, C. (1999). Competitive binding in magnesium coordination chemistry: water versus ligands of biological interest. J. Am. Chem. Soc. 121, 7665–7673. Fraczkiewicz, R., and Braun, W. (1998). Exact and efficient analytical calculation of the accessible surface areas and their gradients for mac-romolecules. J. Comput. Chem. 19, 319–333. Frieden, C., and Patane, K. (1985). Differences in G-actin containing bound ATP or ADP: the Mg2+-induced conformational change requires ATP. Biochemistry 24, 4192–4196. Gershman, L.C., Selden, L.A., and Estes, J.E. (1991). High affinity diva-lent cation exchange on actin. Association rate measurements support the simple competitive model. J. Biol. Chem. 266, 76–82. Guan, J.Q., and Chance, M.R. (2005). Structural proteomics of macro-molecular assemblies using oxidative footprinting and mass spec-trometry. Trends Biochem. Sci. 30, 583–592. Guan, J.Q., Vorobiev, S., Almo, S.C., and Chance, M.R. (2002). Map-ping the G-actin binding surface of cofilin using synchrotron protein footprinting. Biochemistry 41, 5765–5775. Guan, J.Q., Almo, S.C., Reisler, E., and Chance, M.R. (2003). Structural reorganization of proteins revealed by radiolysis and mass spectrom-etry: G-actin solution structure is divalent cation dependent. Biochem-istry 42, 11992–12000. Guan, J.Q., Almo, S.C., and Chance, M.R. (2004). Synchrotron radiol-ysis and mass spectrometry: a new approach to research on the actin cytoskeleton. Acc. Chem. Res. 37, 221–229. Guan, J.Q., Takamoto, K., Almo, S.C., Reisler, E., and Chance, M.R. (2005). Structure and dynamics of the actin filament. Biochemistry 44, 3166–3175. Guex, N., and Peitsch, M.C. (1997). SWISS-MODEL and the Swiss- PdbViewer: an environment for comparative protein modeling. Electrophoresis 18, 2714–2723. Gupta, S., Mangel, W.F., McGrath, W.J., Perek, J.L., Lee, D.W., Taka-moto, K., and Chance, M.R. (2004). DNA binding provides a molecular strap activating the adenovirus proteinase. Mol. Cell. Proteomics 3, 950–959. Holmes, K.C., Popp, D., Gebhard, W., and Kabsch, W. (1990). Atomic model of the actin filament. Nature 347, 44–49. Humphrey, W., Dalke, A., and Schulten, K. (1996). VMD: visual molec-ular dynamics. J. Mol. Graph. 14, 33–38. Joel, P.B., Fagnant, P.M., and Trybus, K.M. (2004). Expression of a nonpolymerizable actin mutant in Sf9 cells. Biochemistry 43, 11554– 11559. Kabsch, W., Mannherz, H.G., and Suck, D. (1985). Three-dimensional structure of the complex of actin and DNase I at 4.5 A° resolution. EMBO J. 4, 2113–2118. Kabsch, W., Mannherz, H.G., Suck, D., Pai, E.F., and Holmes, K.C. (1990). Atomic structure of the actin:DNase I complex. Nature 347, 34–44. Khaitlina, S.Y., and Strzelecka-Golaszewska, H. (2002). Role of the DNase-I-binding loop in dynamic properties of actin filament. Biophys. J. 82, 321–334. Kinosian, H.J., Selden, L.A., Estes, J.E., and Gershman, L.C. (1993). Nucleotide binding to actin. Cation dependence of nucleotide dissoci-ation and exchange rates. J. Biol. Chem. 268, 8683–8691. Kiselar, J.G., Janmey, P.A., Almo, S.C., and Chance, M.R. (2003a). Structural analysis of gelsolin using synchrotron protein footprinting. Mol. Cell. Proteomics 2, 1120–1132. Kiselar, J.G., Janmey, P.A., Almo, S.C., and Chance, M.R. (2003b). Visualizing the Ca2+-dependent activation of gelsolin by using syn-chrotron footprinting. Proc. Natl. Acad. Sci. USA 100, 3942–3947. Klenchin, V.A., Allingham, J.S., King, R., Tanaka, J., Marriott, G., and Rayment, I. (2003). Trisoxazole macrolide toxins mimic the binding of actin-capping proteins to actin. Nat. Struct. Biol. 10, 1058–1063. Laskowski, R.A., MacArthur, M.W., Moss, D.S., and Thornton, J.M. (1993). PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Crystallogr. 26, 283–291. Liu, R., Guan, J.Q., Zak, O., Aisen, P., and Chance, M.R. (2003). Structural reorganization of the transferrin C-lobe and transferrin receptor upon complex formation: the C-lobe binds to the receptor helical domain. Biochemistry 42, 12447–12454. Lorenz, M., Popp, D., and Holmes, K.C. (1993). Refinement of the F-actin model against X-ray fiber diffraction data by the use of a di-rected mutation algorithm. J. Mol. Biol. 234, 826–836. Mannherz, H.G., Gooch, J., Way, M., Weeds, A.G., and McLaughlin, P.J. (1992). Crystallization of the complex of actin with gelsolin seg-ment 1. J. Mol. Biol. 226, 899–901. McLaughlin, P.J., Gooch, J.T., Mannherz, H.G., and Weeds, A.G. (1993). Structure of gelsolin segment 1-actin complex and the mecha-nism of filament severing. Nature 364, 685–692. Moraczewska, J., Wawro, B., Seguro, K., and Strzelecka-Golaszew-ska, H. (1999). Divalent cation-, nucleotide-, and polymerization-dependent changes in the conformation of subdomain 2 of actin. Biophys. J. 77, 373–385. Morton, W.M., Ayscough, K.R., and McLaughlin, P.J. (2000). Latruncu-lin alters the actin-monomer subunit interface to prevent polymeriza-tion. Nat. Cell Biol. 2, 376–378. Muhlrad, A., Cheung, P., Phan, B.C., Miller, C., and Reisler, E. (1994). Dynamic properties of actin. Structural changes induced by beryllium fluoride. J. Biol. Chem. 269, 11852–11858. 50 Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved Structure Model of Mg2+-ATP-Actin Monomer
  • 13. Oda, T., Makino, K., Yamashita, I., Namba, K., and Maeda, Y. (2001). Distinct structural changes detected by X-ray fiber diffraction in stabi-lization of F-actin by lowering pH and increasing ionic strength. Biophys. J. 80, 841–851. Otterbein, L.R., Graceffa, P., and Dominguez, R. (2001). The crystal structure of uncomplexed actin in the ADP state. Science 293, 708– 711. Otterbein, L.R., Cosio, C., Graceffa, P., and Dominguez, R. (2002). Crystal structures of the vitamin D-binding protein and its complex with actin: structural basis of the actin-scavenger system. Proc. Natl. Acad. Sci. USA 99, 8003–8008. Paavilainen, V.O., Bertling, E., Falck, S., and Lappalainen, P. (2004). Regulation of cytoskeletal dynamics by actin-monomer-binding pro-teins. Trends Cell Biol. 14, 386–394. Page, R., Lindberg, U., and Schutt, C.E. (1998). Domain motions in actin. J. Mol. Biol. 280, 463–474. Pecoraro, V.L., Hermes, J.D., and Cleland, W.W. (1984). Stability constants of Mg2+ and Cd2+ complexes of adenine nucleotides and thionucleotides and rate constants for formation and dissociation of MgATP and MgADP. Biochemistry 23, 5262–5271. Rashidzadeh, H., Khrapunov, S., Chance, M.R., and Brenowitz, M. (2003). Solution structure and interdomain interactions of the Saccha-romyces cerevisiae ‘‘TATA binding protein’’ (TBP) probed by radiolytic protein footprinting. Biochemistry 42, 3655–3665. Reutzel, R., Yoshioka, C., Govindasamy, L., Yarmola, E.G., Agbandje- McKenna, M., Bubb, M.R., and McKenna, R. (2004). Actin crystal dy-namics: structural implications for F-actin nucleation, polymerization, and branching mediated by the anti-parallel dimer. J. Struct. Biol. 146, 291–301. Sali, A., and Blundell, T.L. (1993). Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 234, 779–815. Schmid, M.F., Sherman, M.B., Matsudaira, P., and Chiu, W. (2004). Structure of the acrosomal bundle. Nature 431, 104–107. Schoenenberger, C.A., Bischler, N., Fahrenkrog, B., and Aebi, U. (2002). Actin’s propensity for dynamic filament patterning. FEBS Lett. 529, 27–33. Schu¨ ler, H. (2001). ATPase activity and conformational changes in the regulation of actin. Biochim. Biophys. Acta 1549, 137–147. Schu¨ ler, H., Korenbaum, E., Schutt, C.E., Lindberg, U., and Karlsson, R. (1999). Mutational analysis of Ser14 and Asp157 in the nucleotide-binding site of b-actin. Eur. J. Biochem. 265, 210–220. Schutt, C.E., Myslik, J.C., Rozycki, M.D., Goonesekere, N.C., and Lindberg, U. (1993). The structure of crystalline profilin-b-actin. Nature 365, 810–816. Selden, L.A., Estes, J.E., and Gershman, L.C. (1989). High affinity diva-lent cation binding to actin. Effect of low affinity salt binding. J. Biol. Chem. 264, 9271–9277. Shao, Z., Shi, D., and Somlyo, A.V. (2000). Cryoatomic force micros-copy of filamentous actin. Biophys. J. 78, 950–958. Sharp, J.S., Guo, J.T., Uchiki, T., Xu, Y., Dealwis, C., and Hettich, R.L. (2005). Photochemical surface mapping of C14S-Sml1p for con-strained computational modeling of protein structure. Anal. Biochem. 340, 201–212. Sheterline, P., and Sparrow, J.C. (1994). Actin. Protein Profile 1, 1–121. Shi, D., Somlyo, A.V., Somlyo, A.P., and Shao, Z. (2001). Visualizing fil-amentous actin on lipid bilayers by atomic force microscopy in solu-tion. J. Microsc. 201, 377–382. Strzelecka-Golaszewska, H., Moraczewska, J., Khaitlina, S.Y., and Mossakowska, M. (1993). Localization of the tightly bound divalent-cation- dependent and nucleotide-dependent conformation changes in G-actin using limited proteolytic digestion. Eur. J. Biochem. 211, 731–742. Suck, D., Kabsch, W., and Mannherz, H.G. (1981). Three-dimensional structure of the complex of skeletal muscle actin and bovine pancre-atic DNAse I at 6-A° resolution. Proc. Natl. Acad. Sci. USA 78, 4319– 4323. Swamy, N., Head, J.F., Weitz, D., and Ray, R. (2002). Biochemical and preliminary crystallographic characterization of the vitamin D sterol-and actin-binding by human vitamin D-binding protein. Arch. Biochem. Biophys. 402, 14–23. Takamoto, K., and Chance, M.R. (2006). Radiolytic protein footprinting with mass spectrometry to probe the structure of macromolecular complexes. Annu. Rev. Biophys. Biomol. Struct. 35, 251–276. Tirion, M.M., ben-Avraham, D., Lorenz, M., and Holmes, K.C. (1995). Normal modes as refinement parameters for the F-actin model. Biophys. J. 68, 5–12. Valentin-Ranc, C., and Carlier, M.F. (1991). Role of ATP-bound diva-lent metal ion in the conformation and function of actin. Comparison of Mg-ATP, Ca-ATP, and metal ion-free ATP-actin. J. Biol. Chem. 266, 7668–7675. Verboven, C., Bogaerts, I., Waelkens, E., Rabijns, A., Van Baelen, H., Bouillon, R., and De Ranter, C. (2003). Actin-DBP: the perfect struc-tural fit? Acta Crystallogr. D Biol. Crystallogr. 59, 263–273. Vorobiev, S., Strokopytov, B., Drubin, D.G., Frieden, C., Ono, S., Condeelis, J., Rubenstein, P.A., and Almo, S.C. (2003). The structure of nonvertebrate actin: implications for the ATP hydrolytic mechanism. Proc. Natl. Acad. Sci. USA 100, 5760–5765. Wear, M.A., Schafer, D.A., and Cooper, J.A. (2000). Actin dynamics: assembly and disassembly of actin networks. Curr. Biol. 10, R891– R895. Wertman, K.F., Drubin, D.G., and Botstein, D. (1992). Systematic mutational analysis of the yeast ACT1 gene. Genetics 132, 337–350. Williams, R.P. (1970). Biochemistry of sodium, potassium, magnesium and calcium. Q. Rev. Chem. Soc. 24, 331–365. Winder, S.J. (2003). Structural insights into actin-binding, branching and bundling proteins. Curr. Opin. Cell Biol. 15, 14–22. Wriggers, W., and Schulten, K. (1997). Stability and dynamics of G-actin: back-door water diffusion and behavior of a subdomain 3/4 loop. Biophys. J. 73, 624–639. Wu, Y., and Ma, J. (2004). Refinement of F-actin model against fiber diffraction data by long-range normal modes. Biophys. J. 86, 116–124. Xu, G., and Chance, M.R. (2005). Radiolytic modification and reactivity of amino acid residues serving as structural probes for protein foot-printing. Anal. Chem. 77, 4549–4555. Yarmola, E.G., Somasundaram, T., Boring, T.A., Spector, I., and Bubb, M.R. (2000). Actin-latrunculin A structure and function. Differential modulation of actin-binding protein function by latrunculin A. J. Biol. Chem. 275, 28120–28127. Zimmerle, C.T., Patane, K., and Frieden, C. (1987). Divalent cation binding to the high- and low-affinity sites on G-actin. Biochemistry 26, 6545–6552. Structure 15, 39–51, January 2007 ª2007 Elsevier Ltd All rights reserved 51 Structure Model of Mg2+-ATP-Actin Monomer