ArticlePDF Available

Large Amplitude Elastic Motions in Proteins from a Single-Parameter, Atomic AnalysisTirionPhys Rev Lett19967791905190810.1103/PhysRevLett.77.190510063201

Authors:

Abstract and Figures

Normal mode analysis (NMA) is a leading method for studying long-time dynamics and elasticity of biomolecules. The method proceeds from complex semiempirical potentials characterizing the covalent and noncovalent interactions between atoms. It is widely accepted that such detailed potentials are essential to the success of NMA's. We show that a single-parameter potential is sufficient to reproduce the slow dynamics in good detail. Costly and inaccurate energy minimizations are eliminated, permitting direct analysis of crystal coordinates. The technique can be used for new applications, such as mapping of one crystal form to another by means of slow modes, and studying anomalous dynamics of large proteins and complexes.
Content may be subject to copyright.
VOLUME 77, NUMBER 9 PHYSICAL REVIEW LETTERS 26AUGUST 1996
Large Amplitude Elastic Motions in Proteins from a Single-Parameter, Atomic Analysis
Monique M. Tirion*
Department of Membrane Research and Biophysics, Weizmann Institute of Science, Rehovot 76100, Israel
(Received 22 April 1996)
Normal mode analysis (NMA) is a leading method for studying long-time dynamics and elasticity
of biomolecules. The method proceeds from complex semiempirical potentials characterizing the
covalent and noncovalent interactions between atoms. It is widely accepted that such detailed potentials
are essential to the success of NMA’s. We show that a single-parameter potential is sufficient to
reproduce the slow dynamics in good detail. Costly and inaccurate energy minimizations are eliminated,
permitting direct analysis of crystal coordinates. The technique can be used for new applications, such
as mapping of one crystal form to another by means of slow modes, and studying anomalous dynamics
of large proteins and complexes. [S0031-9007(96)01063-0]
PACS numbers: 87.15.By, 87.15.He
Thermal equilibrium fluctuations of the x-ray crystal
coordinates of proteins provide a basis for understanding
the complex dynamics and elasticity of biological macro-
molecules [1]. Analysis of the normal modes of globu-
lar proteins shows an interesting anomaly. The density
of the slow vibrational modes is proportional to their fre-
quency, gsvd,v, rather than gsvd,v
2
as predicted
by Debye’s theory [2]. Yet, the atoms in globular pro-
teins are packed as tightly as in solids. We show that
a single-parameter potential reproduces the slow elastic
modes of proteins obtained with vastly more complex
empirical potentials. The simplicity of the potential per-
mits greater insight and understanding of the mechanisms
that underlie the slow, anomalous motions in biological
macromolecules such as proteins.
To date, normal modes of globular proteins have been
used to reproduce crystallographic temperature factors [3]
and diffuse scatter [4]. Normal mode analyses (NMA’s)
shed light on shear and hinge motions necessary for cata-
lytic reactions, and have been used with some success to
map one crystal form of a protein into another [5]. Finally,
NMA’s yield macroscopic elastic moduli of large protein
assemblies, based on their microscopic structure [6].
NMA studies of macromolecules are handicapped, how-
ever, by the complex phenomenological potentials used to
model the covalent and nonbonded interactions between
atom pairs. The necessary initial energy minimization re-
quires a great deal of computer time and memory, and
is virtually impossible for even moderately large proteins
(with typically thousands of degrees of freedom) with a
reasonable degree of accuracy. This inevitably leads to
unstable modes which must be eliminated through elabo-
rate methods, and which cast doubts on the validity of the
analysis. Moreover, partly because the minimization is
carried out in vacuo, the final configuration disagrees with
the known crystallographic structure, complicating the in-
terpretation of the results of NMA.
A typical example of a semiempirical potential used in
molecular dynamics studies and NMA’s has the form [7]
E
p
1
2
X
bonds
K
b
sb 2 b
0
d
2
1
1
2
X
angles
K
u
su2u
0
d
2
1
1
2
X
dihedrals
K
f
f1 1 cossnf2ddg
1
X
nonbonded pairs
A
r
12
2
B
r
6
1
q
1
q
2
Dr
. (1)
The first three terms describe the energy cost in the
distortion of bond lengths, bond angles, and dihedral
angles, and the last term represents steric repulsions,
van der Waals attractions, and electrostatic interactions
between nonbonded atoms. The various bonded con-
stants, K
b
, b
0
, K
u
, etc., are specific for each type of co-
valent interaction, and the nonbonded constants, A and
B, are specific for every type of interacting atom pairs.
These constants are carefully determined from extensive
theoretical and experimental studies (see Ref. [7]).
In this work, I show that in some cases the usual sophis-
ticated potentials may be replaced by a far simpler pair-
wise Hookean potential, controlled by a single parameter.
Such a formulation is sufficient to fully describe the
anomalous low-frequency motion of large globular pro-
teins, including time scales and eigenfrequencies, as well
as displacements of atoms as predicted by eigenvectors.
The simplified potential provides a very attractive alter-
native for the NMA of macromolecular assemblies. The
derivation of the eigenvalue equation is simple, rapid, and
accurate. Time-consuming and structure-distorting en-
ergy minimization are circumvented, preventing unphysi-
cal instabilities (negative eigenvalues), and the reduction
in the number of fitting parameters yields to the new for-
mulation a stronger predictive value.
The size of the computation depends on the number
of internal coordinates considered. A molecule consist-
ing of N atoms possesses 3N degrees of freedom, defined
by bond lengths, bond angles, and rotations about bonds
(or, alternatively, by the Cartesian coordinates of each of
the constituent atoms). For long-chain molecules such as
0031-9007y96y77(9)y1905(4)$10.00 © 1996 The American Physical Society 1905
VOLUME 77, NUMBER 9 PHYSICAL REVIEW LETTERS 26AUGUST 1996
proteins, bond lengths and angles are constrained within
very narrow limits by the chemical bonding, while ro-
tations about (single) bonds are much less restricted.
Hence, for the analysis of slow modes one typically con-
siders the bond lengths and angles as fixed, permitting
only rotations about bonds. The latter are known as dihe-
dral angles and serve as a convenient set of generalized
degrees of freedom. The number of the dihedral angle co-
ordinates is about Ny2 and is determined by the specific
makeup of the protein in question.
Given a potential energy function E
p
and a set of
n generalized coordinates, q
0
, one first must find a
stable local minimum E
p,min
sq
0
q
0
0
d. The potential
energy is then approximated by the quadratic form: E
p
s1y2d
P
q
i
F
ij
q
j
s1y2dq
y
F
q
sq ; q
0
2 q
0
0
d, where F is
the generalized force matrix
F
i,j
2
E
p
q
i
q
j
Ç
q0
.
Similarly, the kinetic energy is expressed as E
k
s1y2d 3
Ù
q
y
H
Ù
q, where the elements of the “mass” matrix H are
H
i,j
N
X
l1
m
l
r
l
q
i
?
r
l
q
j
.
r
l
are the Cartesian coordinates of atom l, and the sum-
mation runs over all the atoms of the molecule. The
hryqj are moving derivatives which eliminate transla-
tional and rotational motion of the molecule as a whole.
Equations of motion are derived from Lagrange’s
equation, with the Lagrangian L E
k
2 E
p
. Writing
q
j
P
N
k
A
jk
a
k
cossv
k
t 1d
k
d, one obtains the eigen-
value problem
FA LHA , (2)
subject to the normalization condition A
y
HA I. (This
ensures that the eigenmodes diagonalize the system’s
Hamiltonian, H E
k
1 E
p
.) The eigenfrequencies v
i
are given by the elements of the diagonal matrix L, v
2
i
L
ii
, the eigenvectors are the columns of the matrix A, and
the amplitudes and phases, a
k
and d
k
, are determined by
initial conditions.
I replace the habitual detailed potentials, such as
the one in Eq. (1), by the Hookean pairwise potential
(between atoms a and b):
Esr
a
, r
b
d
C
2
sjr
a,b
j 2 jr
0
a,b
jd
2
. (3)
Here r
a,b
; r
a
2 r
b
denotes the vector connecting
atoms a and b, and the zero superscript indicates the
given initial configuration. Thus, the usual minimization
of the potential energy is eliminated.
Expanding to second order about r
0
a,b
yields
Esr
a
, r
b
d
C
2
µ
r
0
a,b
?Dr
a,b
jr
0
a,b
j
2
, (4)
where Dr ; r 2 r
0
. The strength of the potential C is
a phenomenological constant, assumed to be the same for
all interacting pairs.
The potential energy within a molecule is then given by
E
p
X
sa,bd
Esr
a
, r
b
d . (5)
The sum is restricted to atom pairs separated by less
than R
vdW
sad 1 R
vdW
sbd 1 R
c
, where R
vdW
refers to
the van der Waals radii, and R
c
is an arbitrary cutoff
parameter which models the decay of interactions with
distance. R
c
determines the total number of interacting
atom pairs contributing to the potential energy of the
system, and is inversely related to the “bond strength” C.
We shall argue below that best results are obtained with
small cutoff distances. Fortunately, this tends to reduce
the size of the computation.
The proposed potential seems too simplistic. For ex-
ample, it is unclear how interactions between nonbonded
atoms could model the 3-state “knob” potential for ro-
tations about the backbone. The answer lies in the fact
that slow vibrational modes involve coherent motion of
large groups of atoms. The effective force opposing large
scale oscillations stems from the combined effect of nu-
merous interacting atom pairs. The sum of these interac-
tions approaches a universal form, governed by the central
limit theorem, regardless of the details of individual pair-
potentials. Hence, for slow vibrations these details could
be neglected.
To test this hypothesis, I compare the eigenfrequency
and eigenvector data obtained using the potential of
Eqs. (4) and (5) and the detailed potential of Levitt, L79
[7]. In all cases, I refer only to the slowest frequency
modes. The higher frequency modes, pertaining to rapid
oscillations of sidechains and small groups of atoms,
require an accurate analysis at the microscopic level and
could not be modeled by simplified potentials.
I performed extensive tests on the muscle protein,
G-actin. This system has a molecular weight of 44 kD
and contains 375 residues (3500 atoms) in a single
polypeptide chain [8]. The polypeptide chain is folded
so as to form two large domains joined by a narrow neck
region. These two domains are partly held together by
salt bridges and hydrogen bonds provided by a nucleotide
(ADP) and a cation sCa
11
d bound in the cleft between the
two domains. Complete eigenfrequency and eigenvector
data exist for this system using the L79 potential [9].
Focusing first on the density of modes, let us examine
the cumulative density of modes up to frequency v,
Gsvd n
21
R
v
0
dv
0
gsv
0
d. In globular proteins Gsvd
falls under a universal curve [2]. For small frequencies,
Gsvd,v
2
. The difference from regular crystals, where
Gsvd,v
3
, reflects the anomalous dynamics of slow
vibrations in proteins.
Figure 1 shows Gsvd against v for G-actin:ADP:Ca
11
for the slowest 10% of the modes (138 modes). The
dashed curve refers to data obtained using the standard
L79 potential. Superposed are curves obtained with
R
c
1.1, 1.5, 2.0, and 2.5 Å, resulting in 19248, 27310,
1906
VOLUME 77, NUMBER 9 PHYSICAL REVIEW LETTERS 26AUGUST 1996
FIG. 1. The fraction of the total number of modes up
to frequency v scm
21
d for the slowest 150 modes of the
G-actin:ADP:Ca
11
system. The dashed line pertains to data
obtained using the L79 potential, while the four solid curves
are obtained using R
C
values of 1.1, 1.5, 2.0, and 2.5 Å.
The R
c
1.1 Å curve is nearest the dashed line at higher
frequencies, with the fit progressively worsening for the higher
cutoff values.
38654, and 51020 nonbonded interactions, respectively.
To obtain optimal fits to the standard (dashed) curve,
the values of C need to be adjusted to 2.49, 1.29,
0.73, and 0.47 kJyÅ
2
mol, respectively. Curiously, C
seems to scale as 1yR
2
c
rather than 1yR
3
c
. This may
reflect the unusual spectral dimension, i.e., the effective
dimensionality of interatomic interactions, of globular
proteins, d
s
2 [2]. The product CR
2
c
results in a
universal “bond-strength” constant of about 3.0 kJymol.
Figure 1 also shows that larger cutoff values increase
the curvature of the curve, and hence smaller R
c
val-
ues better model the previous eigenvalue data. This em-
phasizes that nonbonded interactions are dominated by
nearest-neighbor interactions, presumably due to screen-
ing effects.
In addition to the frequency of the modes, one can
compute the root mean square (rms) displacements of
atoms from equilibrium at room temperature from the
eigenvector data. Figure 2 shows the rms fluctuations of
all the C
a
atoms as a function of the slowest 30 modes.
The dashed curve shows the data obtained using the
standard, L79 potential. The rms deviations decrease
rapidly with mode number, indicating that the correlation
length of the motion decreases as well. Superposed on
the dashed curve are the four curves obtained with the
same R
c
and C values used in Fig. 1. Values of C
predicted from the eigenfrequency data, Fig. 1, fit the
eigenvector data equally well. These data also indicate
no clear advantage to using larger values of the cutoff
parameter, R
c
. For the slowest modes, therefore, there is
overall consistency and a very good match of the current
FIG. 2. The rms deviation of all mainchain C
a
atoms per
mode, for the slowest 30 modes. The dashed line refers to data
obtained using the L79 potential, and the four solid curves are
obtained with the same cutoff values, R
c
, as in Fig. 1. The rms
fluctuations due to the first four modes contribute over 50% to
the total rms deviations of all C
a
due to all modes [8]. There
is no obvious improvement in choosing one particular cutoff
parameter according to these data.
results with those obtained previously, in terms of both
dispersion spectra as well as rms deviations.
Theoretical temperature factors, B, used to model
x-ray crystallographic temperature factors, are obtained
by computing the rms fluctuations at room temperature
of each C
a
due to a superposition of modes. Figure 3
shows the theoretical temperature factors for each C
a
of
FIG. 3. Comparison of theoretical temperature factors, B,
obtained with the L79 potential (dashed curve) and the
potential of Eq. (1), for the G-actin:ADP:Ca
11
system. The
contributions of the 30 slowest modes are included. The inset
shows the scatter plot of the two data sets: the standard potential
along the ordinate, and the current simplified potential along the
abscissa.
1907
VOLUME 77, NUMBER 9 PHYSICAL REVIEW LETTERS 26AUGUST 1996
TABLE I. CPU time requirements to compute generalized eigenvalue equations.
pdb # of # of # of nonbonded CPU time
Protein identifier residues coordinates interactions (min)
Crambin 1crn 46 139 4817 0.12
Trypsin inhibitor 5pti 58 208 6529 0.45
Ribonuclease A 5rsa 124 455 14946 2.50
Lysozyme 6lyz 129 471 15834 3.10
G-actin 1atn 372 1382 37951 66.0
Myosin (HC) 1mys 780 3010 88653 390
G-actin:ADP:Ca
11
, including the contribution of the 30
slowest modes. The data obtained with the L79 potential
are shown by the dashed curve. The solid line is
obtained using the current potential, with R
c
2.2 Å.
The very good fit argues against the need for additional
parametrization in Eqs. (3)(5).
Finally, I have tested the efficiency of NMA with the
simple potential. Table I shows the central processing
unit (CPU) times required for reading in coordinates and
chemical formulas, indexing degrees of freedom and non-
bonded interactions, and setting up the generalized eigen-
value Eq. (2). The entries show the CPU requirements
on a Convex 220 for the following four proteins (the
Brookhaven protein database label is given in parenthe-
ses): bovine pancreatic trypsin inhibitor (5pti), ribonucle-
ase A (5rsa), G-actin bound with ADP and Ca
11
(1atn),
and myosin subfragment 1 bound with ADP (lmys, with
the sidechains kindly modeled by Michael Lorenz). These
proteins range in size from 58 residues (trypsin inhibitor)
to 780 residues (myosin S1). For this table, I used a cut-
off distance R
c
2.0 Å. These CPU times represent im-
provements of 2 to 3 orders of magnitudes over earlier
NMA. The main effect is due to the absence of mini-
mizations and the faster computation of the force matrix F
with the simple potential. It should be stressed that NMA
of systems as large and poorly resolved as myosin S1
could not be undertaken with “standard” potentials, due to
the accumulation of roundoff errors and distortions during
minimizations.
This work demonstrates the surprising result that a
single-parameter model can reproduce complex vibra-
tional properties of macromolecular systems. The simple
form of the potential dispenses with the need to perform
initial energy minimizations, which are especially detri-
mental for NMA’s due to the absence of solvent. Since
the analysis proceeds directly from the crystal coordinates,
it is now possible to quantitatively test whether two crystal
forms of a protein, as in an “open” and “closed” configu-
ration, are interconvertible using the slow modes as coor-
dinates. Tests performed on a periplasmic maltodextrin
binding protein indicate that the slowest modes do indeed
closely map the open form into the closed form [Tirion, in
preparation].
I thank Daniel ben-Avraham for useful discussions and
ideas, Ken Holmes for use of computer facilities at the
Max-Planck Institute for Medical Research in Heidelberg,
and Michael Lorenz for the full coordinates of myosin S1.
This material is based upon work supported by the
National Science Foundation under Grant No. MCB-
9316109.
*Electronic address: mmt@craft.camp.clarkson.edu
[1] M. Levitt, P. Stern, and C. Sander, J. Mol. Biol. 181,
423 (1985); A. Kidera and N. G
¯
o, J. Mol. Biol. 225, 457
(1992); B. Brooks and M. Karplus, Proc. Natl. Acad. Sci.
U.S.A. 80, 6571 (1983).
[2] D. ben-Avraham, Phys. Rev. B 47, 14559 (1993).
[3] R. Diamond, Acta Crystallogr. A46, 425 (1990);
A. Kidera and N. G¯o, J. Mol. Biol. 225, 457 (1992).
[4] P. Faure et al., Nature Struct. Biol. 1, 124 (1994).
[5] O. Marques and Y.-H. Sanejouand, Proteins 33, 557
(1995); M.M. Tirion, D. ben-Avraham, M. Lorenz, and
K.C. Holmes, Biophys. J. 68, 5 (1995).
[6] D. ben-Avraham and M. M. Tirion, Biophys. J. 68, 1231
(1995).
[7] S. J. Weiner et al., J. Am. Chem. Soc. 106, 765 (1984);
B.R. Brooks et al., J. Comp. Chem. 4, 187 (1983);
M. Levitt, J. Mol. Biol. 168, 595 (1983).
[8] W. Kabsch et al., Nature (London) 347, 37 (1990).
[9] M. M. Tirion and D. ben-Avraham, J. Mol. Biol. 230, 186
(1993).
1908
... The resolution of modelling capability ranges from quantum [60], atomistic [61,62,81], coarsegrained [63] to continuum model [64]. To note a few, methods for manipulating the conformational states viz., metadynamics [65], Transition path sampling [66], Normal mode analysis [67], relative entropy coarse graining [68], Discontinuous molecular potentials [69] have been reported. ...
Preprint
Full-text available
We propose a complex dynamical model to mimic the protein-protein interaction. The signalling pathways that activate the proteins enact the intracellular functions are modelled by random and complimentary linking. The molecular roots of diseases are illuminated on the grounds of protein misfolding. Specifically, amy-loid fibrillated states are emphasized that are analogous to the coalescence of ordered and disordered dynamics. Also, a β− pleated sheet can be inferred from the model reflected as an anti-phase synchronized state. Note that the coupling strength is equivalent to the electromagnetic forces of the individual molecules. Critical clustering coupling strength has been analytically derived. Statistical techniques are imposed to impart an energy perspective of the model. We found that the relaxation time of the model plays a vital role in protein folding. We quote that ‘The quicker the folding, the lesser the entropy’. We validate the model using pathological data of breast carcinoma obtained from https://proteinatlas.org.
Article
Full-text available
Monoclonal antibodies (mAbs) can undergo structural changes due to interaction with oil–water interfaces during storage. Such changes can lead to aggregation, resulting in a loss of therapeutic efficacy. Therefore, understanding the microscopic mechanism controlling mAb adsorption is crucial to developing strategies that can minimize the impact of interfaces on the therapeutic properties of mAbs. In this study, we used MARTINI coarse-grained molecular dynamics simulations to investigate the adsorption of the Fab and Fc domains of the monoclonal antibody COE3 at the oil–water interface. Our aim was to determine the regions on the protein surface that drive mAb adsorption. We also investigate the role of protein concentration on protein orientation and protrusion to the oil phase. While our structural analyses compare favorably with recent neutron reflectivity measurements, we observe some differences. Unlike the monolayer at the interface predicted by neutron reflectivity experiments, our simulations indicate the presence of a secondary diffused layer near the interface. We also find that under certain conditions, protein–oil interaction can lead to a considerable distortion in the protein structure, resulting in enhanced adsorption behavior.
Article
The development of effective therapeutics against COVID‐19 requires a thorough understanding of the receptor recognition mechanism of the SARS‐CoV‐2 spike (S) protein. Here the multidomain collective dynamics on the trimer of the spike protein has been analyzed using normal mode analysis (NMA). A common nanomechanical profile was identified in the spike proteins of SARS‐CoV‐2 and its variants. The profile involves collective vibrations of the receptor‐binding domain (RBD) and the N‐terminal domain (NTD), which may mediate the physical interaction process. Quantitative analysis of the collective modes suggests a nanomechanical property involving large‐scale conformational changes, which explains the difference in receptor binding affinity among different variants. These results support the use of intrinsic global dynamics as a valuable perspective for studying the allosteric and functional mechanisms of the S protein. This approach also provides a low‐cost theoretical toolkit for screening potential pathogenic mutations and drug targets.
Article
Full-text available
The Normal Mode Analysis (NMA) is a standard approach to elucidate the anisotropic vibrations of macromolecules at their folded states, where low-frequency collective motions can reveal rearrangements of domains and changes in the exposed surface of macromolecules. Recent advances in structural biology have enabled the resolution of megascale macromolecules with millions of atoms. However, the calculation of their vibrational modes remains elusive due to the prohibitive cost associated with constructing and diagonalizing the underlying eigenproblem and the current approaches to NMA are not readily adaptable for efficient parallel computing on graphic processing unit (GPU). Here, we present eigenproblem construction and diagonalization approach that implements level-structure bandwidth-reducing algorithms to transform the sparse computation in NMA to a globally-sparse-yet-locally-dense computation, allowing batched tensor products to be most efficiently executed on GPU. We map, optimize, and compare several low-complexity Krylov-subspace eigensolvers, supplemented by techniques such as Chebyshev filtering, sum decomposition, external explicit deflation and shift-and-inverse, to allow fast GPU-resident calculations. The method allows accurate calculation of the first 1000 vibrational modes of some largest structures in PDB ( > 2.4 million atoms) at least 250 times faster than existing methods.
Article
Full-text available
We propose a new way of utilizing normal modes to study protein conformational transitions. Instead of considering individual modes independently, we show that a weighted mixture of low-frequency vibrational modes can reveal dynamic information about the conformational mechanism in more detail than any single mode can. The weights in the mixed mode, termed the allosteric covibrational mode, are determined using a simple model where the conformational transition is viewed as a perturbation of the coupled harmonic oscillator associated with either of the two conformations. We demonstrate our theory in a biologically relevant example of high pharmaceutical interest involving the V617F mutation of Janus 2 tyrosine kinase (JAK2).
Article
Interacting many-body physical systems ranging from neural networks in the brain to folding proteins to self-modifying electrical circuits can learn to perform diverse tasks. This learning, both in nature and in engineered systems, can occur through evolutionary selection or through dynamical rules that drive active learning from experience. Here, we show that learning in linear physical networks with weak input signals leaves architectural imprints on the Hessian of a physical system. Compared to a generic organization of the system components, (a) the effective physical dimension of the response to inputs decreases, (b) the response of physical degrees of freedom to random perturbations (or system “susceptibility”) increases, and (c) the low-eigenvalue eigenvectors of the Hessian align with the task. Overall, these effects embody the typical scenario for learning processes in physical systems in the weak input regime, suggesting ways of discovering whether a physical network may have been trained.
Article
Over the past several decades, biomolecular dynamics has been extensively investigated through both theoretical and experimental approaches to uncover the general relationship between the native structure and biological function of proteins. In this review article, we discuss the elastic network model (ENM) and its variations, which are widely used coarse-grained approaches in biomolecular simulations. First, we summarize the pros and cons of various coarse-grained approaches, considering their computational complexity and reliability. Following that, we introduce ENM based harmonic and anharmonic analyses as tools for studying local intrinsic fluctuations and conformational transitions in proteins, respectively. We also provide a list of ENM based web servers that are open to the public. While all-atom simulations have become the gold standard in biomolecular studies today, the ENM based coarse-grained approach remains a reliable tool for long-time, large-scale protein dynamics simulations.
Chapter
This chapter is divided into four parts: (1) General considerations of statistical mechanics and thermodynamics; (2) Introductory electrostatics; (3) Metropolis Monte Carlo (MMC) method; and (4) Examples of application. Part 1 focuses on fundamental concepts with emphasis on the Potential of Mean Force (PMF) and the radial correlation function g(r). Part 2 covers elementary electrostatics, which is important for understanding molecular interactions such as those of triacylglycerols via their partial charges. Part 3 describes the MMC method and addresses the question of whether the sequence of MMC steps is equivalent to the time-development of the system. Finally, Part 4 presents four examples of systems related to oleogels, utilizing computer simulation and experimental techniques, with a detailed example of using the PMF technique. This chapter is essential for those seeking to deepen their understanding of molecular interactions and MMC computer simulation methods with potential applications in the fields of oleogels.
Article
Full-text available
The atomic models of the complex between rabbit skeletal muscle actin and bovine pancreatic deoxyribonuclease I both in the ATP and ADP forms have been determined by X-ray analysis at an effective resolution of 2.8 A and 3A, respectively. The two structures are very similar. The actin molecule consists of two domains which can be further subdivided into two subdomains. ADP or ATP is located in the cleft between the domains with a calcium ion bound to the beta- or beta- and gamma-phosphates, respectively. The motif of a five-stranded beta sheet consisting of a beta meander and a right handed beta alpha beta unit appears in each domain suggesting that gene duplication might have occurred. These sheets have the same topology as that found in hexokinase.
Article
We present the development of a force field for simulation of nucleic acids and proteins. Our approach began by obtaining equilibrium bond lengths and angles from microwave, neutron diffraction, and prior molecular mechanical calculations, torsional constants from microwave, NMR, and molecular mechanical studies, nonbonded parameters from crystal packing calculations, and atomic charges from the fit of a partial charge model to electrostatic potentials calculated by ab initio quantum mechanical theory. The parameters were then refined with molecular mechanical studies on the structures and energies of model compounds. For nucleic acids, we focused on methyl ethyl ether, tetrahydrofuran, deoxyadenosine, dimethyl phosphate, 9-methylguanine-1-methylcytosine hydrogen-bonded complex, 9-methyladenine-1-methylthymine hydrogen-bonded complex, and 1,3-dimethyluracil base-stacked dimer. Bond, angle, torsional, nonbonded, and hydrogen-bond parameters were varied to optimize the agreement between calculated and experimental values for sugar pucker energies and structures, vibrational frequencies of dimethyl phosphate and tetrahydrofuran, and energies for base pairing and base stacking. For proteins, we focused on Φ,Ψ maps of glycyl and alanyl dipeptides, hydrogen-bonding interactions involving the various protein polar groups, and energy refinement calculations on insulin. Unlike the models for hydrogen bonding involving nitrogen and oxygen electron donors, an adequate description of sulfur hydrogen bonding required explicit inclusion of lone pairs.
Article
The atomic models of the complex between rabbit skeletal muscle actin and bovine pancreatic deoxyribonuclease I both in the ATP and ADP forms have been determined byo X-ray analysis at an effective resolution of 2.8 Å and 3 Å, respectively. The two structures are very similar. The actin molecule consists of two domains which can be further subdivided into two subdomains. ADP or ATP is located in the cleft between the domains with a calcium ion bound to the β- or β- and γ-phosphates, respectively. The motif of a five-stranded (3 sheet consisting of a (3 meander and a right handed βαβ unit appears in each domain suggesting that gene duplication might have occurred. These sheets have the same topology as that found in hexokinase.
Article
A dynamic structure refinement method for X-ray crystallography, referred to as the normal mode refinement, is proposed. The Debye-Waller factor is expanded in terms of the low-frequency normal modes whose amplitudes and eigenvectors are experimentally optimized in the process of the crystallographic refinement. In this model, the atomic fluctuations are treated as anisotropic and concerted. The normal modes of the external motion (TLS model) are also introduced to cover the factors other than the internal fluctuations, such as the lattice disorder and diffusion. A program for the normal mode refinement (NM-REF) has been developed. The method has first been tested against simulated diffraction data for human lysozyme calculated by a Monte Carlo simulation. Applications of the method have demonstrated that the normal mode refinement has: (1) improved the fitting to the diffraction data, even with fewer adjustable parameters; (2) distinguished internal fluctuations from external ones; (3) determined anisotropic thermal factors; and (4) identified concerted fluctuations in the protein molecule.
Article
A method is presented whereby the amplitude coefficients of molecular normal modes of vibration are treated as independent variables in the treatment of thermal effects in X-ray diffraction, and applied to the bovine pancreatic trypsin inhibitor, form II (P2(1)2(1)2(1), a = 74.1, b = 23.4, c = 28.9 A). It is shown that the description of molecular motion furnished by 892 isotropic temperature factors may be largely reproduced using only 19 molecular thermal parameters from which anisotropic temperature factors may be synthesised for every atom. The method shows that motions and/or disorders external to each molecule are the largest single source of apparent motion, and that the internal motions are comparable to those predicted by Levitt, Sander & Stern.
Article
We have developed a new method for modelling protein dynamics using normal-mode analysis in internal co-ordinates. This method, normal-mode dynamics, is particularly well suited for modelling collective motion, makes possible direct visualization of biologically interesting modes, and is complementary to the more time-consuming simulation of molecular dynamics trajectories. The essential assumption and limitation of normal-mode analysis is that the molecular potential energy varies quadratically. Our study starts with energy minimization of the X-ray co-ordinates with respect to the single-bond torsion angles. The main technical task is the calculation of second derivative matrices of kinetic and potential energy with respect to the torsion angle co-ordinates. These enter into a generalized eigenvalue problem, and the final eigenvalues and eigenvectors provide a complete description of the motion in the basic 0.1 to 10 picosecond range. Thermodynamic averages of amplitudes, fluctuations and correlations can be calculated efficiently using analytical formulae. The general method presented here is applied to four proteins, trypsin inhibitor, crambin, ribonuclease and lysozyme. When the resulting atomic motion is visualized by computer graphics, it is clear that the motion of each protein is collective with all atoms participating in each mode. The slow modes, with frequencies of below 10 cm-1 (a period of 3 ps), are the most interesting in that the motion in these modes is segmental. The root-mean-square atomic fluctuations, which are dominated by a few slow modes, agree well with experimental temperature factors (B values). The normal-mode dynamics of these four proteins have many features in common, although in the larger molecules, lysozyme and ribonuclease, there is low frequency domain motion about the active site.
Article
The motion of the atoms in the small protein bovine pancreatic trypsin inhibitor has been simulated for about 60 picoseconds using two different potential energy functions. In one known as HHL the hydrogen bond is purely electrostatic, in the other, known as L79, the hydrogen bond is a directional O . . . H interaction. The energy parameters and techniques used to obtain an accurate, well-equilibrated trajectory are described in detail. The trajectories calculated here with either potential are superior to those obtained in previous simulations on the same protein in that they treat hydrogen bonding realistically and remain closer to the native X-ray structure. Comparison of the two trajectories shows that the potential energy parameters have a significant effect on the shift from the X-ray structure, the distribution of (phi, psi) torsion angles, the pattern of hydrogen bonds and the accessible surface area of individual residues. The L79 potential with directional hydrogen bonds is used to simulate a longer 132 picosecond trajectory that is analysed in the accompanying paper.
Article
A normal mode analysis making use of an empirical potential function including local and nonlocal (nonbonded) interactions is performed for the bovine pancreatic trypsin inhibitor in the full conformational space of the molecule (1,740 degrees of freedom); that is, all bond lengths and angles, as well as dihedral angles, are included for the 580-atom system consisting of all heavy atoms and polar hydrogens. The heavy-atom frequency spectrum shows a dense distribution between 3 and 1,800 cm-1, with 350 modes below 216 cm-1. Most of the low-frequency modes, of which many have significant anharmonic character, are found to be delocalized over the protein. The root-mean-square amplitudes of the atomic fluctuations are calculated at 300 K from the normal modes and compared with those obtained from a solution molecular dynamics simulation based on the same potential function; very good agreement is obtained for the variation in the main-chain fluctuations as a function of residue number, though larger differences occur for the side chains. The fluctuations are generally, though not always, dominated by frequencies below 30 cm-1, in accord with the results of the dynamics simulation. The vibrational contributions to the thermodynamic properties of the protein are calculated as a function of temperature; the effects of perturbations on the spectrum, suggested for ligand or substrate binding, are examined. The analysis demonstrates that, in spite of the anharmonic contributions to the potential, a normal mode description can provide useful results concerning the internal motions of proteins.
Article
Correlated motions of protein atoms are of biological significance in processes involving ligand binding, conformational change and information transmission. X-ray scattering patterns from protein crystals contain diffuse scattering that originates from correlated displacements of atoms. Here we present experimental data on diffuse X-ray scattering from lysozyme crystals. We show that the diffuse scattering is similar in form to scattering derived from molecular dynamics simulation and normal mode analysis of the isolated protein, the normal modes giving the closest agreement with experiment.
Article
The slow normal modes of G-actin were used as structural parameters to refine the F-actin model against 8-A resolution x-ray fiber diffraction data. The slowest frequency normal modes of G-actin pertain to collective rearrangements of domains, motions that are characterized by correlation lengths on the order of the resolution of the fiber diffraction data. Using a small number of normal mode degrees of freedom (< or = 12) improved the fit to the data significantly. The refined model of F-actin shows that the nucleotide binding cleft has narrowed and that the DNase I binding loop has twisted to a lower radius, consistent with other refinement techniques and electron microscopy data. The methodology of a normal mode refinement is described, and the results, as applied to actin, are detailed.