ArticlePDF Available

Molecular Dynamics Simulations of p53 DNA-Binding Domain

Authors:

Abstract and Figures

We have studied room-temperature structural and dynamic properties of the p53 DNA-binding domain in both DNA-bound and DNA-free states. A cumulative 55 ns of explicit solvent molecular dynamics simulations with the particle mesh Ewald treatment of electrostatics was performed. It was found that the mean structures in the production portions of the trajectories agree well with the crystal structure: backbone root-mean-square deviations are in the range of 1.6 and 2.0 A. In both simulations, noticeable backbone deviations from the crystal structure are observed only in loop L6, due to the lack of crystal packing in the simulations. More deviations are observed in the DNA-free simulation, apparently due to the absence of DNA. Computed backbone B-factor is also in qualitative agreement with the crystal structure. Interestingly, little backbone structural change is observed between the mean simulated DNA-bound and DNA-free structures. A notable difference is observed only at the DNA-binding interface. The correlation between native contacts and inactivation mechanisms of tumor mutations is also discussed. In the H2 region, tumor mutations at sites D281, R282, E285, and E286 may weaken five key interactions that stabilize H2, indicating that their inactivation mechanisms may be related to the loss of local structure around H2, which in turn may reduce the overall stability to a measurable amount. In the L2 region, tumor mutations at sites Y163, K164, E171, V173, L194, R249, I251, and E271 are likely to be responsible for the loss of stability in the protein. In addition to apparent DNA contacts that are related to DNA binding, interactions R175/S183, S183/R196, and E198/N235 are highly occupied only in the DNA-bound form, indicating that they are more likely to be responsible for DNA binding.
Content may be subject to copyright.
Molecular Dynamics Simulations of p53 DNA-Binding Domain
Qiang Lu1, Yu-Hong Tan1, and Ray Luo*
Department of Molecular Biology and Biochemistry, University of California, Irvine, CA 92697-3900
Abstract
We have studied room-temperature structural and dynamic properties of the p53 DNA-binding
domain in both DNA-bound and DNA-free states. A cumulative 55ns of explicit solvent molecular
dynamics simulations with the Particle Mesh Ewald treatment of electrostatics were performed. It is
found that the mean structures in the production portions of the trajectories agree well with the crystal
structure: backbone root-mean squared deviations are in the range of 1.6Å and 2.0Å. In both
simulations, noticeable backbone deviations from the crystal structure are observed only in loop L6,
due to the lack of crystal packing in the simulations. More deviations are observed in the DNA-free
simulation, apparently due to the absence of DNA. Computed backbone B-factor is also in qualitative
agreement with the crystal structure. Interestingly little backbone structural change was observed
between the mean simulated DNA-bound and DNA-free structures. Notable difference is only
observed at the DNA-binding interface. The correlation between native contacts and inactivation
mechanisms of tumor mutations is also discussed. In the H2 region, tumor mutations at sites D281,
R282, E285, and E286 may weaken five key interactions that stabilize H2, indicating that their
inactivation mechanisms may be related to the loss of local structure around H2, which in turn may
reduce the overall stability to a measurable amount. In the L2 region, tumor mutations at sites Y163,
K164, E171, V173, L194, R249, I251 and E271 are likely to be responsible for the loss of stability
in the protein. In addition to apparent DNA contacts that are related to DNA binding, interactions
R175/S183, S183/R196, and E198/N235 are highly occupied only in the DNA-bound form,
indicating that they are more likely to be responsible for DNA binding.
INTRODUCTION
p53 is a transcription factor that binds to specific DNA sequences1–4 to activate gene
expression.5–9 In response to activation of certain oncogenes, it can stop cell cycle, prevent
genetic alterations, and induce apoptosis or programmed cell death.10–12 Mutation in the p53
gene is one of the most frequent events in the process of oncogenesis. It is estimated that
approximately 50% of human tumors contain mutations in this pivotal gene.13 The in vivo
importance of p53 transactivation is underscored by the fact that 95% of all known tumor
mutations occur in the DNA-binding domain of the protein.13,14 More interestingly, 75% of
these mutations occur as single missense mutations.13,14 Therefore, the oncogenic form of
p53 is predominantly a full-length protein with a single amino-acid substitution in its DNA-
binding domain. Tumors with inactive p53 mutants are aggressive and often resistant to
ionizing radiation and chemotherapy. Clearly, p53 mutants present one of the most important
clinical targets for drug intervention of tumors. Many experimental approaches are in progress
to develop p53-based therapies.15,16
p53 bears the usual hallmarks of a transcription factor, with an amino-terminal transactivation
domain, a DNA-binding domain, and carboxy-terminal tetramerization and regulatory
* Please send correspondence to: R. Luo. E-mail: rluo@uci.edu. FAX: (949) 856-8551.
1These authors contributed equally to this work.
NIH Public Access
Author Manuscript
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
Published in final edited form as:
J Phys Chem B. 2007 October 4; 111(39): 11538–11545. doi:10.1021/jp0742261.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
domains. The crystal structure of the DNA-binding domain, solved in 1994, provides a starting
point for understanding the nature of mutant p53 (Fig. 1).17 The structure consists of a β-
sandwich scaffold and a DNA-binding surface, including a loop-sheet-helix motif (LSH) and
two loops (L2 and L3) tethered by a single Zinc atom. The loops and the LSH motif form the
DNA-binding surface of p53 and provide contacts to the DNA backbone and bases.
Due to the large number of mutations occurring at the p53 DNA-binding domain, it is
instructive to classify these mutations based on their locations for better understanding of their
effects for drug intervention. It appears that 30% of all mutations fall on six mutation hot spots,
and these hot spots cluster to the DNA-binding surface: two contact DNA directly, R248 (L3)
and R273 (LSH), and four stabilize the surrounding structure, R175 (L2), G245 (L3), R249
(L3) and R282 (LSH).17 These give rise to two classes of mutations, DNA-contact and
structural. At the physiological condition, all six hot spots fail to bind target DNA sequences.
Further classifications of mutants are possible after considering their thermodynamic
properties. Overall, three broad phenotypes are proposed: DNA-contact mutations that have
little or no effect on stability, mutations that disrupt local structure, and mutations that cause
denaturation.17,18 To date only a small fraction of tumor mutations are analyzed by
experimental methods. The inactivation mechanisms of many tumor mutations are still
unknown.17,18
This article reports our effort in the computational analysis of this important protein.
Specifically, molecular dynamics simulations in explicit solvent have been used to study the
dynamic properties of the wild type protein to infer the inactivation mechanisms of tumor
mutations.
METHODS
There are two computational difficulties in molecular dynamics simulations of the p53 DNA-
binding domain (core domain, p53c). The first difficulty is its large size: a monomer p53c is
already around 50Å in diameter, but p53 functions as a tetramer in cell. Simulation of the
complete tetramer p53c in explicit water would be extremely difficult considering the number
of water molecules needed to solvate the protein. Fortunately, it has been shown that the
monomer p53c still binds to DNA, though weaker.19 This indicates that the monomer p53c
can be used a model to study this important interaction. Further, all existing biophysical
analysis of the wild type and mutants of p53c show that the monomer p53c is stable in vitro,
so that the monomer model is also suitable for its stability analysis.18
The second difficulty is accurate modeling of the Zinc-binding interface in p53c. Previous
experiments with metal chelators showed that p53 depends upon the coordination of Zinc for
both correct folding and correct binding to specific DNA in intact cells, especially for binding
in the DNA minor groove.21,22 Therefore the Zinc-binding interface is important for
simulation of p53c and cannot be neglected. Two approaches have been developed for
modeling Zinc coordination interfaces in proteins. The non-bonded approach uses electrostatic
and van der Waals interactions to maintain four-ligand or five-ligand coordinations of Zinc in
proteins.23–26 It is suitable to simulate proteins in which Zinc is required for catalytic function,
though it is quite challenging to maintain the interface stability at room temperature during
long-time simulations. The bonded approach uses covalent bonds between Zinc and its
coordinating ligands to maintain desired geometries.27–29 The advantage of the bonded
approach is that it is easier to preserve desired geometries as in crystal structures and maintain
stability at room temperature by choosing suitable bond and angle parameters. The
disadvantage of this approach is that it cannot model bond breaking between Zinc and ligands.
In this study the bonded approach is chosen in our computational simulations of p53c.
Lu et al. Page 2
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Quantum mechanical analysis of the Zinc interface
The Zinc interface in p53c, [Zn(CYS)3(HIS)1], is chopped out from the crystal structure 1TSR
B (Fig. 1). To minimize perturbation to the interface, four main chain Cα atoms are included
in the model compound for optimization of force field parameters. Since there is no hydrogen
in the crystal structure to determine the total charge, all possible net charges, 2, 1, 0 and 1,
2, are investigated (Table 1).
The geometries of all tested protonation states were optimized with tight option using rational
function optimization (RFO) method.30 After the RFO method, direct inversion in the iterative
subspace (GDIIS)31,32 was used to reach a minimum. B3LYP density functional method33,
34 with the 6-311+G** basis set was employed in the optimization. Finally B3LYP/6-311+G**
was chosen for frequency analysis because it was reported to be successful in reproducing the
hydration energy of Zinc divalent cation35,36 and in studying the farnesyltransferase
inhibitors.37 A scaling factor 0.989 was used for frequency calculation with the basis set 6-311
+G**.38 B3LYP/6-311+G** was also used to obtain ESP potential39–41 in the atomic charge
derivation. All quantum mechanical calculations were performed using Gaussian03(G03)42
at NCSA. It is found that only the protonation state with 1 net charge keeps the Zinc interface
structure intact. The structures with net charges of 2, 1, 0 and 2 lose the tetrahedral structure
after 20 steps of optimization. The final optimized structure of 1 charge is shown in Fig. 2.
Interestingly, the same protonation state with three thiolate cystine residues coordinated with
Zinc were found to facilitate the fourth thiolate cyctine residue to be a chemically active
nucleophilic species in the methyl transfer reactions.20
Molecular mechanics analysis of the Zinc interface
Molecular mechanical optimizations were performed with NMODE43 in Amber944 in order
to obtain normal mode frequencies. The initial geometry is the optimized structure from
G03.42 The Newton-Raphson method was applied to locate the minimum. The convergence
criterion is 106 kcal/mol-Å.
Two bonds and five angle parameters for the Zinc four-ligand coordination interface in p53c
[Zn(CYS)3(HIS)1] were empirically optimized based on the comparison of frequencies
between Amber and G03. We have set all Zinc-related dihedral terms to zero to simplify
parameterization. The van der Walls parameters for the Zinc atom were taken from the
literature: σ = 1.10Å and ε = 0.0125kcal/mol.45 RESP charges46–48 were derived from the
G03 ESP potential with ANTECHAMBER in Amber9.44 Partial charges of CYS (CA,
0.0396; HA1, 0.002; HA2, 0.002; HA3, 0.002; CB, 0.2061; HB1, 0.0231; HB2,
0.0231; SG, 0.7380) and partial charges of HIS (CA, 0.1523; HA1, 0.0249; HA2, 0.0249;
HA3, 0.0249; CB, 0.0693; HB1, 0.0694; HB2, 0.0694; CG, 0.1183; ND1, 0.0151; ZN,
0.8024; CE1, 0.0741; HE1, 0.0950; NE2, 0.3378; HE2, 0.3347; CD2, 0.1782; HD2, 0.1611)
were obtained. The total charge is 1 according to the QM analysis.
Molecular dynamics simulation of DNA-free and DNA-bound p53 core domain
As reviewed, the monomer p53c can bind DNA with approximately one-fifth the affinity of
intact p53.19 This gives us an opportunity to study the DNA binding of p53c with a monomer
model. In our molecular dynamics simulation, the initial complex structure is derived from the
p53c (1TSR chain B) in complex with DNA (5-TAGACTTGCCCA-3).17 1TSR chain B is
chosen as the initial DNA-free p53c structure to study the difference between DNA-bound and
DNA-free structures. Hydrogen atoms were added using LEAP of Amber9.44 In both
simulations the solute molecules were solvated in a TIP3P water box with a 10 Å buffer.49
Both systems were neutralized by adding counter ions. The detailed information of the two
simulation systems are listed in Table 2.
Lu et al. Page 3
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
A revised parm99 force field was used for intramolecular interactions.50,51 Particle Mesh
Ewald was used to treat long-range electrostatics with the default setting in Amber9.52 All
bonds involving hydrogens were constrained by SHAKE,53 so that a time step of 2fs was used
for dynamics integration. Before simulation, the initial coordinates were relaxed by 1000 steps
steepest-descent energy minimization. After minimization the two systems were heated up to
80K in order to adjust the water configurations, while the backbone atoms of protein and DNA
were restrained by weak harmonic forces. Finally the systems were heated up to 293K within
20ps. In the complex simulation, the backbone of DNA was weakly restrained by a weak force
of 1kcal/mol-Å2 throughout the simulation because it is truncated. Constant temperature
(293K) and constant pressure (1 bar) were maintained by the Berendsen’s method.54 A total
of 55ns trajectories at 293K were collected for both systems.
RESULTS AND DISCUSSION
Quality of covalent force field terms
Projections of normal modes onto internal coordinates agree well between G03 and Amber.
The distributions of normal modes over Zinc-related bonds and angles are listed in Supporting
Material (Fig. S-1). Note that variation of Zinc-related parameters only influences normal
modes less than 500 cm1.
Table 3 lists all optimized force parameters of Zinc-related bonds and angles. Literature values
mainly come from Human Carbonic Anhydrase II (3HIS, 1H2O)27,28 and Alcohol
Dehydrogenase (2CYS, 1HIS, 1H2O).55 NA-ZN bond force constant, 40 kcal/mol-Å2, comes
from experimental data.56–58
Similar structures are obtained by Amber and G03 minimization as shown in Fig. 3. After
parameterization, comparison of frequencies over normal modes is performed and shown in
Fig. 4. Frequencies from 0 to 500cm1 correspond to the Zinc-related covalent terms, whose
RMSD is 14.4cm1 between G03 and Amber. The RMSD is 34.50cm1 for frequencies below
1,000 cm1. Overall an excellent agreement can be observed between Amber and G03.
Agreement between simulation and crystal structures
Molecular dynamics simulations of both DNA-bound and DNA-free proteins were run up to
30ns and 25ns, respectively. Their accumulate average potential energies show that the DNA-
bound trajectory has equilibrated after 22.5ns and the DNA-free trajectory has equilibrated
after 5ns, respectively (Fig. 5). Stable structures are observed in both trajectories, as reflected
in their backbone RMSD fluctuations (Fig. 6a).
The backbone RMSD, including all loop residues, of the DNA-bound protein with respect to
chain B of 1TSR (the one in complex with DNA) is about 1.6Å (Fig. 6a). While the backbone
RMSD of the DNA-free protein is around 2.0Å (Fig. 6a), which is higher than that of the
complex, apparently due to the absence of binding to DNA. The mean structures in the
production portion of the trajectories (7.5ns for DNA-bound and 20ns for DNA-free) also agree
with chain B of 1TSR well: the backbone RMSD of the DNA-bound and DNA-free proteins
is 1.61Å and 2.07Å, respectively (Fig. 7). This should be compared with the RMSD differences
among the three chains in 1TSR, around 0.85Å. Considering that the simulations were
performed at room temperature and without crystal packing, the agreement between simulation
and experiment is very good. In both simulations, noticeable backbone deviations from the
crystal structure are observed only in loop L6. This is due to the lack of crystal packing in the
monomer simulations. In the DNA-free simulation, deviations are also observed in L1, H1,
and L5, apparently due to the absence of DNA. In Fig. 7 one can find that the L1 DNA-binding
interface has noticeably deformed in the DNA-free simulation.
Lu et al. Page 4
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Computed backbone B-factor of the DNA-bound protein is in qualitative agreement with that
of chain B in 1TSR (Fig. 8). Major differences between simulation and experiment are all due
to the existence of crystal packing with other chains in 1TSR (Fig. 8). Interestingly, computed
backbone B-factor of the DNA-free protein agrees better with chain A that does not have
specific DNA contacts (Fig. 9). This indicates that chain A might be used as a model of the
DNA-free protein for dynamics analysis. However, it should be pointed out that chain A is a
less accurate structural model than chain B.
RMSD fluctuation of the heavy atoms at the Zinc interface indicates that the tetrahedron
structure is maintained as in 1TSR (Fig. 6b). Further, relative B-factor of the side-chain heavy
atoms with respect to those of the backbone heavy atoms show that the bonded model adopted
here can reproduce the dynamic properties of the Zinc interface in p53c at least for equilibrium
properties at the native state (Table 4).
Differences between DNA-bound and DNA-free room-temperature structures
One of our interests is to understand the binding-induced conformational change in p53c at
room temperature. As shown in Fig. 7, overall there is little change in the backbone between
the mean DNA-bound and DNA-free structures, similar to that observed between crystal
structures 1TSR B and A chains at 98K. Notable difference in the backbone is only observed
in L1, H1, and L3 at the DNA-binding interface.
A detailed comparison of the arrangement of secondary structures in DNA-bound and DNA-
free structures is shown in Fig. 10. The major differences are only observed in places close to
the DNA-binding interface. On the H2 side of the binding interface, the only difference is at
L1 that moves away from H2 in the DNA-free structure. On the L3 side of the binding interface,
L3 is slightly distorted when compared with that in the DNA-bound structure. Another
difference is observed in loop L2, a central structure connecting L3 with L2, H1, NT, L5, and
portion of the β sandwich (Fig. 7). In fact, L2 has most contacts with other secondary structural
units in the DNA-binding region (Table 5). In the DNA-bound structure L2 is closer to L2 and
L3, but it is farther away from L5 and the β sandwich. In addition, the contacts between L2/
NT and L5 are weaker when compared with those in the DNA-free structure (Table 5),
corresponding to the observed higher B-factor on NT and L5 in the complex (Fig. 8). A similar
difference on NT and L5 is also observed in the crystal structure 1TSR. NT is packed well in
chain A but not in chain B, and the B-factor of NT is also lower in chain A than in chain B
(Fig. 11). These differences indicate that contacts involving L2 may be important to DNA
binding even if they are not in direct contact with DNA.
The difference of native contacts between the DNA-bound and DNA-free protein is also
compared to understand the binding-induced conformational changes. This gives us the basis
for the following analysis related to common tumor mutations. In the DNA-binding region, all
side-chain/side-chain or side-chain/main-chain contacts are monitored. Those contacts with
percentage occupancies higher than 50% in the DNA-bond structure or the DNA-free structure
are listed in Table 5 and Table 6. Comparison between the contacts in the DNA-bound and
DNA-free structures reveals many differences. Interestingly, there are ten contacts
significantly weakened by binding to DNA (V97/M169, S116/C124, M169/R213, V173/A161,
R174/E180, R175/D184, D184/R196, D186/R196, L188/Y201, L194/Y236), but only five
contacts are significantly strengthened by binding to DNA (T118/R283, M169/I162, R175/
S183, S183/R196, L188/V203) outside the β sandwich (Table 5). In contrast, few differences
are found in the β-sandwich region as shown in Table 6. Specifically, four contacts (T140/
N235, V143/I255 and E198/H233. I251/V272) are weakened, and two contacts are
strengthened (V143/F270 and E198/N235) by DNA binding.
Lu et al. Page 5
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Implications to common tumor mutations
A more detailed analysis of the native contacts indicates that there are additional differences
between the two room-temperature simulations. These differences may shed lights in
understanding the correlation between tumor mutations and DNA binding or folding stability
in p53c. In the following, we discuss the correlation of native contacts with inactivation
mechanisms of certain tumor mutations in four different regions: β sandwich, H2 region, L2
region, and DNA-binding region.
β sandwich—It is apparent that mutations in the β-sandwich region (hydrophobic core) are
more likely to be responsible for the protein’s overall stability: most native contacts are
hydrophobic in nature. Instead of analyzing all contacts generically, we have chosen to focus
on two centers of the hydrophobic core: I195 and V143. Tumor mutations targeting these two
centers can severely destabilize the p53 core domain (mutation I195T and V143A destabilize
p53c by 4.12 kcal/mol and 3.50 kcal/mol, respectively).59 I195 forms hydrophobic contacts
with A161, Y234, and Y236. None of these change upon DNA binding (Table 6). V143 forms
hydrophobic contacts with L111, Y234, I255, and F270 (Table 6). Most of these hydrophobic
contacts are also well preserved in both simulations.
H2 Region—The inactivation mechanisms of tumor mutations in other parts of p53c are less
clear. Here we have monitored all side-chain contacts to track their occupancies at room-
temperature. In the H2 region, tumor mutations at sites D281, R282, E285, and E286 may
weaken five key interactions that stabilize H2 (Table 5). Our data indicates that their
inactivation mechanism may be related to the loss of local structure around H2, which in turn
may reduce the overall stability to a measurable amount. Our analysis on contacts related to
H2 is in part supported by the structure and stability data from Fersht and co-workers: mutation
R282W affects the packing of Loop-Sheet-Helix,60 and destabilizes p53c by 3.3 kcal/mol.59
L2 Region—In the L2 region, contacts K164/E271, V173/Y163, V173/L194, V173/I251,
and R249/E171 are highly occupied in both DNA-bound and DNA-free simulations. Our
contact analysis thus implies that mutations at sites Y163, K164, E171, V173, L194, R249,
I251 and E271 are likely to be responsible for the loss of stability in p53c. The finding for
R249 is in agreement with available structure and stability data from Fersht and co-workers:
mutation R249S was found to induce substantial structural perturbation around mutation site
in loop L3,61 and to destabilize p53c by 1.92 kcal/mol.59
DNA binding—Many interactions are in direct contact with DNA (Table 7), indicating that
they are responsible for DNA binding. Except N239 the results are in agreement with the crystal
structure from Cho et al: residue K120, S241, R248, R273, R280, and R283 are direct in contact
with DNA.17 In addition, the crystal structure from Fersht and co-workers also shows that
R273C and R273H mutations simply remove DNA contacts without perturbing the
conformation of nearby residues.60,61
In addition to apparent DNA contacts that are related to DNA binding, many interactions are
highly occupied only in the DNA-bound form, indicating that they are more likely to be
responsible for DNA binding. Three such contacts that are not in direct contact with DNA are
found: R175/S183, S183/R196 (Table 5), and E198/N235 (Table 6).
CONCLUDING REMARKS
With optimized force field parameters for the Zinc-binding interface of p53c, a cumulative
55ns of room-temperature molecular dynamics simulations in explicit solvent with the PME
treatment of electrostatics were performed for both DNA-bound and DNA-free p53c. It is found
that the backbone RMSDs, including all loop residues, of both DNA-bound and DNA-free
Lu et al. Page 6
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
p53c with respect to chain B in 1TSR are from 1.6Å to 2.0Å. The mean structures in the
production portions of the trajectories also agree with chain B of 1TSR well. In both
simulations, noticeable backbone deviations from the crystal structure are observed only in
loop L6. This is due to the lack of crystal packing in both simulations. In the DNA-free
simulation, deviations are also observed in L1, H1, and L5, apparently due to the absence of
DNA. Computed backbone B-factor of the DNA-bound protein is also in qualitative agreement
with that of chain B in 1TSR. Major differences between simulation and experiment are due
to the existence of crystal packing with other chains in 1TSR. RMSD fluctuation of the heavy
atoms at the Zinc-binding interface indicates that the tetrahedron structure is maintained as in
1TSR. Relative B-factor of the side-chain heavy atoms with respect to those of the backbone
heavy atoms at the Zinc-binding interface also agrees well between simulation and crystal
structures, indicating that the bonded model adopted here may reproduce equilibrium dynamic
properties of the Zinc-binding interface in p53c.
Our room-temperature simulation shows little backbone change between the mean DNA-
bound and DNA-free structures, similar to that observed between crystal structures 1TSR B
and A chains at 98K. Notable difference in the backbone is only observed on the DNA-binding
interface. On the H2 side of the binding interface, the only difference is L1 that moves away
from H2 in the DNA-free structure. On the L3 side of the binding interface, L3 is slightly
distorted when compared with that in the DNA-bound structure. Another difference is observed
in loop L2. In the DNA-bound structure L2 is closer to L2 and L3, but it is farther away from
L5 and the β sandwich. In addition, the contacts between L2/NT and L5 are weaker when
compared with those in the DNA-free structure, corresponding to the observed higher B-factor
on NT and L5 in the complex. A similar difference on NT and L5 is also observed in the
crystal structure 1TSR. NT is packed well in chain A but not in chain B, and the B-factor of
NT is also lower in chain A than in chain B. These differences indicate that contacts that involve
L2 may be important to DNA-binding even if they are not in direct contact with DNA.
The correlation between native contacts and inactivation mechanisms of certain tumor
mutations is also discussed. It is apparent that mutations in the β-sandwich region (hydrophobic
core) are more likely to be responsible for the protein’s overall stability: most native contacts
are hydrophobic in nature. The inactivation mechanisms of tumor mutations in other parts of
p53c are less clear. Here we have monitored all side-chain contacts to track their occupancies
at room-temperature. In the H2 region, tumor mutations at sites D281, R282, E285, and E286
may weaken five key interactions that stabilize H2. Our data indicates that their inactivation
mechanism may be related to the loss of local structure around H2.
In the L2 region, contacts K164/E271, V173/Y163, V173/L194, V173/I251, and R249/E171
are highly occupied in both DNA-bound and DNA-free simulations. Our contact analysis thus
implies that mutations at sites Y163, K164, E171, V173, L194, R249, I251 and E271 are likely
to be responsible for the loss of stability in p53c.
Many interactions are in direct contact with DNA, indicating that they are responsible for DNA
binding. In addition to apparent DNA contacts that are related to DNA binding, many
interactions are highly occupied only in the DNA-bound form, indicating that they are more
likely to be responsible for DNA binding. Three such contacts that are not in direct contact
with DNA are found: R175/S183, S183/R196, and E198/N235.
Supplementary Material
Refer to Web version on PubMed Central for supplementary material.
Lu et al. Page 7
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Acknowledgements
This work is supported in part by NIH (GM069620) and the State of California (CRCC).
References
1. Bargonetti J, Friedman PN, Kern SE, Vogelstein B, Prives C. Cell 1991;65:1083. [PubMed: 1646078]
2. Kern SE, Kinzler KW, Bruskin A, Jarosz D, Friedman P, Prives C, Vogelstein B. Science
1991;252:1708. [PubMed: 2047879]
3. Funk WD, Pak DT, Karas RH, Wright WE, Shay JW. Molecular and Cellular Biology 1992;12:2866.
[PubMed: 1588974]
4. Eldeiry WS, Kern SE, Pietenpol JA, Kinzler KW, Vogelstein B. Nature Genetics 1992;1:45. [PubMed:
1301998]
5. Fields S, Jang SK. Science 1990;249:1046. [PubMed: 2144363]
6. Raycroft L, Wu H, Lozano G. Science 1990;249:1049. [PubMed: 2144364]
7. Farmer G, Bargonetti J, Zhu H, Friedman P, Prywes R, Prives C. Nature 1992;358:83. [PubMed:
1614538]
8. Kern SE, Pietenpol JA, Thiagalingam S, Seymour A, Kinzler KW, Vogelstein B. Science
1992;256:827. [PubMed: 1589764]
9. Unger T, Nau MM, Segal S, Minna JD. Embo Journal 1992;11:1383. [PubMed: 1314165]
10. Harris CC. Journal of the National Cancer Institute 1996;88:1442. [PubMed: 8841019]
11. Ko LJ, Prives C. Genes & Development 1996;10:1054. [PubMed: 8654922]
12. Prives C, Hall PA. Journal of Pathology 1999;187:112. [PubMed: 10341712]
13. Martin ACR, Facchiano AM, Cuff AL, Hernandez-Boussard T, Olivier M, Hainaut P, Thornton JM.
Human Mutation 2002;19:149. [PubMed: 11793474]
14. Olivier M, Eeles R, Hollstein M, Khan MA, Harris CC, Hainaut P. Human Mutation 2002;19:607.
[PubMed: 12007217]
15. Nikolova PV, Wong KB, DeDecker B, Henckel J, Fersht AR. Embo Journal 2000;19:370. [PubMed:
10654936]
16. Hupp TR, Sparks A, Lane DP. Cell 1995;83:237. [PubMed: 7585941]
17. Cho YJ, Gorina S, Jeffrey PD, Pavletich NP. Science 1994;265:346. [PubMed: 8023157]
18. Bullock AN, Fersh AR. Nature Reviews Cancer 2001;1:68.
19. Pavletich NP, Chambers KA, Pabo CO. Genes & Development 1993;7:2556. [PubMed: 8276238]
20. Min DH, Liu YS, Carbone I, Yang W. Journal of Chemical Physics 2007;126
21. Hainaut P, Mann K. Antioxidants & Redox Signaling 2001;3:611. [PubMed: 11554448]
22. Meplan C, Richard MJ, Hainaut P. Biochemical Pharmacology 2000;59:25. [PubMed: 10605931]
23. Vedani A, Huhta DW. Journal of the American Chemical Society 1990;112:4759.
24. Aqvist J, Warshel A. Journal of the American Chemical Society 1990;112:2860.
25. Stote RH, Karplus M. Proteins-Structure Function and Genetics 1995;23:12.
26. Wasserman ZR, Hodge CN. Proteins-Structure Function and Genetics 1996;24:227.
27. Hoops SC, Anderson KW, Merz KM. Journal of the American Chemical Society 1991;113:8262.
28. Lu DS, Voth GA. Proteins-Structure Function and Genetics 1998;33:119.
29. Pang YP. Journal of Molecular Modeling 1999;5:196.
30. Simons J, Jorgensen P, Taylor H, Ozment J. Journal of Physical Chemistry 1983;87:2745.
31. Csaszar P, Pulay P. Journal of Molecular Structure 1984;114:31.
32. Farkas O, Schlegel HB. Journal of Chemical Physics 1999;111:10806.
33. Lee CT, Yang WT, Parr RG. Physical Review B 1988;37:785.
34. Becke AD. Journal of Chemical Physics 1993;98:5648.
35. Hehre, WJ.; Radom, L.; Schleyer, PVR.; Pople, J. Ab initio molecular orbital theory. New York:
1986.
36. Pavlov M, Siegbahn PEM, Sandstrom M. Journal of Physical Chemistry A 1998;102:219.
Lu et al. Page 8
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
37. Pang YP, Xu K, Yazal JE, Prendergast FG. Protein Science 2000;9:2583.
38. Foresman, AFJB. Exploring Chemistry with Electronic Structure Methods. 2. Gaussian; 1996.
39. Momany FA. Journal of Physical Chemistry 1978;82:592.
40. Cox SR, Williams DE. Journal of Computational Chemistry 1981;2:304.
41. Singh UC, Kollman PA. Journal of Computational Chemistry 1984;5:129.
42. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery JA,
TVreven J, Kudin KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, et al. Gaussian 03. 2003
43. Kottalam J, Case DA. Biopolymers 1990;29:1409. [PubMed: 2361153]
44. Case DA, Darden TA, Cheatham TEI, Simmerling CL, Wang J, Duke RE, Luo R, Merz KM, Pearlman
DA, Crowley M, Walker RC, Zhang W, Wang B, Hayik S, Roitberg A, Seabra G, Wong FK, Paesani
F, Wu X, Brozell S, Tsui V, Gohlke H, Yang L, Tan C, Mongan J, Hornak V, Cui G, Beroza P,
Mathews DH, Schafmeister C, Ross WS, Kollman aPA. AMBER 9. 2006
45. Merz KM. Journal of the American Chemical Society 1991;113:406.
46. Bayly CI, Cieplak P, Cornell WD, Kollman PA. Journal of Physical Chemistry 1993;97:10269.
47. Cornell WD, Cieplak P, Bayly CI, Kollman PA. Journal of the American Chemical Society
1993;115:9620.
48. Cieplak P, Cornell WD, Bayly C, Kollman PA. Journal of Computational Chemistry 1995;16:1357.
49. Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. Journal of Chemical Physics
1983;79:926.
50. Wang JM, Cieplak P, Kollman PA. Journal of Computational Chemistry 2000;21:1049.
51. Lwin TZ, Luo R. Protein Science 2006;15:2642. [PubMed: 17075138]
52. Darden T, York D, Pedersen L. Journal of Chemical Physics 1993;98:10089.
53. Ryckaert JP, Ciccotti G, Berendsen HJC. Journal of Computational Physics 1977;23:327.
54. Berendsen HJC, Postma JPM, Vangunsteren WF, Dinola A, Haak JR. Journal of Chemical Physics
1984;81:3684.
55. Ryde U. Proteins-Structure Function and Genetics 1995;21:40.
56. Nakamoto K, Takemoto J, Chow TL. Applied Spectroscopy 1971;25:352.
57. Sacconi L, Gans P, Sabatini A. Inorganic Chemistry 1964;3:1772.
58. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds. New Youk:
1978.
59. Bullock AN, Henckel J, Fersht AR. Oncogene 2000;19:1245. [PubMed: 10713666]
60. Joerger AC, Ang HC, Fersht AR. Proceedings of the National Academy of Sciences of the United
States of America 2006;103:15056. [PubMed: 17015838]
61. Joerger AC, Ang HC, Veprintsev DB, Blair CM, Fersht AR. Journal of Biological Chemistry
2005;280:16030. [PubMed: 15703170]
Lu et al. Page 9
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1.
Ribbon representation of chain B of 1TSR with the Zinc interface.
Lu et al. Page 10
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2.
Optimized Zinc interface [Zn(CYS)3(HIS)1] in B3LYP/6-311+G**.
Lu et al. Page 11
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 3.
Superposition of optimized Zinc-binding interface structure between Amber and Gaussian.
Black is for Gaussian and white is for Amber. Root mean square deviation for all heavy atoms
is 0.94 Å.
Lu et al. Page 12
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 4.
Comparison of normal mode frequencies of the Zinc interface between Gaussian and Amber.
Lu et al. Page 13
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 5.
Cumulative average potential energies for DNA-free and DNA-bound p53c. The moving
window is 1ns.
Lu et al. Page 14
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 6.
(a) The black and grey lines are backbone RMSD of all the residues of DNA-bound and DNA-
free simulations respectively. (b) The black and grey lines are RMSD of the five atoms (ZN,
3S, N) at the Zinc-binding interface.
Lu et al. Page 15
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 7.
Superposition of mean simulated structures and 1TSR B chain. Right: DNA-free p53c; left:
DNA-bound p53c. Gray: crystal; black: simulation.
Lu et al. Page 16
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 8.
B-factors for DNA-bound p53c between simulation and crystal structure (chain B). P: for
crystal packing position. The four vertical lines mark the four residues at the Zinc-binding
interface.
Lu et al. Page 17
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 9.
B-factors for DNA-free p53c between simulation and crystal structure (chain A).
Lu et al. Page 18
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 10.
Diagrams of secondary structure interaction networks. Upper: DNA-free; lower: DNA-bound.
Lu et al. Page 19
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 11.
(a) B-factors between DNA-free and DNA-bound simulations. (b) B-factors in crystal structure
(chains A and B).
Lu et al. Page 20
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 21
Table 1
Possible total charges (a. u.) and their corresponding total energies (Hartree).
Total charge +2 +1 0 12
HIS 0 0 0 0 1
CYS 0 0 0 11
CYS 0 0 111
CYS 0 1111
Total energy 3518.046193 3517.767790 3517.386853 3516.865598 3516.231373
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 22
Table 2
Information of the p53 and p53 complex MD simulations.
With DNA Without DNA
Total Atoms 43058 24993
Protein & DNA Residues 218 194
Protein & DNA Atoms 3769 3007
Water Molecules 13329 7328
Counter Ions 22 Na+ 2 Cl
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 23
Table 3
Empirically optimized force constants for ZN related bonds and angles.
This work Literature
Kb/Kθ, beq/θeq Kb/Kθ, beq/θeq
ZN-SG 60.0, 2.360 81.0, 2.29055
ZN-ND1 26.0, 2.270 40.0, 2.10056–58
SG-ZN-SG 32.0, 116.6 50.0, 146.555
SG-ZN-ND1 36.0, 100.8 30.4, 103.055
ZN-SG-CB 30.0, 103.5 19.0, 111.655
ZN-ND1-CE1 30.0, 113.0 20.0, 111.645
ZN-ND1-CG 30.0, 121.7 20.0, 111.645
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 24
Table 4
Ratios of B-factors between side chain and main chain heavy atoms at the Zn interface.
Residues Simulation Crystal
176 0.947 0.949
179 0.994 0.944
238 1.355 1.231
242 0.888 0.940
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 25
Table 5
Percentage occupancies of native contacts outside the β-sandwich region. The table only lists percentage occupancies that are higher than
50%. w DNA: DNA-bound simulation; w/o DNA: DNA-free simulation; IP: ion pair; HP: hydrophobic pair.
Secondary Structures Residues w/o DNA w DNA Type
H2 S2 E286 S127 97.1 98.0 IP
H2 S2E285 K132 97.1 100.0 IP
H2 S2 R282 Y126 99.0 98.5 IP
H2 S2 R282 T125 100.0 93.5 IP
H2 S10 D281 R273 99.5 100.0 IP
L3 L2 R249 E171 100.0 99.5 IP
L3 S4 R249 Y163 99.2 95.1 IP
L3 S5 M246 L194 74.9 99.5 HP
L3 S9 M246 I251 100.0 100.0 HP
L3 ZN G245 C242 97.5 92.5 HB
L3 S10 S240 V274 99.5 97.5 HB
L3 S3 M237 A138 99.6 100.0 HP
L2S8 L194 Y236 74.1 0.0 HP
L2S6 P190 Y205 100.0 98.4 HP
L2S6 A189 Y205 100.0 100.0 HP
L2S6 L188 Y201 71.3 9.7 HP
L2L5 L188 V203 53.0 92.0 HP
L2S6 L188 Y205 99.2 100.0 HP
L2S5 D186 R196 88.7 1.0 IP
L2S5 D184 R196 91.5 1.0 IP
H1 S5 S183 R196 0.0 77.3 IP
L2 L2R175 S183 0.0 87.9 IP
L2 L2R175 D184 100.0 0.0 IP
L2 H1 R174 E180 100.0 1.0 IP
L2 S4 V173 A161 99.2 65.2 HP
L2 S4 V173 Y163 100.0 100.0 HP
L2 S5 V173 L194 100.0 100.0 HP
L2 S9 V173 I251 98.7 99.2 HP
L2 L3 V173 M246 91.5 83.0 HP
L2 L5T170 R213 50.0 49.5 IP
L2 L5M169 R213 98.5 1.0 IP
L2 S4 M169 I162 64.0 97.6 IP
L2 S10 K164 E271 100.0 98.4 IP
S2S10 K132 E271 100.0 99.2 IP
S2 S2Y126 N131 100.0 100.0 HB
L1 H2 T118 R283 27.0 100.0 IP
L1 S2 G117 T125 100.0 79.0 HB
L1 S2 S116 C124 100.0 42.5 HB
NT L5V97 M169 91.5 43.0 HP
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 26
Table 6
Same as Table 5, but for native contacts inside the β-sandwich region.
Secondary Structure Residues w/o DNA w DNA Type
S3 S1 V143 L111 78.0 89.5 HP
S3 S8 V143 Y234 94.5 68.5 HP
S3 S9 V143 I255 66.5 4.0 HP
S3 S10 V143 F270 45.0 93.5 HP
S3 S7 V147 Y220 100.0 98.5 HP
S4 S3 R156 E258 100.0 100.0 IP
S4 S3 V157 L145 99.5 98.0 HP
S4 S9 R158 E258 100.0 100.0 IP
S5 S4 I195 A161 95.0 95.0 HP
S5 S8 I195 Y236 57.0 47.5 HP
S5 S8 I195 Y234 94.0 99.0 HP
S5 S8 E198 N235 0.0 99.6 HB
S3 S8 T140 N235 53.4 0.0 HB
S5 S3 E198 T140 97.6 100.0 IP
S5 S8 E198 H233 84.2 0.0 IP
L5 L6 R202 E221 91.9 92.7 IP
S6 L5 Y205 V203 98.0 100.0 HP
S6 S7 Y205 V216 100.0 99.0 HP
S9 S10 I251 V272 93.5 65.5 HP
S9 S1 I255 F109 94.5 100.0 HP
S9 S8 I255 Y234 100.0 100.0 HP
S9 L4 L257 P151 99.5 99.0 HP
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lu et al. Page 27
Table 7
Same as Table 5, but for native contacts with DNA.
Secondary Structures Residues w DNA Type
L1 DNA K120 G7 96.8 IP
L1 DNA K120 G8 100.0 IP
L3 DNA N239 G13 100.0 HB
L3 DNA S241 T12 93.5 HB
L3 DNA S241 G13 89.1 HB
L3 DNA R248 T11 100.0 IP
L3 DNA R248 T12 81.4 IP
L3 DNA R248 T26 100.0 IP
L3 DNA R248 G25 65.2 IP
S10 DNA R273 T12 100.0 IP
H2 DNA R280 G13 100.0 IP
H2 DNA R283 G19 100.0 IP
J Phys Chem B. Author manuscript; available in PMC 2008 October 4.
... To fulfill the physiological pH condition (i.e., pH 7.4 ) [41] in each full-length p53 form, Asp and Glu residues were deprotonated, Arg and Lys residues were protonated, and His residues kept neutral. To model the coordinated Zn 2+ complex state [42], three Cys residues were deprotonated (i.e., Cys176, Cys238, and Cys242 were assigned a −1 charge). This zinc model, with the protonation state of −1 , was found to be the only model that kept the zinc interface structure intact (i.e., maintaining the tetrahedral structure) [42]. ...
... To model the coordinated Zn 2+ complex state [42], three Cys residues were deprotonated (i.e., Cys176, Cys238, and Cys242 were assigned a −1 charge). This zinc model, with the protonation state of −1 , was found to be the only model that kept the zinc interface structure intact (i.e., maintaining the tetrahedral structure) [42]. The detailed atomic charge calculations for the amino acids residues in the zinc model, using B3LYP density functional method with the 6-311+G** basis set, were reported [42]. ...
... This zinc model, with the protonation state of −1 , was found to be the only model that kept the zinc interface structure intact (i.e., maintaining the tetrahedral structure) [42]. The detailed atomic charge calculations for the amino acids residues in the zinc model, using B3LYP density functional method with the 6-311+G** basis set, were reported [42]. Both N-and C-termini of each full-length p53 form were capped with acetyl (ACE) and N-methyl (NME) groups, respectively to keep them neutral. ...
Article
Full-text available
p53, a tumor suppressor protein, is essential for preventing cancer development. Enhancing our understanding of the human p53 function and its modifications in carcinogenesis will aid in developing more highly effective strategies for cancer prevention and treatment. In this study, we have modeled five human p53 forms, namely, inactive, distal-active, proximal-active, distal-Arg175His mutant, and proximal-Arg175His mutant forms. These forms have been investigated using Gaussian accelerated molecular dynamics (GaMD) simulations in OPC water model at physiological temperature and pH. Our observations, obtained throughout [Formula: see text] of production run, are in good agreement with the relevant results in the classical molecular dynamics (MD) studies. Therefore, GaMD method is more economic and efficient method than the classical MD method for studying biomolecular systems. The featured dynamics of the five human p53-DBD forms include noticeable conformational changes of L1 and [Formula: see text]-[Formula: see text] loops as well as [Formula: see text]-[Formula: see text] and [Formula: see text]-[Formula: see text] turns. We have identified two clusters that represent two distinct conformational states in each p53-DBD form. The free-energy profiles of these clusters demonstrate the flexibility of the protein to undergo a conformational transition between the two clusters. We have predicted two out of seven possible druggability pockets on the clusters of the Arg175His forms. These two druggability pockets are near the mutation site and are expected to be actual pockets, which will be helpful for the compound clinical progression studies.
... The copyright holder for this preprint this version posted November 10, 2023. ; https://doi.org/10.1101/2023.11.09.566481 doi: bioRxiv preprint Zn developed by Luo's group (57); parameters for NAD + developed by Walker et al (58) and Pavelites et al (59) ; parameters for acetylated lysine developed by Papamokos et al (60). Amber14 tools (Antechamber program) are employed for parametrizing the non-standard residues and tleap program for creating topology and force field parameters for the standard residues. ...
Preprint
Full-text available
Mammalian sirtuins (SIRT1-SIRT7) are a family of nicotinamide adenine dinucleotide (NAD ⁺ )-dependent protein deacylases that play critical roles in lifespan and age-related diseases. The physiological importance of sirtuins has stimulated intense interest in designing sirtuin activating compounds. However, except for allosteric activators of SIRT1-catalyzed reactions that are limited to specific substrates, methodologies for the rational design of sirtuin activating compounds -- including compounds that activate mitochondrial sirtuins implicated in the age-related decline of cellular metabolism -- have been lacking. Here, we use computational high-throughput screening methodologies and a biophysical model for activation of the major mitochondrial sirtuin SIRT3 to identify novel small molecule activators of the human SIRT3 enzyme from a 1.2 million compound library. Unlike previously reported SIRT3 activators like Honokiol, which only transiently upregulate SIRT3 under non-steady state conditions and reduce the steady state catalytic efficiency of the enzyme, several of the novel compounds identified here are potent SIRT3 activators in both the steady and non-steady states. Two such compounds can almost double the catalytic efficiency of the enzyme with respect to NAD ⁺ , which would be sufficient to almost entirely compensate for the loss in SIRT3 activity that occurs due to the reduction in mitochondrial coenzyme concentration associated with aging, and display AC50s (concentrations of half-maximal activation) as low as 100 nM. The current work thus reports first-in-class, non-allosteric steady state activators that activate SIRT3 through a novel, mechanism-based mode of activation and that may be developed further for therapeutic applications.
... Moreover, we thought that the combination of simulation approaches with the current information regarding the role of this TP53 variant in apoptosis induction, suppression of cell growth and transcriptional activation might help in solving the discrepancies associated with this infamous polymorphism (Volodko et al., 2015). Consequently, for the first time, we generated the full-length p53 protein model, as it has already been reported that the oncogenic form of this protein typically consists of the full-length protein with a single amino-acid substitution in its DNA-binding domain (Lu et al., 2007). On that account, the total energy of the mutant Pro-72 model showed higher energy than the wild-type Arg-72 model, suggesting its lesser stability (Rajasekaran & Sethumadhavan, 2010), which was also supported by the backbone as well as C-alpha RMSD differences, where the mean RMSD was increased in the mutant model speculating destabilization of the structure (Figure 2(A,B)). ...
Article
Head and neck cancer (HNC), the sixth most common cancer globally, stands first in India, especially Northeast India, where tobacco usage is predominant, which introduces various carcinogens leading to malignancies by accumulating DNA damages. Consequently, the present work aimed to predict the impact of significant germline variants in DNA repair and Tumour Suppressor genes on HNC development. WES in Ion ProtonTM platform on ‘discovery set’ (n = 15), followed by recurrence assessment of the observed variants on ‘confirmation set’ (n = 40) using Sanger Sequencing was performed on the HNC-prevalent NE Indian populations. Initially, 53 variants were identified, of which seven HNC-linked DNA damage response gene variants were frequent in the studied populations. Different tools ascertained the biological consequences of these variants, of which the non-coding variants viz. EXO1_rs4150018, RAD52_rs6413436, CHD5_rs2746066, HACE1_rs6918700 showed risk, while FLT3_rs2491227 and BMPR1A_rs7074064 conferred protection against HNC by affecting transcriptional regulation and splicing mechanism. Molecular Dynamics Simulation of the full-length p53 model predicted that the observed coding TP53_rs1042522 variant conferred HNC-risk by altering the structural dynamics of the protein, which displayed difficulty in the transition between active and inactive conformations due to high-energy barrier. Subsequent pathway and gene ontology analysis revealed that EXO1, RAD52 and TP53 variants affected the Double-Strand Break Repair pathway, whereas CHD5 and HACE1 variants inactivated DNA repair cascade, facilitating uncontrolled cell proliferation, impaired apoptosis and malignant transformation. Conversely, FLT3 and BMPR1A variants protected against HNC by controlling tumorigenesis, which requires experimental validation. These findings may serve as prognostic markers for developing preventive measures against HNC.
Article
p53 is a tumor suppressor protein that plays a crucial role in inhibiting cancer development and maintaining the genetic integrity. Within the cell nucleus, four p53 molecules form a stable tetrameric active structure through highly cooperative interactions, bind to DNA via its DNA-binding domain, and transcriptionally activate or inhibit their target genes. However, in most human tumor cells, there are numerous p53 mutations. The majority of these mutations are formed in the p53 DNA-binding domain, importantly, the p53 DNA-binding domain is critical for p53 to form the tetrameric active structures and to regulate the transcription of its downstream target genes. This study exploited the all-atom molecular dynamics simulations to investigate the mechanisms of interactions within the wild-type p53 tetramers. This study indicates that the symmetric dimers on either side of the DNA are stable ones, keep stable structures both before and after DNA bindings. The binding of two monomers on the same side of the DNA depends on proteinprotein interactions provided by two contact surfaces. DNA scaffold stabilizes the tetrameric active structure. Such interactions play a crucial role in helping the tetramer formation. This study clarifies the internal interactions and key residues within the p53 tetramer during dynamic processes, as well as the critical sites at various interaction interfaces. The findings of this study may provide a significant foundation for us to further understand the p53's anticancer mechanisms, to explore the effective cancer treatment strategies, and in near future, to develop the effective anti-cancer drugs.
Article
In epithelial tumors, oncoprotein E6 binds with the ubiquitin ligase E6AP to form E6/E6AP heterodimer, then this heterodimer recruits p53 to form E6/E6AP/p53 heterotrimer and induces p53 degradation. Recent experiments demonstrated that three E6 single-site mutants (F47R, R102A and L50E) can inhibit the E6/E6AP/p53 heterotrimer formation and rescue p53 from the degradation pathway. However, the molecular mechanism underlying mutation-induced heterotrimer inhibition remains largely elusive. Herein, we performed extensive molecular dynamics simulations (totally ∼13 μs) on both heterodimer and heterotrimer to elucidate at an atomic level how each p53-degradation-defective HPV16 E6 mutant reduces the structural stabilities of the two complexes. Our simulations reveal that the three E6 mutations destabilize the structure of E6/E6AP/p53 complex through distinct mechanisms. Although F47RE6 mutation has no effect on the structure of E6/E6AP heterodimer, it results in an electrostatic repulsion between R47E6 and R290p53, which is unfavorable for E6-p53 binding. R102AE6 mutation destabilizes the structure of E6/E6AP heterodimer and significantly disrupts hydrophobic and cation-π interactions between F47E6 and E286p53/L298p53/R290p53. L50EE6 mutation impairs both E6 inter-domain interactions (especially F47-K108 cation-π interaction) and E6-E6AP intermolecular interactions important for the stabilization of E6/E6AP heterodimer. This study identifies the intra- and inter-molecular interactions crucial for the complex stability, which may provide mechanistic insights into the inhibition of complex formation by the three HPV16 E6 mutations.
Article
Full-text available
Human papillomavirus (HPV) is a common virus, and about 5% of all cancers worldwide is caused by persistent high-risk HPV infections. Here, we reported a comprehensive analysis of the molecular features for HPV-related cancer types using TCGA (The Cancer Genome Atlas) data with HPV status. We found that the HPV-positive cancer patients had a unique oncogenic process, tumor microenvironment, and drug response compared with HPV-negative patients. In addition, HPV improved overall survival for the four cancer types, namely, cervical squamous cell carcinoma (CESC), head and neck squamous cell carcinoma (HNSC), stomach adenocarcinoma (STAD), and uterine corpus endometrial carcinoma (UCEC). The stronger activity of cell-cycle pathways and lower driver gene mutation rates were observed in HPV-positive patients, which implied the different carcinogenic processes between HPV-positive and HPV-negative groups. The increased activities of immune cells and differences in metabolic pathways helped explain the heterogeneity of prognosis between the two groups. Furthermore, we constructed HPV prediction models for different cancers by the virus infection score (VIS) which was linearly correlated with HPV load and found that VIS was associated with drug response. Altogether, our study reveals that HPV-positive cancer patients have unique molecular characteristics which help the development of precision medicine in HPV-positive cancers.
Preprint
Full-text available
The dissociation process of the DNA binding domain of p53 (p53-DBD) from a DNA duplex that contains the consensus sequence, which is the specific target of p53-DBD, was investigated by a combination of dissociation parallel cascade selection molecular dynamics (dPaCS-MD) and the Markov state model (MSM). Based on an all-atom model including explicit solvent, we first simulated the p53-DBD dissociation processes by 75 trials of dPaCS-MD, which required an average simulation time of 11.2 ± 2.2 ns per trial. By setting the axis of the DNA duplex as the Z-axis and the binding side of p53-DBD on DNA as the + side of the X-axis, we found that dissociations took place along the +X and −Y directions (−Y directions) in 93% of the cases, while 7% of the cases moved along +X and +Y directions (+Y directions). Toward the −Y directions, p53-DBD dissociated first from the major groove and then detached from the minor groove, while unbinding from the minor groove occurred first in dissociations along the +Y directions. Analysis of the free energy landscape by MSM showed that loss of the minor groove interaction with p53-DBD toward the +Y directions incurred a relatively high energy cost (1.1 kcal/mol) upon a critical transition, whereas major groove detachment more frequently occurred with lower free energy costs. The standard binding free energy calculated from the free energy landscape was −10.9 ± 0.4 kcal/mol, which agrees with an experimental value of –11.1 kcal/mol. These results indicate that the dPaCS-MD/MSM combination can be a powerful tool to investigate dissociation mechanisms of two large molecules. Minor groove binding is mainly stabilized by R248, identified as the most important residue that tightly binds deep inside the minor groove. Analysis of the p53 key residues for DNA binding indicates high correlations with cancer-related mutations, confirming that impairment of the interactions between p53-DBD and DNA can be frequently related to cancer.
Article
Molecular recognition is fundamental to transcription regulation. As a transcription factor, the tumor suppressor p53 has to recognize either specific DNA sequences or repressor protein partners. However, the molecular mechanism underlying the p53 conformational switch from the DNA-bound to repressor-bound states is not fully characterized. The highly charged nature of these interacting molecules prompted us to explore the non-bonded energy contributions behind molecular recognition of either a DNA or the repressor protein iASPP by p53 DNA-binding domain (p53DBD), using molecular dynamics (MD) simulation followed by rigorous analyses of energy terms. Our results illuminate the allosteric pathway by which iASPP binding to p53 diminishes binding affinity between p53 and DNA. Even though the p53DBD uses a common framework of residues for recognizing both DNA and iASPP, a comparison of the electrostatics in the two p53DBD complexes revealed significant differences in residue-wise contributions to the electrostatic energy. We found that an electrostatic allosteric communication path exists in the presence of both substrates. It consists of evolutionarily conserved residues, from residue K120 of the binding loop L1 to a distal residue R213 of p53DBD. K120 is near the DNA in the p53DBD-DNA complex, whereas iASPP binding moves it away from its DNA-binding position in the p53DBD-iASPP complex. The ‘energy-Hubs’ (the residues show a higher degree of connectivity with other residues in the electrostatic networks) determined from the electrostatic network analysis established that this conformational change in K120 completely rewires the electrostatic network from K120 to R213, thereby impeding DNA binding. Furthermore, we found shifting populations of hydrogen bonds and salt bridges reduce pair-wise electrostatic energies within p53DBD in its DNA-bound state.
Article
Full-text available
Geometry optimization has become an essential part of quantum-chemical computations, largely because of the availability of analytic first derivatives. Quasi-Newton algorithms use the gradient to update the second derivative matrix (Hessian) and frequently employ corrections to the quadratic approximation such as rational function optimization (RFO) or the trust radius model (TRM). These corrections are typically carried out via diagonalization of the Hessian, which requires O(N3) operations for N variables. Thus, they can be substantial bottlenecks in the optimization of large molecules with semiempirical, mixed quantum mechanical/molecular mechanical (QM/MM) or linearly scaling electronic structure methods. Our O(N2) approach for solving the equations for coordinate transformations in optimizations has been extended to evaluate the RFO and TRM steps efficiently in redundant internal coordinates. The regular RFO model has also been modified so that it has the correct size dependence as the molecular systems become larger. Finally, an improved Hessian update for minimizations has been constructed by combining the Broyden–Fletcher–Goldfarb–Shanno (BFGS) and (symmetric rank one) SR1 updates. Together these modifications and new methods form an optimization algorithm for large molecules that scales as O(N2) and performs similar to or better than the traditional optimization strategies used in quantum chemistry. © 1999 American Institute of Physics.
Article
An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolutions using fast Fourier transforms. Timings and accuracies are presented for three large crystalline ionic systems. The Journal of Chemical Physics is copyrighted by The American Institute of Physics.
Article
Describes and discusses the use of theoretical models as an alternative to experiment in making accurate predictions of chemical phenomena. Addresses the formulation of theoretical molecular orbital models starting from quantum mechanics, and compares them to experimental results. Draws on a series of models that have already received widespread application and are available for new applications. A new and powerful research tool for the practicing experimental chemist.
Article
The Raman spectra of [64Zn(NH3)4]I2 and its 68Zn analog have been measured in the crystalline state. The Zn-N stretching vibrations have been assigned based on the observed metal isotope shifts. These shifts have been compared with theoretical values obtained from a complete normal coordinate analysis on the [Zn(NH3)4]2+ ion.
Article
We present a new approach to generating electrostatic potential (ESP) derived charges for molecules. The major strength of electrostatic potential derived charges is that they optimally reproduce the intermolecular interaction properties of molecules with a simple two-body additive potential, provided, of course, that a suitably accurate level of quantum mechanical calculation is used to derive the ESP around the molecule. Previously, the major weaknesses of these charges have been that they were not easily transferable between common functional groups in related molecules, they have often been conformationally dependent, and the large charges that frequently occur can be problematic for simulating intramolecular interactions. Introducing restraints in the form of a penalty function into the fitting process considerably reduces the above problems, with only a minor decrease in the quality of the fit to the quantum mechanical ESP. Several other refinements in addition to the restrained electrostatic potential (RESP) fit yield a general and algorithmic charge fitting procedure for generating atom-centered point charges. This approach can thus be recommended for general use in molecular mechanics, molecular dynamics, and free energy calculations for any organic or bioorganic system.
Article
In order to study the coordinative behavior of doubly charged metal ions in water, a few representative metals have been chosen for theoretical studies. These are the group 2 metal ions beryllium, magnesium, and calcium and the group 12 zinc ion. The density functional method B3LYP has been used with very large basis sets. It is found that the water dipole moment and polarizabilities, which are critical for the accuracy of the binding energies, are very well reproduced provided that the basis set on the metal is included in the calculations. One of the main points of the present investigation has been to study the boundary between the first and second hydration shells. Trends of binding energies and structures are also discussed.
Article
Free energy perturbation calculations of the catalytic effects associated with substitutions of the active site Ca2+ ion in staphylococcal nuclease are reported. The calculated changes in the activation barrier for different ions are found to be consistent with kinetic measurements, and the catalytic rate of enzyme indeed appears to be optimized for Ca2+. Our results indicate that the more electrophilic ions (with large hydration free energy) increase the activation barrier as a result of overstabilization of the intermediately created OH- nucleophile and that the enzymatic rate is more affected by these ions than by those that are less electrophilic than Ca2+. A simple model for treating transition-metal ions is also presented and calibrated for the Mn2+ ion in solution. The calculated decrease in activity when Mn2+ is bound to the enzyme agrees fairly well with experimental observations. Simple free energy relationships are outlined in order to classify different types of metal-catalyzed enzymatic reactions. These relationships demonstrate that the optimization of the catalytic efficiency for a particular ion is related to its multiple tasks during the reaction; i.e., the ion must stabilize the negatively charged nuclephile as well as the subsequent transition state. Several other metalloenzymes are discussed in these terms, and it is argued that such free energy relationships can provide qualitative predictions of the effects associated with metal substitutions. Finally, a tentative qualitative classification of metalloenzymes is presented in terms of the interplay between metal and general-base catalysis, again based on linear free energy concepts.