Jumper Uchicago 0330D 13647
Jumper Uchicago 0330D 13647
Jumper Uchicago 0330D 13647
A DISSERTATION SUBMITTED TO
THE FACULTY OF THE DIVISION OF THE PHYSICAL SCIENCES
IN CANDIDACY FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY
DEPARTMENT OF CHEMISTRY
BY
JOHN M. JUMPER
CHICAGO, ILLINOIS
MARCH 2017
Copyright c 2017 by John M. Jumper
All Rights Reserved
To Sarah, Everett, Carolyn, and my grandfather Fred House
“All models are wrong but some are useful” – George Box
TABLE OF CONTENTS
LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Computational trends that favor statistical potentials . . . . . . . . . . . . . 2
1.2 Contributions of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3 CONTRASTIVE DIVERGENCE . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2 Coarse-grained model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.3 Contrastive divergence method . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4 Handling crystallographic artifacts . . . . . . . . . . . . . . . . . . . . . . . 49
3.5 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.6 Accuracy of de novo Folding . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.7 Structural characterization of low temperature conformations . . . . . . . . . 55
3.8 Melting behavior and unfolded states . . . . . . . . . . . . . . . . . . . . . . 57
3.9 Accuracy of de novo Folding . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.10 Related Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.11 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.12 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
v
3.13 Training data and optimization . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.14 Details of test proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.15 Clustering of protein structures . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.16 Derivation of contrastive divergence . . . . . . . . . . . . . . . . . . . . . . . 70
3.17 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5 FUTURE DIRECTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.1 Future improvements to the Upside model and training . . . . . . . . . . . . 99
5.1.1 Enhanced equilibration for simulations and training . . . . . . . . . . 99
5.1.2 Improved treatment of hydrogen bonding . . . . . . . . . . . . . . . . 100
5.1.3 Understanding the important physics of protein folding . . . . . . . . 100
5.2 Applications of the Upside model . . . . . . . . . . . . . . . . . . . . . . . . 101
5.2.1 Prediction of native state fluctuations and hydrogen exchange . . . . 101
5.2.2 Prediction of protein-protein binding . . . . . . . . . . . . . . . . . . 101
5.2.3 Co-prediction of structure for homologous sequences . . . . . . . . . . 102
5.2.4 Computational simulation of mutational scans . . . . . . . . . . . . . 103
5.2.5 Prediction of conformational change . . . . . . . . . . . . . . . . . . . 104
5.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
vi
LIST OF FIGURES
2.1 Inner loop of Upside calculation. Executing this computational loop in the present
simulations takes between 0.25 CPU milliseconds (BBA) and 0.82 CPU millisec-
onds (ubiquitin) per iteration (each simulation replica is run on single 2.60 GHz
processor core). The side chain potential enters into the integration step simply
as a complicated, many-body energy function that may be treated with standard
techniques of molecular simulations. . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Fragment of protein G with associated interaction graph (Rcutoff = 7Å). A pair
of residues has a connection whenever their side chain beads are within Rcutoff
for any side chain states. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Example of optimized coarse states for arginine overlaid on the PDB distribution
of the rotamer angles χ1 and χ2 . Each of the six coarse states contains only a
single fine state that has high probability, so that the variance of dihedral angles
within each coarse state is small. . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Error in the decomposition of rotamer states into coarse-grained states as a func-
tion of the number of side chain states. The position uncertainty is σ. The relative
uncertainty is the position uncertainty for each number of states divided by the
accuracy at 3 states. For residues without a rotable χ2 , such as valine, it is not
possible to coarse-grain. One, three or six rotamer states are used, depending on
the residue type. The computational time to compute the pairwise interactions
and solve for the free energy scales roughly as the number of coarse rotamer states
squared, so there is an incentive to use as few coarse states as possible. The table
(above) summarizes the number of states chosen for each amino acid type. . . . 20
2.5 Left panel depicts side chain interaction coordinates for equation (2.24). Right
panel depicts side chain interactions with backbone hydrogen (blue) and oxygen
(red) sites; axes labels indicated on lower right plot. Within each group of 4
plots, the curves on the left hand side present the radial components of the
interaction. The thin lines provide the unif(r) interaction and the thick line is
the directional interaction unif(r) + dir(r). The right plots are heat maps of the
angular interactions ang1 (θH/O ) ang2 (θSC ), where the lower left corner depicts the
interaction where θ1 = θ2 = 0◦ . Within each group of 4 plots, the upper plots
represent side chain interactions with hydrogen and the bottom plots represent
side chain interactions with oxygen. . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6 Accuracy of the model (probability assigned to true χ̃ state) versus regularization
strength (essentially the inverse of kr ). Regularization as employed in this work
does not appear to improve the accuracy of the side chain packing significantly.
The regularization, especially the lower bound to the energy, may have a strong
effect on the accuracy of protein simulation, even if does not increase packing
accuracy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.7 Comparison of χ1 prediction accuracy for Upside and SCWRL4. The “No in-
teractions” line represents the accuracy of the only the NDRD rotamer library
without any interactions; this library is used in both Upside and SCWRL. Note
that Upside is approximately 150x faster than SCWRL at side chain prediction,
in addition to return a probability distribution instead of a single answer. . . . . 28
vii
2.8 Accuracy over short duration simulations. The backbone hydrogen bonding
strength is not determined by the packing optimization, so we search for the
strength that gives the best simulation accuracy. This is the only parameter in
the model directly optimized for simulation accuracy. To assess accuracy, we look
at the fraction of each simulation with RMSD to native of less than 5 Å, and we
compute the median accuracy over all proteins in the test set. All other results
shown are for a backbone hydrogen bond energy of −1.8 kT . . . . . . . . . . . . 31
2.9 RMSD to native over replica exchange simulation trajectories with hydrogen bond
energy −1.8 kT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
viii
4.1 Probability density of Ramachandran angles for the central residue of a free ala-
nine chain with no Ramachandran potential for the central residue (colors scale
gives the probability density values in units of 1/degrees2 ). The chain experience
only short-range excluded volume interactions, and interactions between adjacent
residues are excluded from the potential. The depression near φ = ψ = 0 occurs
due to a high probability of i − 1, i + 1 steric overlap when the i-th residue is near
the origin of φ/ψ space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
ix
LIST OF TABLES
3.1 Lowest RMSD for de novo folding simulations. . . . . . . . . . . . . . . . . . . . 54
x
ACKNOWLEDGMENTS
This thesis would be impossible without my family. My wife Carolyn is unbelievably wise and
supportive. She is my best counsel and greatest strength. Sarah and Everett are absolutely
wonderful, and they gently remind me of the important of work-life balance. When I spend
too much time on research, Sarah closes my laptop and reminds me, “Daddy, you can do
that later.” I also want to thank my parents for putting my education first when I was
young, giving me a great foundation on which I am still building.
I owe a great intellectual debt to my advisors, Tobin Sosnick and Karl Freed. It is
essential for theoretical work to have a close connection to experiment evidence, and Tobin
is excellent at interpreting between these two cultures of science. He further provided me
with enormous latitude to address protein modelling using an idiosyncratic mix of physics,
statistics, and computer science techniques, even when such approaches little resemble to
familiar techniques. Karl is a consummate theorist and is an inspiration to a lifetime of
tackling hard problems with insightful techniques. I also want to thank my lab mates for
vigorous discussion and a willingness to argue forcefully in and out of group meetings.
Finally, I want to thank my friends in physical chemistry for making Hyde Park a great
place to spend my time.
xi
CHAPTER 1
INTRODUCTION
Proteins are challenging physical systems to model because of an interplay between statistical
mechanics that explores conformational space compatible with a protein’s sequence and
evolution that explores sequence space compatible with a protein’s structure and function.
While some aspects of protein structure are amenable to deduction from simple principles,
such as Linus Paulings’ prediction of helical and sheet structure, much of our knowledge of
protein physics comes from a great number of generally weak energetic preferences rather
than hard rules, requiring a more flexible approach to understanding structure and dynamics.
Since the pioneering work of Karplus on molecular dynamics, scientists have attempted to
characterize the wiggling and jiggling of protein atoms by watching them do so on their
computer in order to transfer computational observations to experimental hypotheses. This
thesis will explore new, rigorous methods to characterize both the motions of proteins and
the parameters that control them.
Sampling a conformation from the Boltzmann ensemble of proteins is extremely chal-
lenging owing to three separate factors. The first challenge is to represent the complicated
potential energy surface of the protein. The cooperative folding behavior of proteins[2] in-
dicates that proteins are often finely balanced between widely-separated states, the native
state and the unfolded state. We must represent the energetics of these states carefully in
order to reproduce this behavior. The tremendous number of potential conformations re-
quires us to separate the native state from all other conformations by a large free energy
gap in order to observe the high fraction of the native state observed in the experimental
ensemble.
The second challenge is that a realistic protein potential energy will be subject to the
separation of time scales[7] that typifies protein dynamics. It is plausible that one can
create a unrealistic potential energy function that folds nearly downhill to the native state
for even large proteins with great accuracy; this often is the basis of structure prediction
1
methodologies that make no attempt to represent the Boltzmann ensemble of a protein. For
models that seek to represent the totality of protein physics, we likely must accept that
folding is often cooperative, with the attendent extended conformational search.
This thesis and the Upside model described herein tackles the potential and sampling
challenges simultaneously. By carefully designed to be physical, this procedure should mostly
extract physical information from statistical regularities (deviations from randomness) of
the training data. In some cases, however, the potential energy will lack the necessary
flexibility in its functional form to model the underlying physics. In these cases, unphysical
interaction energies of the terms present in the potential may be required to best match
experimental data. While such unphysical interaction energies provide the best prediction of
training data by definition, they may limit the ability of the resulting models to generalize to
related prediction (e.g. while a potential energy may provide the best native-state structure
predictions, it may not generalized to correctly predict the properties of extended states).
A similar situation occurs in other potential functions. Three-molecule interactions are
very important to the energetics of water clusters and water interfaces, yet are absent in many
common water models like TIP3P[3]. To obtain the impressive agreement of water models
to bulk water observables[5], three-molecule interactions must be incorporated into two-
molecule interactions in an averaged sense. This approximation works for many properties
of interest but can give flawed predictions at interfaces. The general solution is to identify
the deficiencies and propose modified potentials to reduce the impact of these errors. The
solution for protein statistical potentials would be similar.
3
use easily. The trends in computational resources make this approach increasingly viable, so
long as we may equilibrate of single proteins on a small number of computational cores in
reasonable time. Our newly-created Upside software and parameterized potential described
in this thesis are expressly intended to enable sufficiently-fast PDB-scale training on current
computational resources.
The resulting Upside model of protein physics enables a great variety of protein studies,
from folding to conformational change to binding to predictions of protein structure using
Upside as a strong Bayesian prior on protein structures. The focus of this thesis has been
to develop techniques that allow researchers to easily obtain a sample from the Boltzmann
distribution for proteins of small-to-moderate size. Most of the dividends of this research
program are still to be reaped and a selection of potential applications are discussed in
chapter 5.
This thesis makes contributions to both the sampling and parameterization problems for
coarse-grained protein potentials.
In chapter 2, we solve a major limitation of protein simulations, the inability to have
detailed side chain interactions without the difficult equilibration of very rough energy sur-
face. We define and approximate a novel scheme for instantaneously-equilibrated side chain
free ensembles. This side chain free energy allows rapid sampling in the smoother potential
energy surface of the protein backbone conformations for molecular dynamics while retaining
the detailed energetics of the side chain conformations. Importantly, the free energy is inher-
ently many-body, as is typical for interacting systems with pair interactions, capturing the
non-additivity of side chain interactions due to the inability to simultaneously adopt multiple
rotamer conformations simultaneously. As an associated benefit, direct optimization of the
rotamer prediction accuracy of the side chain model yields a side chain rotamer predictor
competitive to the state of the art in accuracy while running two orders of magnitude faster.
4
In chapter 3, we address the parameterization challenge by using the contrastive diver-
gence method from machine learning to parameterize a detailed potential for protein folding.
We show that only optimizing near-crystal conformations is sufficient to define a potential
capable of de novo folding of small proteins to high accuracy. Furthermore, we examine the
ability of our model to represent temperature-denatured states of proteins in preparation for
future work on folding pathways. It should be noted that chapters 2 and 3 are intended for
publication, so that they contain redundancies with other parts of the thesis.
Chapter 4 describes the details of the Upside model. In particular, the subtleties of
working with an existing torsional potential for the Ramachandran angles is discussed, as
well as the need for careful treatment of side chain-backbone interactions. We also discuss a
number of decisions in the design of a coarse-grained potential necessary to ensure effective
and rapid sampling.
Integral to the theoretical work described above, we have developed an entirely new simu-
lation package for the simulation and training of coarse-grained physics models, including our
semi-implicit side chain model. This simulation packages applies lessons from the machine
learning community that it is essential to have a single framework for training and testing
statistical models so the same implementation used for protein dynamics can be used to com-
pute the parameter derivatives needed to train the model. This simulation package is also
extensible so that arbitrarily sophisticated potential energy functions or virtual coordinates
may be created, which is intended to be an extensible platform to integrate sophisticated
new sources of information, such as that available from statistical predictions of secondary
or tertiary structure. This software is publically available under an open-source license at
https://psd-repo.uchicago.edu/freed-and-sosnick-lab/upside-md.
Finally, chapter 5 concludes with future work that builds on the methods and perspec-
tives developed in this work. Both future applications to protein dynamics and enhanced
parameterization techniques are presented.
5
1.3 References
[1] John D Chodera and Frank Noé. Markov state models of biomolecular conformational
dynamics. Current opinion in structural biology, 25:135–144, 2014.
[2] Hüseyin Kaya and Hue Sun Chan. Polymer principles of protein calorimetric two-state
cooperativity. Proteins: Structure, Function, and Bioinformatics, 40(4):637–661, 2000.
[3] R Kumar and JL Skinner. Water simulation model with explicit three-molecule interac-
tions. The Journal of Physical Chemistry B, 112(28):8311–8318, 2008.
[4] Stefan M Larson, Christopher D Snow, Michael Shirts, and Vijay S Pande. Folding@home
and genome@home: Using distributed computing to tackle previously intractable prob-
lems in computational biology. arXiv preprint arXiv:0901.0866, 2009.
[5] Pekka Mark and Lennart Nilsson. Structure and dynamics of the tip3p, spc, and spc/e
water models at 298 k. The Journal of Physical Chemistry A, 105(43):9954–9960, 2001.
[6] David E Shaw, JP Grossman, Joseph A Bank, Brannon Batson, J Adam Butts, Jack C
Chao, Martin M Deneroff, Ron O Dror, Amos Even, Christopher H Fenton, et al. Anton
2: raising the bar for performance and programmability in a special-purpose molecu-
lar dynamics supercomputer. In Proceedings of the International Conference for High
Performance Computing, Networking, Storage and Analysis, pages 41–53. IEEE Press,
2014.
[7] David E Shaw, Paul Maragakis, Kresten Lindorff-Larsen, Stefano Piana, Ron O Dror,
Michael P Eastwood, Joseph A Bank, John M Jumper, John K Salmon, Yibing Shan,
et al. Atomic-level characterization of the structural dynamics of proteins. Science,
330(6002):341–346, 2010.
6
CHAPTER 2
2.1 Introduction
Two major challenges must be overcome in order to confront the difficulties of accurately
simulating protein dynamics. The first is the necessity of balancing the large and compet-
ing sources of energy and entropy whose total determines both the thermodynamics and
the native conformation of the protein. The second challenge involves the intensive sam-
pling required to obtain a Boltzmann ensemble of conformations. The sampling challenge
is addressed here by integrating out the side chain free energy to produce a coarse-grained
configuration defined just in terms of the backbone N, Cα , and C atoms. Consequently, back-
bone motions evolve on a smoother coarse-grained free energy surface with greatly reduced
side chain rattling (molecular friction) compared to that for standard all-atom molecular
dynamics simulations (see article on long side chain timescales).
The uncertainty in the position of coarse-grain interactions introduces the difficulty of ac-
curately parameterizing a coarse-grained model to represent the physical interactions. More-
over, all-atom force fields produce conformations that deviate from experiment, especially for
unfolded proteins[20]. Hence, rather than following the customary process of matching the
energies of the coarse-grained model to approximate the already inexact energies of atomistic
force fields or trying to interpret raw statistics for the distribution of interatomic distances in
the Protein Data Bank (PDB)[1] and defining the correct reference state appropriate to the
statistical potential[7], our side chain energies are determined as those that best reproduce
the side chain conformations observed in the PDB given fixed backbone configurations. This
maximum likelihood approach has key advantages: (1) it directly provides an interpretation
of the structural information as a sample from the statistical mechanical ensemble of side
7
chain packing, and (2) it can be evaluated quickly since we will show that approximating
the Boltzmann distribution for the side chains in a fixed backbone configuration does not
require laborious discrete sampling of the χ angles. This work also presents a computation-
ally extremely inexpensive, coarse-grained approximation for describing side chain packing,
thereby allowing the model to be directly used as the free energy function for molecular
dynamics simulations. Our method enables rapidly equilibrating coarse-grained simulation
that can nonetheless contain significant molecular detail.
The positions of the N, Cα , and C atoms constitute the backbone trace. The native backbone
trace determines the fold of the protein, and the free energy of the trace is likely to preserve
the major barriers that determine the slow degrees of freedom in the protein. The strategy
in our method, called Upside, is to perform dynamics simulations of the backbone trace,
while still including sufficient structural details (side chain structures and free energies, etc.)
necessary to compute realistic forces on the three atoms of the backbone trace. This strategy
yields the advantage that the inclusion of the side chain free energy, rather than the side
chains themselves, greatly smooths the potential governing the dynamics of the backbone
trace, especially because of the reduction of steric rattling attributable to the side chains as
they try to sample multiple substates in the condensed state.
First consider a representation of the protein configurations in terms of the coordinates
({bi }, {χi }) where bi represents the positions of the backbone N, Cα , and C atoms on the
i-residue and χi represents the side chain χ-angles on the i-th residue. Since bond lengths
and angles are approximately constant for proteins, the positions of the protein atoms can
be reconstructed with high accuracy from the ({bi }, {χi }) coordinates. Given a potential
8
Add*atoms*to*AA* 3*or*6*side*chain*
3*atoms/AA* posi:ons*@*each*χ1*
C* Cα χ1 1*bead*/*side*chain*
Cα N *
N
C*
Authen:c*H<bonds**
&*(φ,ψ)*angles
Move*C,*N*&*
Cα:*F=ma"
(Langevin*
dynamics)*
Computa:onal*loop* Op:mize*P(χ1)*to*
pack*the*core
Cα N
C* Un
(pull*b decorate*
ack*f
plane orces*onto
*(C,*N, * (χ 1)**
*Cα )* Prob ed**
ht
weig :ons*
C* ac
Cα inter
N
Figure 2.1: Inner loop of Upside calculation. Executing this computational loop in the
present simulations takes between 0.25 CPU milliseconds (BBA) and 0.82 CPU milliseconds
(ubiquitin) per iteration (each simulation replica is run on single 2.60 GHz processor core).
The side chain potential enters into the integration step simply as a complicated, many-body
energy function that may be treated with standard techniques of molecular simulations.
9
energy V ({bi }, {χi }), we define the free energy as a function of the backbone configuration,
Z
V̄ ({bi }) = − log dχ1 · · · χN e−V ({bi },{χi }) . (2.1)
Natural energy units are used so that kB T = 1. An intermediate step of this derivation
requires the introduction of a discrete approximation {χ̃i } for our χ-angles and a discrete
approximation V̄ ({bi }, {χ̃i }) for the potential.
Rather than directly calculate (2.1), we define an intermediate discrete approximation
to V̄ that is amenable to approximation techniques. Consider a discrete coarse-graining
function g so that χ̃i = g(χi ), where χ̃i is a small integer (χ̃i ∈ {1, . . . , 6} in this work). The
coarse-grain potential Ṽ is defined so that
Z !
e−Ṽ ({bi },{χ̃i }) ≈ e−V ({bi },{χi }) .
Y
dχ1 · · · χN δχ̃i f (χi ) (2.2)
i
In principle, any coarse-graining function for the side chains may be used; however the
discrete approximation Ṽ to the potential provides a more accurate approximation whenever
the distribution of χ-angles is sharply peaked (in the true potential V ) within each discrete
state χ̃. See Figure 2.3 for an example of a coarse-graining function and see section 2.4 where
an optimized coarse-graining function f is derived. We make the following assumptions
on the form of Ṽ . First, we assume there is an explicit function yi (bi , χ̃i ) for the side
chain coordinates based only on the backbone coordinates and side chain state for residue
i. We may relax the requirements to depend on only a single residue’s backbone position,
but the requirement that yi depend on only a single side chain state χ̃i is firm. These
directed coordinates are approximately side chain centers of mass with direction given by
the Cβ –Cγ bond vector direction. However, a parameterization of these side chain position
functions separately for each amino acid type enables the maximization of the accuracy of
the approximation in equation (2.2). A further assumption is that Ṽ can be expressed in
10
the form
(2)
where the pair interaction Vij (yi , yj ) = 0 for the side chain is taken to vanish beyond a
cutoff Rcutoff . Notice that the dependence of the potential on the backbone is completely
general except the potential is assumed to contain at most a pairwise dependence on the
discrete rotamer states χ̃i . Explicit parameterizations for yi and Ṽ are defined in section 2.5
using the principle of maximum likelihood.
One can simulate the Boltzmann ensemble for Ṽ using molecular dynamics for the back-
bone {bi } and Monte Carlo moves for the side chain states {χ̃i }, but the strong steric
interactions lead to a slow equilibration and dynamics for both the side chains and back-
bone. Since we are predominantly interested in backbone motions, we return to the free
energy V̄ in (2.1), now summing over discrete side chain states instead of integrating over
continuous side chain angles,
The benefits of running dynamics with the coarse grained V̄ enter at great cost because
using even three coarse-grained states per side chain implies a summation over 3N χ̃-states
in equation (2.6). Furthermore, the vast majority of those 3N states have steric clashes or
other large energies and, therefore, contribute little to the free energy of the side groups.
To approximate the free energy of the side chains V̄ , it is convenient to express our
problem in the language of Ising models so that we can apply standard techniques developed
in that context. For a fixed backbone configuration {bi },
where the potentials v̄ are written in lowercase to indicate suppression of the dependence
on the fixed backbone coordinates {bi } in order to focus on the side chain contribution.
(1)
Notice that with the backbone positions fixed, each single-residue potential vi is simply
a vector with as many components as the number of possible states for χ̃i (e.g. length-6
(2)
vectors). Similarly, each of the pair potentials vij is a small 6x6 matrix of potential energies
to cover the 36 possibilities. These single and pair potentials are calculated only once before
evaluating the free energy as described in section 2.5. Moreover, the pair summation in
equation (2.7) only applies for residues pairs i and j that are neighbors spatially. A pair
of residues (i, j) are neighbors if inter-residue distance |yi (χ̃i ) − yj (χ̃j )| is less than a cutoff
Rcutoff for any of their possible discrete states (χ̃i , χ̃j ). In this work, we use Rcutoff = 7Å
12
Figure 2.2: Fragment of protein G with associated interaction graph (Rcutoff = 7Å). A pair
of residues has a connection whenever their side chain beads are within Rcutoff for any side
chain states.
for side chain-side chain interactions and Rcutoff = 5Å for side chain-backbone interactions.
The potential Ṽ may be visualized as an energy function on a graph with one discrete
site per amino acid. The graph has a connection between any two residues that are within
the cutoff separation Rcutoff as defined above. This graph is illustrated in Figure 2.2 for a
model protein configuration. The structure of this graph varies dynamically over the course
of a simulation because the definition of neighboring residues depends on the backbone con-
figuration {bi }. The potential varies smoothly as the backbone moves so long as the pairwise
potential functions are continuous in the backbone coordinates. The potential Ṽ is continu-
ous despite the changing connections of the graph, because the strength of the potential for
each interaction approaches zero at Rcutoff just before the connection is eliminated from the
graph. Problems such as this, with discrete potentials on an arbitrary graph, are extensively
studied in both statistical mechanics (as variants of the Ising model) and machine learn-
ing (as undirected graphical models or Markov random fields)[24]. Below we adopt some
well studied approximations from these fields to provide accurate and tractable methods for
computing our coarse-grain potential V̄ .
Two approximations (see [24]) are invoked to compute the free energy from
The first approximation is to express the free energy GSC in terms of the entropy and average
13
energy of the Boltzmann ensemble where the entropy has been replaced by an approximation,
GSC = hv̄i − S
where hv̄i and S approx are defined below. Both the average energy and a mutual information
approximation to the entropy may be expressed using the single-residue probabilities pi (χ̃i )
that residue i is in state χ̃i in the Boltzmann ensemble of v̄ and similarly for the joint
probabilities pij (χ̃i , χ̃j ). Using pi and pij , the approximate energy and entropy are
XX (1)
hv̄i = pi (χ̃i )vi (χ̃i )+
i χ̃i
X X (2)
pij (χ̃i , χ̃j )vij (χ̃i , χ̃j ) (2.10)
i,j χ̃i ,χ̃j
neighbors
S approx = −
XX
pi (χ̃i ) log pi (χ̃i )+
i χ̃i
X X pij (χ̃i , χ̃j )
− pij (χ̃i , χ̃j ) log . (2.11)
pi (χ̃i ), pj (χ̃j )
i,j χ̃i ,χ̃j
neighbors
The mutual information approximation to the entropy ignores contributions from three-
residue and higher correlations. We intend to minimize the approximate free energy (2.9)
over all putative Boltzmann probability distributions for the side chain states {χ̃i }. Notice
that only the 1-side chain probabilities pi and 2-side chain probabilities pij are required to
compute the average energy and approximate entropy; we do not need the more complicated
full joint probability distribution of the {χ̃i } states for all side chains. In addition to the
mutual information approximation of the entropy, we assume that any pair probability pij
represents possible pair probabilities from a Boltzmann distribution, so that the only task
is to minimize the free energy with respect to the pair probabilities. The only constraints
14
imposed are that they must satisfy the obvious consistency conditions for probabilities,
X
pi (χ̃i ) = pij (χ̃j ) (2.12)
j
X
pij (χ̃i , χ̃j ) = 1 (2.13)
χ̃i ,χ̃j
X X
pij (χ̃i , χ̃j ) = pjk (χ̃j , χ̃k ) (2.14)
χ̃i χ̃k
However, use of only the conditions (2.12)-(2.15) is insufficient to ensure that a joint prob-
ability distribution exists for all the variables consistent the with the choices of pi and pij .
As an explicit example,
1/3 0 0
p12 = p23 =
0 1/3 0
(2.16)
0 0 1/3
1/9 1/9 1/9
p13 =
1/9 1/9 1/9
(2.17)
1/9 1/9 1/9
15
Thus, we now have a tractable approximation to free energy of the side chain. We can
minimize that free energy using a self-consistent iteration technique called belief propagation;
see appendix 2.11 for details. The iteration typically converges rapidly, often in 10-20 steps.
Molecular dynamics simulations require calculations of the forces on the backbone coor-
dinates, − ddbV̄ . The derivatives can be computed simply using the chain rule, noting that
i
several terms will be zero because the pair probabilities minimize the free energy,
simplifications occur because S approx is independent of the backbone coordinates. While the
underlying side chain interactions are pairwise additive and vanish outside the cutoff radius
Rcutoff , the free energy (2.9) is a many-body potential that can interact over arbitrary
distances.
Since the approximate free energy due to the side chains is not a convex function of
the probabilities, local minima may arise and impair the the self-consistent iteration from
finding the global minimum. To reduce the danger posed by the presence of local minima,
calculations are begun from a carefully initialized state, as detailed in appendix 2.11. Other
self-consistent approximations exist for the side group free energy, such as tree-reweighted
belief propagation[23], that are typically less accurate but always converge to the global
minimum of their approximate free energy. Another limitation of the present approximation
16
Fine Rotamer States Coarse Rotamer States (6-states)
χ3 and χ4 not pictured minimize uncertainty in atom position
240 240
120 120
χ2
0 0
−120 −120
−120 0 120 240 −120 0 120 240
χ1 χ1
Figure 2.3: Example of optimized coarse states for arginine overlaid on the PDB distribution
of the rotamer angles χ1 and χ2 . Each of the six coarse states contains only a single fine
state that has high probability, so that the variance of dihedral angles within each coarse
state is small.
scheme arises when a bi-stable or multi-stable energy landscape is possible for the rotamer
states. If well-separated and equally important minima are present for a single backbone
configuration in the rotamer free energy surface, the probabilities only converge to a single
minumum and thus underestimate the entropy of the side chains. While this does not
appear to occur near the native well, we have not extensively searched for special backbone
configurations that would result in bi-stable rotamer energies. The characterization of such
problematic configurations, likely near free energy barriers, is left to future work.
The χ-angles for the side chains are partitioned into discrete states in an optimized manner.
The NDRD rotamer library[18] provides a set of approximate discrete states for each residue
type according to their frequencies of occurrence in a non-redundant set of high resolution
protein structures in the PDB. However, the number of rotamer states in the NDRD library
can be quite large. For instance, naively using all 81 rotamers for each arginine, means that
computing the pair interaction vi,j for two arginines would require computing 812 = 6561
energy values. Consequently, instead of using all possible rotamer states, several NDRD ro-
tamer states are combined into 3–6 coarse-grained rotamer states for the sake of manageable
computational cost.
17
We choose aggregate the rotamer states of the side chain to minimize the positional
uncertainty of side chain atoms in each state. A search over all possible aggregations is
conducted to find the aggregation that provides the smallest possible error. More formally,
f
the NDRD rotamer library[18] is used to define the atomic positions xij (φ, ψ), where i is
the atom (such as Cβ ), j is the coordinate (x, y, or z), and f is the fine-grained rotamer
state. Each rotamer state has a probability pf (φ, ψ) specified in the NDRD library from
frequencies in the PDB for each fine-grained rotamer state as a function of the backbone
dihedral angles (φ, ψ). Each fine-grained state f may belong to exactly one coarse-grained
state c (i.e. the c states form a partition of the f states). Given the choice of a coarse-grained
state c, an average is performed over the fine-grained atomic positions, and sum is taken over
the probabilities of all fine-grained states f grouped into c according to the prescription,
q c (φ, ψ) = pf (φ, ψ)
X
(2.20)
s∈t
c (φ, ψ) = 1 X f f
yij pij (φ, ψ) xij (φ, ψ), (2.21)
q c (φ, ψ)
s∈t
The error incurred by coarse-graining is defined as the variance of the atom positions
within each coarse-grained state, weighted by the frequency of occurrence of the coarse-
grained state in the PDB. Specifically, the error σ 2 (φ, ψ) is defined as,
X pf (φ, ψ) X f c(f )
σ 2 (φ, ψ) = (xij (φ, ψ) − yij (φ, ψ))2 , (2.22)
s
Natom
ij
where Natom is the number of atoms in the side chain and c(f ) is the coarse-grained state c
that contains the fine-grained state f . The error depends implicitly on the state decompo-
sition c(f ) and measures the deviation of the atoms within each state. This error favors the
fine-grained states f that have higher frequency of occurrences in the PDB.
The division of fine-grained states into coarse-grained states is restricted for simplicity
18
to be independent of the Ramachandran angles for the residue,
Z
σ2 = pRama (φ, ψ) σ 2 (φ, ψ) dφ dψ, (2.23)
where pRama (φ, ψ) is the frequency of each Ramachandran angle taken from the PDB coil
library. Note that this error term depends implicitly on the decomposition c(f ) and weights
for the (φ, ψ) pairs according to the frequency in the coil library.
An optimal coarse-grained representation of the side chain rotamer states is obtained by
minimizing σ 2 for each residue type over all partitions c(f ). We force the coarse-graining
c(f ) to obey a few conditions, essentially to make sure that c(f ) is easily interpretable in
terms of χ1 and χ2 as well as limiting the number of possibilities that must be checked by the
brute-force minimization. In particular, the mapping from coarse-states back to χ1 rotamer
states is unambiguous because no single coarse state contains two different χ1 rotamer states.
We impose the following conditions,
1. c(f ) depends only on the χ1 and χ2 rotamer states of f (i.e. if f1 and f2 states differ
only in their χ3 or χ4 states, then c(f1 ) = c(f2 ))
2. Each coarse state c must contain only a single χ1 state (i.e. if f1 and f2 have different
χ1 states, then c(f1 ) 6= c(f2 ))
3. Each coarse state c must contain a contiguous range of χ2 values. This greatly reduces
the number of possible coarse-grainings for residues with non-rotameric χ2 angles like
asparagine.
Optimizing the decomposition of the coarse-grained state c(f ) proceeds by completely enu-
merating all possible decompositions into coarse-grained states that contain no more than
six fine-grained states and by imposing the three conditions.
We optimize the decomposition of the coarse-grained state c(f ) by completely enumer-
ating all possible decompositions into coarse-grained states that satisfy the three conditions
19
Restype States Restype States
0.7 1.0
ALA 1 LEU 3
Figure 2.4: Error in the decomposition of rotamer states into coarse-grained states as a
function of the number of side chain states. The position uncertainty is σ. The relative
uncertainty is the position uncertainty for each number of states divided by the accuracy at
3 states. For residues without a rotable χ2 , such as valine, it is not possible to coarse-grain.
One, three or six rotamer states are used, depending on the residue type. The computational
time to compute the pairwise interactions and solve for the free energy scales roughly as the
number of coarse rotamer states squared, so there is an incentive to use as few coarse states
as possible. The table (above) summarizes the number of states chosen for each amino acid
type.
Paralleling the necessity of coarse-graining the rotamer states, side chain atoms also require
coarse-graining in order to obtain an inexpensive side chain model. This reduction in the
number of degrees of freedom may further be justified since the atomic positions of the side
chains are uncertain due to the discretization and aggregation of the rotamer states, meaning
that there is little value in assigning precise positions for all atoms. We instead use a single
oriented bead (location and direction coordinates) to represent each side chain (note that
the direction is independent of the side chain, e.g. in aromatic residues it may be the ring
normal unit vector). The locations and directions of the side chain beads are changed by the
optimizer during the optimization of the potential. The improvement in prediction accuracy
from using optimized side chain positions rather than the static positions is substantial, and
the accuracy of predicting crystallographic rotamer states rose from 48% to 60% as a result
20
of using optimized positions.
While the side chain belief propagation can handle essentially any pairwise residue po-
tential, considerable care is required to design an interaction form that complex enough to
represent a wide range of protein physics and adequately represent side chain packing, yet
simple enough to be trained by gradient descent. To do so, we gather key requirements to
develop a simple protein model.
The basic component is a radial pair interaction between side chain beads. This term can
reproduce an effective side chain excluded volume but it is less obvious that such a model
is appropriate to describe attractive interactions. The attractive protein interactions are
often directional, arising from the dipole moments of polar residues and the ring stacking of
aromatic residues. For this reason, we add a separate directional interaction that depends
on the cosine of the angle between the bead direction and the inter-bead separation vector.
The bead direction is chosen by the optimizer, and the starting point of the optimization is
for the bead direction to be the Cβ –Cγ bond vector. This interaction form is sufficient to
capture simple angular-dependent interactions of side chains, providing a large boost to the
ability of the model to handle side chain hydrogen bonding and other dipolar interactions.
Concretely, each interaction pair is described by positions y1 and y2 and directions n1 and
n2 . From this the distance r12 = |y1 − y2 | and displacement unit vector n12 = (y1 − y2 )/r12
are calculated. The form of the interaction is given by
V = κ( unif(r12 )+
where unif, ang1 , ang2 , and dir are smooth curves. The smooth curves are represented by
cubic splines. The prefactor κ is 1 for interactions between two side chain beads.
For side chain-backbone interactions, however, it is advantageous to have the prefactor
κ depend on the hydrogen bonding state of the backbone residue because the presence of
21
RMSD
alignment
x1
SC
1
ref
pos
θ1
backbone
posi;on
in
simula;on
r
SC
2
θ2
ref
pos
RMSD
x2
Cα
alignment
Cα
N
C
N
C
V (r, θ1, θ 2 ) = unif(r)
+ ang1 (θ1 )•ang 2 (θ 2 )•dir(r)
Figure 2.5: Left panel depicts side chain interaction coordinates for equation (2.24). Right
panel depicts side chain interactions with backbone hydrogen (blue) and oxygen (red) sites;
axes labels indicated on lower right plot. Within each group of 4 plots, the curves on the
left hand side present the radial components of the interaction. The thin lines provide the
unif(r) interaction and the thick line is the directional interaction unif(r) + dir(r). The right
plots are heat maps of the angular interactions ang1 (θH/O ) ang2 (θSC ), where the lower left
corner depicts the interaction where θ1 = θ2 = 0◦ . Within each group of 4 plots, the upper
plots represent side chain interactions with hydrogen and the bottom plots represent side
chain interactions with oxygen.
22
one hydrogen bond inhibits forming another. Specifically, the interaction between a back-
bone hydrogen or oxygen is given a hydrogen bond confidence score f , a number that is
typically close to 0 for non-Hbonded and 1 for Hbonded residues. We set κ = 1 − f so that
the interaction is only turned on for hydrogens or oxygens that are not participating in a
backbone-backbone hydrogen bond. The physical motivation is that the directional interac-
tion primarily describes the effects of the dipole interactions, and in a hydrogen bond the
C=O and N–H dipoles approximately cancel each other. While it is theoretically possible for
the algorithm to learn carefully balanced hydrogen and oxygen interactions that themselves
cancel out on hydrogen bonded pairs, it is much easier to achieve a physically-reasonable
model if we enforce the zeroing of directional interactions with already hydrogen-bonded
pairs.
The side chain-backbone interactions are needed to describe helix capping. We have
observed that a proper description of these capping effects is required to avoid helix fraying.
Furthermore, Harper and Rose[8] have observed that N-terminal capping of a helix by side
chains is more likely to be observed than is C-terminal capping of the side chain. This finding
is consistent with our maximum likelihood training (below), where side chain-amide hydrogen
interactions are fit with stronger (i.e. higher confidence) potentials than side chain-oxygen
interactions. Harper and Rose also note that hydrophobic residues play a strong role in helix
capping by covering exposed protein backbone at the ends of helices. To provide our model
with the freedom to describe this effect, an additional side chain-backbone interaction is
added with three beads representing the hydrophobic portion of the backbone. The location
of the three beads are initialized from the reference position of N, Cα , and C and are
optimized with the rest of the parameters. For this interaction, κ = 1.
23
2.6 Maximum likelihood training
The side chain model is trained by the maximum likelihood principle. Specifically, we deter-
mine the set of parameters that maximizes the log probability of the true side chain states χ̃p
in the Boltzmann ensemble of all possible side chain states χ̃ for the fixed backbone positions
Xp for each protein p.
e−V (χ̃p )
p(χ̃p ) = P (2.25)
−V (χ̃)
χ̃ e
e−V (χ̃)
X
− log p(χ̃p ) = V (χ̃p ) + log (2.26)
χ̃
= Egap . (2.28)
The evaluation of Egap requires the evaluation of the free energy of the side chains, a quantity
that is intractable to calculate exactly. Fortunately, our side chain energy (2.18) approxi-
mates the true side chain free energy GSC that appears in (2.27). Furthermore, the expression
for the parametric derivative (2.19) allows for gradient descent optimization to minimize the
average gap energy.
The side chain packing interaction is trained using a large, non-redundant collection of
crystal structures from the PDB with 50–500 residues and resolution less than 2.2Å. From
a training set of protein structures, we extract the sequences sp , backbone trace positions
Xp , and true coarse-grained side chain states the χ̃p for each protein p. The proteins are
further filtered using PISCES[25] so that all pairs of proteins have sequence similarity less
than 30%. Non-globular structures in the dataset are removed, as we suspect that the side
chain packing of these structures are more strongly influenced by other chains in the crystal
24
structures. We define non-globular structures as outliers in the linear relationship between
log(Nres ) and log(Rg ); the outliers are identified using the RANSAC algorithm[6]. After
filtering, 6255 chains remained, containing approximately 1.4 million residues.
2.6.2 Regularization
Since there are 210 types of amino acid pairs and the potential for each pair has 10s of
associated parameters, the model contains more than 10000 parameters just for side chain
pair interactions. When maximizing the likelihood for a system with such a large numbers
of parameters, it is often beneficial to add a penalty term, called a regularizer or a maximum
a posteriori prior, that favors simpler models, greatly reduces overfitting, as well as encour-
ages models that will better generalize to molecular dynamics simulations. Two types of
regularization penalties are contrasted. The first penalty simply encourages lower energies
except at the repulsive core of the interaction.
The second penalty is a lower bound penalty for the energy that very strongly discour-
ages energies below a certain threshold, where the lower bound is chosen to be as strict
as possible without significantly reducing accuracy. The term is needed empirically as the
model sometimes learns very low energies for certain side chain-backbone hydrogen bonding
for residues such as aspartic acid, on the order of -15 kT. We suspect that these large en-
ergies reflect systematic differences between crystallographic structures and the Boltzmann
ensemble of proteins in physiological conditions. If an aspartic acid can hydrogen bond to
the backbone, the electron density of rotamers with unbound states are likely to be quite
small and the crystallographer refining the structure will likely not register the minor alter-
native rotamer (additionally, we do not use multiple states for a single residue). This effect
is likely exacerbated by the low temperature of crystallization. As a result, an unrealistically
favorable energy is learned for side chain-backbone hydrogen bonding. This over-estimation
is allowable if the only objective is accurate predictions of side chain rotamers, but such large
energies (greater than the backbone-backbone hydrogen bond energy by a large amount) are
25
problematic when running dynamics. Enforcing a lower bound on the energies suppresses
these large energies.
5kT
Z
Ereg =kr f V (r, θ1 , θ2 ) − dr cos(θ1 )dθ1 cos(θ2 )dθ2 +
1. + e r/(0.2Å)−10
Z
2 (V (r, θ1 , θ2 ) − Vlb )2− dr cos(θ1 )dθ1 cos(θ2 )dθ2 (2.29)
(x)− is zero whenever x is positive and x otherwise. The function f is the logarithm of the
student-t distribution, which is convenient for regularizing while still allowing large energies
when necessary. The hyperparameters kr and Vlb are chosen below based on the results of
the maximum likelihood training.
To check for overfitting, 20% of the proteins are randomly chosen to be left out of the
optimization as a validation set. Decisions on the functional form of the interactions and
the regularization parameters kr and Vlb are made based on data that is not used in the
gradient descent optimization.
The Adam optimizer[12] is used to minimize the energy gap. This optimizer is convenient
because it automatically adjusts the gradient descent step size for each parameter according
to the typical scale of the gradient in that dimension. This rescaling is important because
spline coefficients at large radii tend to have much larger gradient magnitudes than param-
eters at small radii. For full details of the optimization including initialization, please see
section 2.10.
total total
acc = e−Egap /Nres . (2.30)
26
0.61
0.60
validation accuracy
0.59
0.58
0.57
0.56 lower bound -3
lower bound -6
0.55 1
10 102 103
prior width
Figure 2.6: Accuracy of the model (probability assigned to true χ̃ state) versus regularization
strength (essentially the inverse of kr ). Regularization as employed in this work does not
appear to improve the accuracy of the side chain packing significantly. The regularization,
especially the lower bound to the energy, may have a strong effect on the accuracy of protein
simulation, even if does not increase packing accuracy.
This represents roughly the geometric mean over residues in the test set of the probability
assigned to the true side chain state.
Figure 2.5 depicts the optimized side chain interactions with backbone hydrogen and oxy-
gen. Most of the trends in the figure can be explained by helix capping motifs described in
[8]. The prominent role of side chain-hydrogen interactions is consistent with the observation
that side chain-backbone hydrogen bonding is more common on the N-terminus than the
C-terminus. The strong valine-oxygen interaction is more puzzling, since we do not expect
strong, favorable interactions of oxygen atoms with hydrophobic residues. The strong valine
interaction may be caused by statistical correlations of valine with the geometry of oxygen
atoms that are associated with the hydrophobic capping of helices by valine. The valine
interaction represents a likely weakness in the training, as the statistical training that maxi-
mizes side chain packing accuracy may not capture the physics that drives protein backbone
dynamics. The physical reasonableness of the other terms, however, is encouraging.
To compare to state of the art side chain prediction methods, we compare to SCWRL4[14]
on its training and validation set of side chains conformations. As per SCWRL’s validation
procedure, the side chains with less than 25th percentile electron density are excluded. To
27
Model No interactions SCWRL4 Upside
1.00
0.75
accuracy 0.50
0.25
0.00
GLU SER ARG MET GLN HIS LYS TRP ASN CYS TYR LEU ASP PHE THR VAL ILE
restype
Figure 2.7: Comparison of χ1 prediction accuracy for Upside and SCWRL4. The “No
interactions” line represents the accuracy of the only the NDRD rotamer library without
any interactions; this library is used in both Upside and SCWRL. Note that Upside is
approximately 150x faster than SCWRL at side chain prediction, in addition to return a
probability distribution instead of a single answer.
avoid biasing the comparison toward Upside, the SCWRL set of proteins is split so that 20%
of the proteins are withheld for validation, while the rest are used for maximum likelihood
training of Upside. The accuracy metric chosen is to calculate the fraction of side chains,
excluding glycine, alanine, and proline, for which the Upside or SCWRL predicted χ1 angle
agrees with the crystallographic conformation. This accuracy metric is typically used to
assess side chain prediction and is always larger than the geometric mean accuracy used for
Upside training for a properly calibrated method.
As seen in Fig 2.7, Upside is very accurate, predicting the correct χ1 conformation 87.6%
of the time. SCWRL4 is slightly more accurate, achieving 89.4% correct predictions, but
with a number of limitations. First, Upside is approximately 150x faster than SCWRL4 at
predicting side chain conformations with 98% of the accuracy. This enables us to compute
the distribution of side chain positions at every step of molecular dynamics at modest cost.
Second, Upside provides a Boltzmann probability distribution over rotamer states, enable
28
molecular dynamics using exact forces from the approximate side chain ensemble. SCWRL4
is not well suited for continuous dynamics because it provides only the lowest energy confor-
mation, causing inevitable discontinuities in any attempt to compute forces using SCWRL’s
predicting side chain conformations.
Strictly speaking, the parameters obtained from the maximum likelihood training are only
optimal for side chain packing for a fixed, native-like backbone geometry. However, we
believe that they also encode energetics that can be applied to molecular dynamics simula-
tions. Specifically, in the limit that the model is flexible enough to model the true side chain
interactions and there is unlimited training data, the maximum likelihood method would
recover the true side chain interaction. Even without having the true form of the side chain
interaction, the maximum likelihood parameters assign high probability to the observed ro-
tamer states, thereby including at least some of the underlying physics. The degree to which
packing suffices provides a useful energy function to simulate correct backbone structures
will be investigated below.
There are caveats to using side chain packing parameters for protein dynamics, even
though the parameters are in principle governed by the same physics. The first issue is that
a free backbone may move to an unlikely conformation that differs qualitatively from the
configurations in the crystallographic data set (e.g., poorly packed or less dense) and the
parameters may assign an inappropriately low energy to this conformation. This happens
most commonly with the strong parameters of side chain-backbone hydrogen bonding. The
backbone may adjust to enable more side chain-backbone hydrogen bonds than are physi-
cally realistic because this interaction is geometrically-constrained and hence, inordinately
strong. This problematic effect has been observed and may have multiple causes. The first
cause is that insufficient strength in other interactions (such as hydrophobic burial) may pro-
vide insufficient penalty to the unusual configurations that maximize side chain-backbone
29
hydrogen bonding. The second cause is that the data are trained on low temperature crystal
structures. Hydrogen bonds and other interactions may stabilize at low temperature, addi-
tional structure may form, and/or the minor populations with side chain-backbone hydrogen
bonds broken may not be identified or present in the electron density. We anticipate that for
the case of side chain-backbone hydrogen bonding, hydrogen exchange protection factors or
NMR observables may be able to resolve the true populations of these interactions. Crystal
packing artifacts, as well as limiting training to crystallizeable sections of proteins, may also
affect the generated parameters. These artifacts are expected to have a weak effect, though
they may bias the model to unphysically bias against unstructured loop regions.
To test the suitability of adapting the side chain packing model to molecular dynamics,
simulations were run from the native state of a set of small, fast-folding proteins (protein
set adapted from [15]. To create a reasonable protein dynamics model, backbone springs,
backbone sterics, hydrogen bond energy, and a basic Ramachandran potential were added
to the side chain model. The Ramachandran potential is derived from a coil library[22] as
a statistical potential. The hydrogen bond strength is chosen using trial simulations. We
choose the hydrogen bond strength −1.8 kT to maximize the median fraction of simulation
frames with RMSD less than 5Å from the native state over the course of a short simulation.
For simulation details, see the appendix 2.10. Note that because alanine and glycine have
no side chain rotamer states, and hence no training to match the native χ-angles can be
conducted, the ALA-ALA, ALA-GLY, and GLY-GLY potentials are completely determined
by the regularization. Interactions of ALA and GLY with other residue types are optimized,
however, as rotamer states of the other residues provide information on the ALA-X and
GLY-X interactions.
The simulations results show that for the majority of proteins, the Upside model does not
assign the lowest free energy to the native structure (Fig 2.9). For most proteins, the native
structure in the Upside model is however temporarily stable, indicative a local minimum of
the free energy surface. The model relaxes quickly to its preferred structure.
30
0.2
0.0
2.4 2.2 2.0 1.8 1.6
backbone-backbone HBond energy (kT)
Figure 2.8: Accuracy over short duration simulations. The backbone hydrogen bonding
strength is not determined by the packing optimization, so we search for the strength that
gives the best simulation accuracy. This is the only parameter in the model directly optimized
for simulation accuracy. To assess accuracy, we look at the fraction of each simulation with
RMSD to native of less than 5 Å, and we compute the median accuracy over all proteins in
the test set. All other results shown are for a backbone hydrogen bond energy of −1.8 kT .
Figure 2.9: RMSD to native over replica exchange simulation trajectories with hydrogen
bond energy −1.8 kT .
31
2.8 Related Work
In the vast literature of coarse-grained modeling, we highlight several strands of work that
relate to our modeling. The major features of our model include the following: molecular
dynamics on three atoms but with a dynamic ensemble of side chains, optimized discretiza-
tion of the side chain states to best represent the protein interactions in the coarse-grained
model, statistical potential with optimized and state-dependent bead locations and orienta-
tions, training a protein interaction model for folding using side chain packing accuracy, and
a side chain model with an explicit side chain entropy.
A large body of work, exemplified by SCWRL[14], have studied the prediction of side
chain configurations by discrete rotamer states. SCWRL achieves greater than 90% χ1 ac-
curacy for predicting the most likely rotamer states by minimizing the energy that combines
observed rotamer state frequencies and an atomic interaction model[14]. A variety of algo-
rithms have been developed for solving for the highest probability side chain states given the
pair interaction values[4, 26]. Kamisetty et al.[11] have worked on scoring protein interaction
complexes using a self-consistent approximation to the side chain interactions. Earlier sim-
ulation work by Koehl and Delarue[13] use 1-residue mean field techniques to approximate
ensembles of side chain conformations but fail to account for the pairwise correlations of the
side chain rotamer states. All of these works use atomically-detailed descriptions of the side
chains paired with simple or molecular dynamics interaction forms. Their highly detailed
side chain with many χ-angles for each residue makes it difficult to perform dynamics suf-
ficiently quickly for folding, and the use of existing interactions (instead of a newly-trained
interaction model) makes it difficult to use reduced detail to speed computation. There has
also been extensive work in reconstructing backbone positions from side chain beads[5] in
lattice models, but these models do not perform a proper summation over possible rotamer
states.
There have also been a large variety of coarse-grained techniques that use a variety of
non-isotropic potentials for reduced side chain interactions. One of the most successful is the
32
coarse-grained united residue model (UNRES)[16]. The model also uses statistical frequen-
cies to determine the positions of the side chains but it emphasizes the parameterization of
the coarse-grained model from physics-based calculations instead of statistical information.
Though the potential form (Gay-Berne) used in UNRES is quite different from our work,
UNRES also uses non-isotropic side chain potentials[19].
Similar to our work, Dama, Sinitskiy, et al.[3] investigate mixed continuous-discrete dy-
namics, where the states of molecules jump according to a discrete Hamilitonian. Their
method differs from our work in a number of important ways: the authors use discrete
jumps in state instead of a free energy summation over all states we employ; they do not
optimize the rotamer states as we do; and they train parameters from force matching of
molecular dynamics trajectories rather than from the statistical analysis of experimental
data in our method.
2.9 Conclusion
We have demonstrated a fast, principled method to coarse-grain discrete side chain states
to create a smooth backbone potential. This procedure results in a considerable decrease
in computational time as it removes the side chain rattling and friction normally associated
with a polypeptide chain moving in a collapsed state. This tracking and instantaneous
equilibration of the side chains is analogous to the instantaneously-equilibrated electronic
degrees of freedom with respect to the nuclear motions employed in the adiabatic Born-
Oppenheimer approximation[2]. Motions are calculated only for three heavy backbone atoms,
yet the model contains considerable structural detail including hydrogen bonds involving
both the backbone and side chains. Further, we have shown how to parameterize both a
tuneable discretization of the rotamer states, and a maximum liklihood procedure to obtain
physically-reasonable parameters for our coarse-grain model from X-ray structures. The
resulting method is capable of rapid molecular dynamics sampling of protein structures.
The importance of optimizing the bead locations and directions in our model illustrates
33
the principle that chemical intuition can only be a partial guide to accurate coarse-graining
of protein interactions. The location of the interaction sites has a strong effect on our model’s
ability to achieve high-packing accuracy, and we expect similarly strong effects to be observed
had we directly optimized for backbone conformational accuracy.
While the side chain packing optimization shows promise as a route to accurate and
inexpensive molecular simulation, pairing the resulting potentials with simple Ramachandran
and hydrogen bond potentials can maintain the structure of a minority of small proteins
tested. Future work will develop co-training of both side chain and backbone parameters to
improve simulation accuracy.
All simulations are run with Upside, a custom simulation engine that implements the belief
propagation of side chain interactions as well as the parameter derivatives needed for gradient
descent. Upside is freely available and open source[9].
The temperature is 0.7 natural units. The Ramachandran potential uses the NDRD
TCB coil library[22]. The backbone hydrogen bond interaction uses both distance and
angle criteria to determine hydrogen bonds. The H-O bond distance interaction starts at
approximately 1.4Å and ends at 2.5Å. Both the N-H-O and H-O-C criteria half-heights are
at approximately 47 degrees off of collinear.
We use Verlet integration with a time step of 0.009 units. We use the random number
generator Random123 [17] to implement the Langevin dynamics with a thermalization time
scale of 0.135 time units. The thermalization time scale (related to Langevin friction) is
chosen to maximize the effective diffusion rate of chains while effectively thermostatting the
simulation. As Langevin dynamics with any friction coefficient produces the same Boltzmann
ensemble, we chose to maximize equilibration of our system rather than attempt to match a
solvent viscosity.
The derivative calculations need for regularization and coordinate transforms necessary
34
to ensure positive coefficients are handled with the Theano framework[21].
The cutoff radius for side chain-side chain interactions is 7Å, and the cutoff radius for
side chain-backbone interactions is 5Å. The distance splines are zero-derivative-clamped
cubic splines with a knot spacing of 0.5Å. The angular splines have a knot spacing of 0.167
in cos θ, which ranges over [−1, 1].
We use the following settings for the Adam optimizer: minibatch size 256 proteins,
α = 0.03, β1 = 0.90, β2 = 0.96, = 10−6 . Positivity constraints on the angular coefficients
are enforced by a exponential transform. The regularization integrals over all space are
approximated by sums at the knot locations of the radial and angular splines.
For convenience, this appendix contains a brief description of the equations used to imple-
ment belief propagation for the side chain free energies. Given 1-residue energies vi (χ̃i ) and
2-residue energies vij (χ̃i , χ̃j ), we seek probabilities pi (χ̃i ) and pij (χ̃i , χ̃j ) to minimize the
free energy (2.18).
It is helpful to first understand the intuition behind the belief propagation process. We
seek a consistent set of one- and two-side chain probabilities for the residues compatible with
the interaction potential (2.7). The probability of each residue state χ̃i for residue i is deter-
mined by two factors. The first factor is the 1-residue energy vi (χ̃i ) that would determine
the probabilities exactly in the absence of interactions. The second factor is consistency
with the side chain states of the residues in contact with residue i, where consistency is
determined by the potentials vij (χ̃i , χ̃j ). Using these factors, the probabilities for residue i
are estimated as
P
pi (χ̃i ) ∝ e−vi (χ̃i )− j wij (χ̃i ) (2.31)
35
where wij (χ̃i ) is the effective 1-body potential that residue i feels due to the interaction with
residue j. The wij depends implicitly on the probability distribution pj (χ̃j ) of residue j, so
the equations (2.31) and (2.32) must be solved by self-consistent iteration until convergence of
the {pi }. This algorithm is distinguished from a standard mean-field iteration, which would
P
be identical except the mean-field algorithm would set wij (χ̃i ) = χ̃j pi (χ̃j )vij (χ̃i , χ̃j ). It
should be emphasized that, despite the appeal of the intuitive explanation above, the real
justification of belief propagation is that the process minimizes the approximate free energy
(2.18) as derived in [27]. The iteration is described more formally below, including a damping
term λ to suppress oscillations during the self-consistent iteration.
For 1-residue beliefs, define bri (χ̃i ) to be the round r “belief” that the i-th residue is in
state χ̃i . For the 2-residue beliefs, we have two beliefs for each pair of interacting residues
(i.e. any pair of residues that have non-zero interaction in any rotamer states). Define brij (χ̃j )
to be the round r belief for the residue pair (i,j) that residue j is in state χ̃j . The belief
bji (χ̃i ) is defined similarly.
To initialize the algorithm at round 0, we take
We compute the round r + 1 beliefs from the round r beliefs according to the following
equations.
The products in equation (2.35) should be understood as taken only over residues j that
interact with residue i. The damping constant λ suppresses oscillatory behavior that hin-
36
der convergence (λ = 0.4 is used in the present work). The equations are iterated until
|br+1 r
i (χ̃i ) − bi (χ̃i )| < 0.001 for all residues i and states χ̃i .
From the converged beliefs bi (χ̃i ) and bij (χ̃j ), we can compute the marginal probabilities
The free energy of the model is obtained by using the marginal probabilities above in equation
(2.18).
Mutations from the indicated PDB structures are indicated in bold. The NuG2 sequence is
from reference [15].
Name PDB ID Length Sequence
alpha3d 2a3d 73 MGSWAEFKQRLAAIKTRLQALGGSEAELAAFEKEIAA
FESELQAYKGKGNPEVEALRKEAAAIRDELQAYRHN
BBA 1fme 28 EQYTAKYKGRTFRNEKELRDFIEKFKGR
BBL 2wxc 47 GSQNNDALSPAIRRLLAEWNLDASAIKGTGVGGRLTREDVEKHLAKA
homeodomain 2p6j 52 MKQWSENVEEKLKEFVKRHQRITQEELHQYAQRLGLNEEAIRQFFEEFEQRK
lambda 1lmb 80 PLTQEQLEDARRLKAIYEKKKNELGLSQESVADKMGMGQS
GVGALFNGINALNAYNAALLAKILKVSVEEFSPSIAREIY
NTL9 2hba 39 MKVIFLKDVKGMGKKGEIKNVADGYANNFLFKQGLAIEA
protein B 1prb 53 TIDQWLLKNAKEDAIAELKKAGITSDFYFNAINKAKTVEEVNALKNEILKAHA
protein G 1pga 56 MTYKLILNGKTLKGETTTEAVDAATAEKVFKQYANDNGVDGEWTYDDATKTFTVTE
NuG2 (Shaw) 1mi0 57 MDTYKLVIVLNGTTFTYTTEAVDAATAEKVFKQYANDAGVDGEWTYDAATKTFTVTE
protein L 2ptl 61 VTIKANLIFANGSTQTAEFKGTFEKATSEAYAYADTLKKDNGEYTVDVADKGYTLNIKFAG
ubiquitin 1ubq 76 MQIFVKTLTGKTITLEVEPSDTIENVKAKIQDKEGIPP
DQQRLIFAGKQLEDGRTLSDYNIQKESTLHLVLRLRGG
WW domain 2f21 33 KLPPGWEKRMSADGRVYYFNHITNASQWERPSG
2.13 References
[1] Helen M Berman, John Westbrook, Zukang Feng, Gary Gilliland, Talapady N Bhat,
Helge Weissig, Ilya N Shindyalov, and Philip E Bourne. The protein data bank. Nucleic
acids research, 28(1):235–242, 2000.
37
[2] Max Born and Robert Oppenheimer. Zur quantentheorie der molekeln. Annalen der
Physik, 389(20):457–484, 1927.
[3] James F Dama, Anton V Sinitskiy, Martin McCullagh, Jonathan Weare, Benoı̂t Roux,
Aaron R Dinner, and Gregory A Voth. The theory of ultra-coarse-graining. 1. general
principles. Journal of Chemical Theory and Computation, 9(5):2466–2480, 2013.
[4] Johan Desmet, Marc De Maeyer, Bart Hazes, and Ignace Lasters. The dead-end elimi-
nation theorem and its use in protein side-chain positioning. Nature, 356(6369):539–542,
1992.
[5] Michael Feig, Piotr Rotkiewicz, Andrzej Kolinski, Jeffrey Skolnick, and Charles L.
Brooks. Accurate reconstruction of all-atom protein representations from side-chain-
based low-resolution models. Proteins: Structure, Function, and Bioinformatics,
41(1):86–97, 2000.
[6] Martin A Fischler and Robert C Bolles. Random sample consensus: a paradigm for
model fitting with applications to image analysis and automated cartography. Commu-
nications of the ACM, 24(6):381–395, 1981.
[7] Thomas Hamelryck, Mikael Borg, Martin Paluszewski, Jonas Paulsen, Jes Frellsen,
Christian Andreetta, Wouter Boomsma, Sandro Bottaro, and Jesper Ferkinghoff-Borg.
Potentials of mean force for protein structure prediction vindicated, formalized and
generalized. PloS one, 5(11):e13714, 2010.
[8] Edwin T Harper and George D Rose. Helix stop signals in proteins and peptides: the
capping box. Biochemistry, 32(30):7605–7609, 1993.
38
[10] John M Jumper, Karl F Freed, and Tobin R Sosnick. Maximum-likelihood, self-
consistent side chain free energies with applications to protein molecular dynamics.
arXiv preprint arXiv:1610.07277, 2016.
[11] Hetunandan Kamisetty, Eric P Xing, and Christopher J Langmead. Free energy esti-
mates of all-atom protein structures using generalized belief propagation. Journal of
Computational Biology, 15(7):755–766, 2008.
[12] Diederik Kingma and Jimmy Ba. Adam: A method for stochastic optimization. arXiv
preprint arXiv:1412.6980, 2014.
[13] Patrice Koehl and Marc Delarue. Application of a self-consistent mean field theory
to predict protein side-chains conformation and estimate their conformational entropy.
Journal of molecular biology, 239(2):249–275, 1994.
[14] Georgii G. Krivov, Maxim V. Shapovalov, and Roland L. Dunbrack. Improved prediction
of protein side-chain conformations with scwrl4. Proteins: Structure, Function, and
Bioinformatics, 77(4):778–795, 2009.
[15] Kresten Lindorff-Larsen, Stefano Piana, Ron O Dror, and David E Shaw. How fast-
folding proteins fold. Science, 334(6055):517–520, 2011.
[16] Adam Liwo, Jaroslaw Pillardy, Cezary Czaplewski, Jooyoung Lee, Daniel R Ripoll,
Malgorzata Groth, Sylwia Rodziewicz-Motowidlo, Rajmund Kamierkiewicz, Ryszard J
Wawak, Stanislaw Oldziej, et al. Unres: a united-residue force field for energy-based
prediction of protein structure?orgin and significance of multibody terms. In Proceedings
of the fourth annual international conference on Computational molecular biology, pages
193–200. ACM, 2000.
[17] John K Salmon, Mark A Moraes, Ron O Dror, and David E Shaw. Parallel random
numbers: as easy as 1, 2, 3. In 2011 International Conference for High Performance
Computing, Networking, Storage and Analysis (SC), pages 1–12. IEEE, 2011.
39
[18] Maxim V Shapovalov and Roland L Dunbrack. A smoothed backbone-dependent ro-
tamer library for proteins derived from adaptive kernel density estimates and regressions.
Structure, 19(6):844–858, 2011.
[19] Adam K Sieradzan, Pawel Krupa, Harold A Scheraga, Adam Liwo, and Cezary
Czaplewski. Physics-based potentials for the coupling between backbone-and side-
chain-local conformational states in the united residue (unres) force field for protein
simulations. Journal of chemical theory and computation, 11(2):817–831, 2015.
[20] John J Skinner, Wookyung Yu, Elizabeth K Gichana, Michael C Baxa, James R Hin-
shaw, Karl F Freed, and Tobin R Sosnick. Benchmarking all-atom simulations using
hydrogen exchange. Proceedings of the National Academy of Sciences, 111(45):15975–
15980, 2014.
[21] Theano Development Team. Theano: A Python framework for fast computation of
mathematical expressions. arXiv e-prints, abs/1605.02688, May 2016.
[22] Daniel Ting, Guoli Wang, Maxim Shapovalov, Rajib Mitra, Michael I Jordan, and
Roland L Dunbrack Jr. Neighbor-dependent ramachandran probability distributions of
amino acids developed from a hierarchical dirichlet process model. PLoS Comput Biol,
6(4):e1000763, 2010.
[23] Martin J Wainwright, Tommi S Jaakkola, and Alan S Willsky. Tree-reweighted belief
propagation algorithms and approximate ml estimation by pseudo-moment matching.
In AISTATS, 2003.
[24] Martin J Wainwright and Michael I Jordan. Graphical models, exponential families,
and variational inference. Foundations and Trends R in Machine Learning, 1(1-2):1–305,
2008.
[25] Guoli Wang and Roland L Dunbrack. Pisces: a protein sequence culling server. Bioin-
formatics, 19(12):1589–1591, 2003.
40
[26] Jinbo Xu and Bonnie Berger. Fast and accurate algorithms for protein side-chain pack-
ing. Journal of the ACM (JACM), 53(4):533–557, 2006.
[27] Jonathan S Yedidia, William T Freeman, and Yair Weiss. Understanding belief propa-
gation and its generalizations. Exploring artificial intelligence in the new millennium,
8:236–239, 2003.
41
CHAPTER 3
CONTRASTIVE DIVERGENCE
3.1 Introduction
A major challenge is to extract from the collection of crystallized proteins a suitable potential
that captures the physics that lead to the protein structures. We attack this problem directly
by showing that a strong connection exists between the shape and location of the native
basin and the rest of the protein’s conformational landscape, and this connection is strong
enough to train a potential energy function accurate enough for de novo folding simulations.
Furthermore, the resulting potential is inexpensive enough to converge simulations of small
proteins while running for hours to days on a single computer.
Traditional reference-state methods for determining statistical potentials are difficult to
use correctly or to systematically improve. These methods compute a statistical potential
using the ratio of the observed counts of residues at a specific distance as compared to the
“reference” probability distribution that would be observed when the potential is absent.
As Hamelryck[9] points out, these methods can be formally valid, but the true reference
frequency depends on all of the other potential energy terms. This creates a self-consistency
problem as the reference state for each potential depends on the values of all other potentials.
The traditional solution to this difficulty is to avoid self-consistency by simply postulating a
reference frequency as is done, for example, in the DOPE model[21]. The great difficulty in
providing a good reference state for even simple coordinates is reflected in the great variety of
reference states that have been proposed in the literature. To use the self-consistent reference-
state method correctly requires computations significantly more involved than those proposed
in this work, and this is probably too expensive for current computational resources.
In contrast to traditional reference-state methods, we develop a trajectory-based method
to parameterize protein force-fields. Separately fitting individual parameters in the force-
field inevitably leads to balance issues. Large-scale issues of energy balance, such as the
42
hydrogen bond terms overwhelming the side chain interactions to make very long helices,
may fought by weighting each term in the force-field, but more subtle issues, such as the
balance between hydrophobic and charged interactions within the side chains terms, will be
missed. The essence of these ideas is that the only way to correctly parameterize a statistical
potential is to jointly choose the parameters to optimize the resulting protein ensemble. A
contribution of this paper will be to show that we do not need to be able to converge
the entire ensemble from proteins in our training set, but we may focus our attention on
fluctuations near the native well. By choosing parameters to optimize the near-native basin
of the protein, we will obtain a balanced and accurate set of parameters that is sufficient to
fold proteins.
Further, trajectory-based training fundamentally alters the relationship between the com-
putational speed of a model and its accuracy. Traditionally, the level of detail in both atoms
represented and interaction terms are increased in order to increase accuracy, leading to
slower simulation for more accurate models. When optimizing parameters based on finite
trajectories, this speed-accuracy tradeoff is replaced to some extent by a speed-accuracy
synergy. Models that can be equilibrated more quickly, in CPU-time, are more thoroughly
explored by finite trajectories and therefore optimized closer to the ideal parameters for
the set functional form. Furthermore, training with inexpensive models allows us to opti-
mize using trajectories from approximately 500 crystal structures in the Protein Data Bank
(PDB). We use the Upside model that has been shown to equilibrate extremely quickly while
retaining molecular detail, as shown in chapter 2.
We will use a two part strategy for building our model. First, we will describe a physically-
plausible coarse-grained model that incorporates many of the key features of protein struc-
ture and folding. To find the parameters to populate our model, we will rely on holistic
assessment of accuracy, rather than parameterizing based on transfer studies or quantum
calculation. We incorporate proven techniques from the machine learning community, who
train very complex models on few examples with success in de novo prediction. Our modi-
43
fied contrastive divergence method is capable of training the model from exploration around
the native state. Even though we only trained for a very limited metric of quality (local
exploration), our parameters are plausible as evidenced by the reasonable folding thermody-
namics.
We make the following contributions in this paper. We give a clear and proper statistical
framework that is practical for training highly detailed protein models. We extend previous
contrastive divergence work with molecular dynamics, handling of crystallographic artifacts,
and much larger number of parameters. We characterize the ability of contrastive divegence
to handle extreme numbers of force-field parameters in a simulation setting. We demon-
strate that careful statistical handling of the parameters results in a model with transferable
accuracy in three ways: (1) it extends to de novo folding despite only being trained on native
stability, (2) it produces cooperative folding without being trained for it, and (3) it produces
realistic unfolded states with high Rg , which has proven difficult for molecular dynamics
models.
This section recapitulates the Upside model and describes the differences between the Upside
model used in this chapter, and the model used for folding simulations in the previous
chapter. For exhaustive details of the Upside model, see chapter 4.
The key insight of chapter 2 is that a suitable self-consistent free energy can be defined
from the side chain interactions, and that free energy provides an smooth potential for the
simulation of backbone dynamics. Our previous work parameterized the model to maximize
the accuracy of side chain conformations. While these parameters are used as the initial
parameters for optimization, all interaction parameters are retrained in the current work
except for the locations and directions of the side chain beads. Our current work extends
the Upside model with an additional pseudo-solvation term and better-adapted backbone
Ramachandran potential.
44
The majority of interaction parameters belong to the pair interactions between side
chains. All of the pairwise interactions have the same functional form
V = κ( unif(r12 )+
where unif, ang1 , ang2 , and dir are arbitrary curves represented by cubic splines represented
by cubic splines. For details of the distance coordinate r12 and the angular coordinates n1 ,
n2 , and n12 , see the Fig. 2.5. There are 20
2 = 210 types of amino acid pairs and 62 spline
coefficients per pair, giving 13020 side chain-side chain interaction parameters. There are
also five interaction sites on the backbone, roughly representing the H, O, N, Cα , and C
atoms, with 54 parameters per interaction due to a smaller cutoff radius. The total number
of side chain-backbone interaction parameters is 5400. The cutoff radius of the side chain-
side chain interactions is 7Å and the cutoff radius for the side chain-backbone interactions
is 5Å. The smaller cutoff radius for side chain-backbone interactions encourages the model
to use this term to describe side-chain backbone sterics and hydrogen bonding, as well as
reducing the computation expense from these terms.
We add an additional term, similar in spirit to reference [1], to capture solvation effects in
a simple-minded way. For each residue, we compute the number of side chain beads within
an approximate hemisphere above the Cβ . While we could include other heavy atoms in the
hemisphere calculation, it would needlessly increase the cost of the calculation. To handle
the uncertainty of rotameric state that can affect the count of side chain beads, the count
for different rotameric states are weighted by the prior probabilities of the rotamer states.
45
Mathematically, the count is given by
where S is sigmoid-like cutoff function. Low values of Ni are typical of residues exposed to
solvent and high values of Ni are characteristic of buried residues. A arbitrary energy curve
is coupled to the value of Ni so that
vaenv
X
Venv = i
(Ni ). (3.3)
i
While many more sophisticated solvation potentials have been derived, this term has the
advantages being very fast and easily optimized by the contrastive divergence procedure,
while remaining flexible enough to represent many of the solvation effects omitted by the
side chain potential.
The core backbone Ramachandran potential is a simple
P rama (φ , ψ ), where Virama
i Vi i i
46
n"
MD"
Energy"
t=0" update"
parameters"
t=1" " %
αi+1 − αi = ε $# dV dV
( xt=1 )− ( xt=0 )'
&
Conforma)on" dα dα
e−w right
Virama (φi , ψi ) = − log(pTCB
ai−1 ai ai+1 (φi , ψi ) + (fai ai−1 + faleft sheet
ai+1 )pai−1 ai ai+1 (φi , ψi )) (3.4)
2 i
1 right
pai−1 ai ai+1 (φi , ψi ) = (pai ai−1 (φi , ψi ) + pleft
ai ai+1 (φi , ψi )) (3.5)
2
where faleft
i ai+1
is the fraction of (ai ,ai+1 ) dimers in the PDB where the i-th amino acid
is in a β-sheet secondary structure. Note that this model uses the arithmetic average to
combine dimer probabilities rather than the geometric average recommended in the NDRD
model. Using the arithmetic average results in smoother, less sharply peaked probabilities
to encourage the side chain potential to exert more influence over the backbone geometry.
The backbone non-bonded interactions are governed by a hydrogen-bonding potential
and a steric repulsion. The hydrogen bonding potential depends on both distance and angle
for the participating atoms, and the energy of forming a hydrogen bond is a single parameter
that is chosen by contrastive divergence. The backbone atoms N, Cα , C, and Cβ feel an
repulsive interaction at approximately 1.5Å.
47
The motivation of contrastive divergence is to consider two ensembles, one closely linked
to the crystal structures and one free to diffuse away under Langevin dynamics. Under
an ideal physical model, the crystal-based ensemble and the free ensemble would be quite
similar, up to artifacts of crystallographic structure determination. For an inexact coarse-
grained model, systematic differences will arise between the crystal ensemble and simulation
ensembles. For example, the simulation ensembles may have a higher number of backbone-
backbone hydrogen bonding than are present in crystal structures. In such a case, making
the energy of forming a hydrogen bond less favorable will shift the simulation ensemble
to better resemble the crystal ensemble. Such a modification of the potential will have
additional effects; weakening the hydrogen bonding interaction may increase the amount of
hydrophobic burial. Thus, we iteratively modify all the parameters to shift the simulation
ensemble to better match the crystal ensemble. The algorithm converges when there is no
parameter that can distinguish the simulation ensemble from the crystal ensemble.
The unrestrained simulation ensemble is obtained through finite-time Langevin dynamics.
For each protein in the training set, the simulation ensemble given by 5000 time units of
dynamics, of which the first half is discarded as equilibration. This corresponds to about
ten minutes of wall-clock time for each simulation. Unless the native state is particularly
unstable, this time is insufficient to explore the conformational landscape much beyond the
native basin. Instead this typically creates a locally-equilibrated ensemble that relaxes the
crystal conformation and explores fluctuations in the near-crystal conformational basin.
The crystal ensemble is traditionally defined in contrastive divergence as consisting solely
of the true data points. This δ-function distribution is problematic for protein structures
because of the sharpness of the protein energy function. A slight inaccuracy in the crystal-
lographic reconstruction, or a slightly incorrect geometry in the Upside protein interactions,
will cause a very unfavorable energy. Furthermore, due to crystal packing and reconstruc-
tion artifacts, we would always expect some relaxation of the crystal structure under aqueous
conditions. To reduce the impact of these issues, we replace the exact ensemble of crystal
48
structures with the ensemble of running Langevin dynamics restrained to be near the crystal
structure, approximately 1Å RMSD.
To shift the free simulation ensemble toward the crystal ensemble, we change parameters
αi in proportion to the amount that they differentiate the simulation and crystal ensembles,
M
X dV dV
αi+1 = αi + (X) − (X) . (3.6)
M dαi restrained dαi free
a=1
D E D E
The quantity dV − dV
dαi (X) restrained dαi (X) free represents a pseudo-derivative of the free
energy of restraining the simulation to be near the crystal structure. In the limit that the
simulation duration is infinite, this difference is the exact derivative of the free energy. In
practice, this difference chooses a suitable direction to improve the parameters.
The derivation of contrastive divergence presented above makes the assumption that the
conformations Xa are equilibrium samples from the Boltzmann distribution of each protein,
but in reality, we must work with crystal structures of proteins. While it has been shown
that that the static diversity of crystal structures for different proteins conveys significant
information about the dynamic ensembles of individual proteins [12]. Crystal structures
deviate in a number of systematic ways from equilibrium samples, but we are most concerned
about crystal packing artifacts, crystallizability bias, and errors in published structures.
We expect that our bias in working only with crystallizable sequences, thus missing
intrinsically disordered regions from training, likely biases the resulting potential to disfavor
coil states. The loop-stabilizing effects of crystal packing somewhat counteract this effect,
as it allows longer loop regions to exist in crystal structures.
49
6 1.8
avg RMSD to native (A)
2 2
4 4
6 6
0 50 100 150 200 0 50 100 150 200
optimizer step optimizer step
Figure 3.2: Progress of contrastive divergence training. In all plots, the blue curves indicate
larger step-size training and the green plots indicate smaller step-size fine-tuning. The upper
left plot show the decline in minibatch-averaged RMSD over the course of the optimization.
The remaining plots show the convergence of the hydrogen bonding and side chain-side
chain parameters over the optimization. The larger step-size optimization of the side chain
parameters uniformly shows large oscillations that inhibit convergence.
50
3.5 Optimization
For practical reasons, the contrastive divergence simulations are run for a short time that does
not permit complete exploration of configuration space. Each contrastive divergence simu-
lation is run for 4000 time units, corresponding to approximately ten minutes of wallclock
time per iteration. The simulations use temperature replica exchange with eight replicas to
enhance barrier crossing of the contrastive divergence[20], while the temperature intervals
√
of the replicas scale with 1/ Nres to encourage good replica exchange efficiency for proteins
of various sizes. The progress of the replica exchange is monitored by the average best-fit
RMSD-to-crystal over the simulation for each minibatch.
The initial parameters from the potential come from optimizing side chain accuracy of
the model using a procedure similar to our work in the previous chapter. The contrastive
divergence training rapidly improves this model as there is a quick decline in average RMSD
over a minibatch from 5Å to 3Å. This decline is accompanied by rapid movement of the
parameters, especially the hydrogen bond strength. At the same time, the side chain param-
eters show much greater fluctuations. This is likely because there are far fewer interactions
for any particular pair of residue types (say ALA-GLY) than there are hydrogen bond inter-
actions. For this reason, we expect the gradients of the side chain-side chain parameters to
be far noisier than those of the hydrogen bond interactions. To reduce the fluctuations and
fine-tune the results, we reduce the optimizer step size by a factor of four after two epochs.
While the change in the observed RMSD has become small, there are indications that we
have not converged the value of the parameters and thus are stopping the contrastive di-
vergence early. Earlier tests showed that continuing the contrastive divergence to complete
convergence does not necessarily produce better results, which has been observed in [5]. Par-
ticulary, if large barriers have been built around the native states by contrastive divergence,
there may be little relaxation of the conformation during the short simulations of contrastive
divergence and hence little useful information to optimize the parameters. Instead, the fur-
ther fine-tuning of the contrastive divergence results may reduce the accuracy of the model
51
by further optimizing against only very near-native results. Secondly, early stopping of opti-
mization has been observed in a number of contexts to function as a regularizer that favors
simpler models[6].
There is a curious behavior of the hydrogen bond strength, where it appears to converge
to a significantly smaller value during the fine-tuning than during the larger optimizer steps.
We speculate that the extra noise in the side chain interactions during the larger optimizer
steps may in aggregate cause stronger side chain interactions for the protein. This would
necessitate a large hydrogen bond energy to balance against the side chain interactions.
The previous training only optimized the parameters’ ability to hold a protein conformation
for a short period of time, less than ten minutes of wallclock time. While contrastive di-
vergence training has been shown to train models well for many machine learning problems
[4], the accuracy must be demonstrated for our model. As a stringent test of the protein
model and our training procedure, we attempt de novo folding of a benchmark set of small,
fast-folding proteins similar to those used in [15, 1]. Before training, we removed all proteins
from the training set that were homologous to any protein in the benchmark set to ensure
that this would be a de novo prediction. The proteins are simulated using replica exchange
using 16 replicas for approximately three days of wall-clock time with one processor per
replica.
Two replica exchange simulations are launched for each protein with the lowest tem-
perature of the replicas the same as the contrastive divergence temperature. The first is
initialized from the native configuration of the protein to assess the stability of the exper-
imental structure. The second simulation is initialized from a random unfolded state with
φ and ψ angles chosen uniformly at random. The range of temperatures were chosen to be
large enough to span to signficant populations of unfolded states for all proteins.
We judge the accuracy and equilibration of the model from the histogram of best-fit
52
alpha3d BBA BBL homeodomain
0.4 1.5
2.0 0.20
0.4
1.0 1.5
0.4
0.3
density
1.0
0.2
0.5 0.2
0.1 0.5
Figure 3.3: RMSD distributions after equilibration phase. Green indicates simulations
started from the crystallographic native structure and red indicates simulations started from
a random unfolded state.
53
Name Length Lowest RMSD (Å)
alpha3d 73 2.0
BBA 28 0.7
BBL 47 1.7
homeodomain 52 1.5
lambda 80 4.3
NTL9 39 2.6
protein B 53 1.7
protein G 56 3.7
NuG2 (Shaw) 57 0.9
protein L 61 2.6
ubiquitin 76 2.0
WW domain 33 0.9
Table 3.1: Lowest RMSD for de novo folding simulations.
RMSD deviations from the native structure after discarding the initial third of the simulation
as equilibration. When the native-initialized and unfolded-initialized structures have similar
RMSD distributions, the simulation has likely converged. In protein L and ubiquitin, we note
that the ensembles are relatively far from convergence as the native- and unfolded-initialized
simulations disagree strongly in their RMSD distributions.
It should be noted that, while no parameters are set based on observing the benchmark
folding results, decisions on the functional form of energies are made based on the effect
on folding results. While we expect that any statistical bias toward higher accuracies that
arises from making force-field decisions after seeing folding results is small, we do note these
“researcher degrees of freedom”[22] for full disclosure.
Multiple conformations are observed for many of the simulated proteins. This is con-
sistent with the known properties of maximum likelihood training, where inability of the
functional form of the model to represent the true Boltzmann distribution of the model
will result in smaller, broader energy functions and thus unsurprising to see multiple con-
formations in the ensemble. Since contrastive divergence is an approximation to maximum
likelihood, it is unsurprising that it inherits this conservative property.
54
3.7 Structural characterization of low temperature
conformations
To characterize the structural ensemble at low temperature, we cluster each trajectory into
five clusters and choose a conformation in a high density region of each cluster to represent
the cluster, as detailed in section 3.15.
The majority of the proteins, excluding BBL, show a small number of well-defined and
stable basins that represent the dominant conformations of Upside for each protein. These
well-defined clusters are clearly indicated in the principal component plots in Fig. 3.12. While
the simulation often produces many of the these conformations quickly, the equilibration of
their populations takes time, likely dominating the relaxation time of the RMSD distribu-
tions. Still, this relaxation time is on the order of hours to days of wallclock time, making it
extremely quick in comparison to typical molecular dynamics simulations.
As indicated by the clustered structures, the Upside simulations tend to correctly repre-
sent the secondary structure of the proteins even as it provides a small number of distinct
tertiary arrangements. This tertiary diversity is illustrated in the mirror three helix bundles
for α3d and protein B, as well as the subtle re-arrangements in NuG2. As these structures
coexist with similar probabilities at low temperature, we hypothesize that the short-time
contrastive divergence we are using does not provide a sufficient library of large changes in
the tertiary structure to enable the potential to properly distinguish the various conforma-
tions. It is difficult for short-time simulation to produce sweeping changes that preserve
secondary structure but cause permutations of the tertiary structure.
As these results are obtained with an almost untuned backbone energy fucntion (only
a single scalar parameter to control the amount of sheet in the Ramachandran potential),
it is likely that loop conformations would be significantly improved with a more carefully
tuned backbone potential. Presumably the backbone potential can be optimized using the
same contrastive divergence procedure as we have used in this work. It remains only to
55
alpha3d BBA BBL homeodomain
N 1 N 1 N 1 N 1
2 3 2 3 2 3 2 3
4 5 4 5 4 5 4 5
2 3 2 3 2 3 2 3
4 5 4 5 4 5 4 5
2 3 2 3 2 3 2 3
4 5 4 5 4 5 4 5
Figure 3.4: Structures of the native state (N) and a representative structure within each
cluster of the de novo trajectory in the lowest temperature replica. The clusters are ordered
by average RMSD-to-native within the cluster. Representative are chosen from regions of
high density in the principal component analysis plots of the trajectory.
56
Figure 3.5: RMSD trajectories within each cluster, indicated by color. Clusters are ordered
by the average RMSD-to-native. Within each cluster, the data points are in trajectory order.
choose a suitable backbone model, such as TorusDBN[3], that allows sufficient freedom for
the optimization. In future work, we intend to co-adapt an existing backbone statistical
potential to the Upside energy as part of the contrastive divergence procedure.
The point of the contrastive divergence training is to capture the energy of fluctuations
about the native state of the protein. As we have shown above, the energies learned by
contrastive divergence are sufficient to assign significant probability to the native state at
low temperatures. We are also interested in the melting and folding behavior of the model,
specifically whether the excited states of the model are consistent with experimental data.
The model typically exhibits concerted melting behavior over a small range of tempera-
tures. While the temperature of the model in Upside is not exactly comparable to a physical
temperature, it is reasonable to assume T = 1 corresponds roughly to a temperature of 300 K.
The ubiquitin transition occurs over a temperature range of approximately 0.07 temperature
57
Figure 3.6: Time vs RMSD at the temperature of peak heat capacity. Heat capacity for each
temperature is estimated by the fluctuation formula Cp = (var E)/T 2 .
Figure 3.7: Fraction of formed HBonds vs Rg at the temperature of peak heat capacity
58
alpha3d BBA BBL homeodomain
25 20.0
14 20
17.5
20 12
15.0 15
10
15 12.5
10
280 300 320 340 360 280 300 320 340 360 280 300 320 340 360 280 300 320 340 360
lambda NTL9 NuG2 (Shaw) protein B
25
30 18
16
25 20 20
14
Rg
20
12 15 15
15 10
10 10
280 300 320 340 360 280 300 320 340 360 280 300 320 340 360 280 300 320 340 360
protein G protein L ubiquitin WW domain
25 17
25 15
20 20
20 13
15 15
15 11
10 9
280 300 320 340 360 280 300 320 340 360 280 300 320 340 360 280 300 320 340 360
310 * T
0.4
0.3
0.2
0.1
0.0
0.4
density
0.3
0.2
0.1
0.0
0.4
0.3
0.2
0.1
0.0
−2 0 2 4 −2 0 2 4 −2 0 2 4 −2 0 2 4
energy/N_res
59
alpha3d BBA BBL homeodomain
100
75
50
25
100
75
Cp/N_res
50
25
100
75
50
25
280 300 320 340 360 280 300 320 340 360 280 300 320 340 360 280 300 320 340 360
310 * T
Figure 3.10: Heat capacity as a function of temperature. Heat capacity for each temperature
is estimated by the fluctuation formula Cp = (var E)/T 2 .
60
Figure 3.11: Constant temperature reversible folding trajectory of ubiquitin at a tempera-
ture of 1.003. Note that the structure becomes completely extended at high RMSD before
returning to a structure with an RMSD-to-native of 2.3Å.
We are able to show constant temperature, reversible folding to the experimental native state
for a number of proteins in our test set. This is illustrated in Figures 3.6 and 3.11. Each of
the simulations are run on a single core for a few days. The time scales of folding implied by
these reversible folding trajectories implies that the time scales of the contrastive divergence
simulations are far less (often a factor of 100 or more) than required to equilibrate these
proteins. It is clear that the contrastive divergence is not optimizing over all conformations
of the system, but only local fluctuations. The reversible folding at constant temperature
also indicates that we have an model to explore folding pathways in reversible conditions.
The power of our optimization procedure also means that we can take a number of qualitative
models and make them quantitative by optimizing the parameters for de novo folding, so
that we can explore folding pathways for a number of reasonable folding models without
having to resort to strong native bias.
We have a model that produces a wide range of protein-like behavior for folding, but
it is not yet clear which parts of the model are associated with which behavior, so that
we lack a truly reduced description of the protein physics that identifies certain features as
key (I could argue that this is not possible since statistical correlations tend to confound
61
mechanistic description in ways that practitioners do not appreciate). Given that these
features are produced without any specific attempt to generate them suggests that we are
approaching a good model of protein dynamics. Furthermore, contrastive divergence, though
its relationship to maximum likelihood, is tuned to pick up fluctuation scales in addition to
the best structure. Emprically, we have noticed that a large increase in cooperativity is
observed by going to optimized, precise side chain locations for the side chain model, as
show in the previous chapter. This suggests that we are able to predict the improvement in
the behavior of the protein physics by observing an increase in side chain prediction accuracy,
which is a much easier criterion to assess. Future work will address the quality of predicted
pathways and observables for individual proteins.
Note that conditional on low hydrogen bonding, the Rg at high temperature and at the
peak heat capacity are quite similar. This suggests the increase in Rg for the unfolded state
as temperature increases is driven by a reduction in backbone-backbone hydrogen bonds
rather than side chain effects.
Based on the above results, there are two facts that must be reconciled. The first fact
is the sharp phase transition with a single peak for the heat capacity. The shape of the
phase transition, but not its amplitude, is consistent with a cooperative folding transition.
Additionally, the relationship of average Rg to average number of hydrogen bonds is ap-
proximately linear over the transition region. The second fact is the large residual hydrogen
bonding in the denatured state at the heat capacity peak, which suggests that the transition
is quite non-cooperative. We propose that this may be explained by the essential feature
of the contrastive divergence process, that it must balance the competing energy terms of
the model so that no one energy dominates. We suspect that the very close temperatures
of disrupting tertiary contacts by leaving the native state and then melting hydrogen bonds
results from balancing the interaction energies of side chains pair interactions and backbone-
backbone hydrogen bonds. A small tweak to the contrastive divergence training may be
able to push the temperature of melting secondary structure lower so that the folding is
62
significantly more cooperative.
The most similar work to ours is a contrastive divergence optimization of a Gō-like protein
potentials sampled with crankshaft Monte Carlo moves[19, 27]. These works optimized only
tens of parameters using contrastive divergence, and the resulting model is only used to fold
protein G and 16-residue peptides.
Early work focused on training protein energy terms against a library of decoys. Such
efforts are problematic for a number of reasons. The first is realistic molecular dynamics
energy functions have extremely rugged energy landscapes, so that the energy of a decoy
may be much higher than another structure less than an angstrom away in RMSD. This
ruggedness means that it is improper to score decoy prediction by energy without first relax-
ing the decoys to the center of the nearby energy well. The significance of using unrelaxed
decoys depends on how close the decoy generation model is to the force field being trained.
Another difficulty is that the conformational space of proteins is vast; where it not so, finding
the lowest energy conformation for a given energy model would be easy. As a result, decoy
sets are unable to exhaust conformational space and are likely to miss some conformations
that incorrect protein energy functions find extremely favorable. These effects suggest that
the best decoy set may be obtained simply by conformational sampling of the protein en-
ergy function. Decoy generation may provide convenient seed structures for exploration, but
ultimately protein conformational sampling is needed to evaluate potential energy functions
and suggest directions for improvement.
A more technical distinction against traditional protein training methods, such as Z-
score optimization [16], is that it is relatively unimportant to know how well the model
handles average decoys. The important task is to produce a low energy ensemble (those
structures no more than a few kB T above the relaxed experimental structure) having a high
population of low RMSD structures. While improving Z-score should correlate with a better
63
conformational ensemble early in training, in later training it will likely become important to
focus on the remaining conformations that are lower in energy than the native, irrespective of
the effect on Z-score. Finally, we note that Z-score is inherently defined relative to a decoy
ensemble (it is near meaningless with respect to the Boltzmann ensemble of the model).
The Z-score is not defined comparing the native to all other conformations, as denatured
conformations are exponentially more numerous than folded conformations. Instead, one
must define a class of decoy structures that are somehow folded-protein-like in order to
define the Z-score. Methods based on simulations ensembles and the associated probability
density (such as maximum likelihood and contrastive divergence) are well-defined and do
not need to create a class of decoys.
There has been previous work on estimating protein statistical potentials using simu-
lation trajectories. Following the long history of work for using contrastive divergence on
machine learning models, [18] have applied contrastive divergence to few-parameter protein
models. The present work extends these methods to handle large-scale molecular dynamics
training, where the scale is large in both the number of parameters and the amount of pro-
tein simulation training used. Furthermore, we go beyond existing work by demonstrating
that our method is able to fit a model transferable to realistic unfolded states and thermal
melting behavior, consistent with a cooperative, pseudo-first-order model of protein folding.
There have been several attempts to train a force-field usingmaximum likelihood. These
methods are inherently limited by the need to compute the derivative of the free energy, which
involves a summation over an equilibrium sample of the configurations. Such a requirement
necessitates a very long simulation to update parameters. Still, the maximum likelihood
approach can be viable when used with very small proteins on which the simulations converge
quickly. A variant of maximum likelihood is given in [31], where decoys are generated and a
maximum likelihood model is fit to adjust the parameters to distinguish between near-native
and far-from-native conformations (smoothed with a Gaussian cutoff of nearness). The
potential is trained on a single protein, tryphtophan cage, and then the resulting potential
64
is applied to a number of α-helical proteins.
3.11 Conclusions
These data suggest that the Upside model may be representing the physics of protein folding
well enough to be useful far from its training temperature. In particular, future work will
investigate if the folding pathways of the Upside model are consistent with experimental
data.
We have shown that extremely short simulations, paired with optimization of contrastive
divergence, are capable of parameterizing an accurate coarse-grained potential. Replica
exchange molecular dynamics with this energy is capable of folding many small, fast-folding
proteins to sub-3Å accuracy in a matter of days on a single machine. Training the model
by contrastive divergence is successful precisely because the Upside model has extremely
quick equilibration for many proteins. The usual trade-off in molecular simulation is to
choose between accurate, expensive simulation models (such as explicit water molecular
dynamics) and inexpensive, less accurate models. With contrastive divergence training of
the model, the situation is reversed to some extent. Less expensive models allow more
extensive exploration during the training phase, which allows the contrastive divergence
better maximize the accuracy of the chosen parameters. Coarse-grained models must still
retain sufficient flexibility to represent the underlying physics, but faster models will come
closer to their ideal parameters due to the enhanced exploration. Simultaneously, we have
show that very large numbers of parameters (approximately 25000 in our case) are no obstacle
to producing accurate proteins models using trajectory-based training. While overfitting is
still a real concern, the severity is greatly reduced because contrastive divergence is training
against the vast possibilities of alternative protein conformations explored by conformational
sampling. Overstrengthing any particular interaction will lead to that interaction being
present in a large number of the structures found by molecular simulations, which will cause
negative feedback in the training that reduces the magnitude of that parameter. By this
65
mechanism, contrastive divergence automatically obtains balanced parameters such that no
particular interaction overwhelms the others. Even if we have clear experimental evidence
about the strength of a few interactions (such as backbone-backbone hydrogen bonding),
it is inadvisable to force those known interactions to their experimental values. It is more
important for the different interaction terms to have balanced strength than to have precise
accuracy for a subset of the model.
We have also clearly shown that straightforward contrastive divergence training produces
a protein model with both cooperative folding behavior and highly-expanded unfolded states.
These correlations to experiment are notable precisely because the contrastive divergence op-
timizations is only concerned with native state stability, not the properties of the unfolded
state or the folding transition. Instead, these properties emerge from the training sponta-
neously, and do not depend on an accurate solvation model. Based on these successes, it is
likely that this methodology can be easily extended to treat disordered regions of proteins,
possibly after augmenting the training set with examples of well-characterized disordered
systems.
Despite the success of this training procedures, there are also a number of weaknesses that
will be addressed in future work. In the machine learning literature, contrastive divergence
is noted to obtain peak accuracy before fully converging then declining accuracy afterwards
[20]. This is attributed to growing interaction strengths as training proceeds that result
in larger barriers that reduce exploration of the model. With reduced exploration, the
contrastive divergence trains against less interesting alternative structures. There have been
many suggestions in the literature and many point to continuous simulations that do not
reset to the native structure on every iteration, termed Persistent Contrastive Divergence[25].
We intend to explore such alternatives with larger computing resources to better maximize
the accuracy that we can extract from the Upside model.
Future work will also attempt to address equilibration difficulties for α/β-proteins. These
difficulties are typified by the underconverged simulations of protein L and ubiquitin in the
66
benchmark simulations. More expensive training like persistent contrastive divergence may
still be unable to train optimal parameters when simulations like those of protein L are un-
coverged after three days of continuous simulation. The origin of the sampling difficulty may
be traced to the sharp folding transition in these proteins. Temperature-based techniques
for accelerating simulation convergence, such as replica exchange, are noted to have diffi-
culties accelerating first-order phase transitions [8]. The essential reason is that few to no
replicas have any significant population in the transition state energies between folded and
unfolded conformations. As first-order, cooperative folding transitions are experimentally
well-supported [13], we would like to preserve this emergent property of our model rather
than modifying the training to somehow avoid it. Instead, we intend to pursue alternative
sampling techniques, such as Wang-Landau sampling [29, 28], that are known to gracefully
handle first-order phase transitions. With accelerated sampling, we may expect improve
accuracy from contrastive divergence.
The force is integrated using Verlet integration with a time step of 0.009 time units. Tem-
perature is maintained using a Langevin thermostat with a thermalization timescale of 0.135
time units.
The Adam optimizer is used to perform gradient descent on the objective function, using
the contrastive divergence pseudo-gradient in place of the true maximum likelihood gradient.
The Adam parameters used are β1 = 0.8, β2 = 0.96 and = 10−6 . The α parameter is varied
based on the type of term to ensure stability, αSC = 0.5, αenv = 0.1, αHBond = 0.02, and
αsheet = 0.03. The α parameters are multiplied by 0.25 for the fine-tuning optimization.
Regularization and derivative propagation for contrastive divergence optimization are
handled using the Theano library[24].
67
3.13 Training data and optimization
The contrastive divergence training is conducted with 456 crystal structures from the Protein
Data Bank. The initial selection of structures uses the PISCES server[30] to select proteins
with X-ray resolution less than 2.2Å and pairwise sequence similarity less than 30%. In
structures with multiple chains, a single chain is chosen by the PISCES server. To avoid
non-globular proteins or proteins with strong interactions with other subunits in the struc-
ture, random sample consensus linear regression [7] is used to identify outliers based on the
relationship between log Nres and log Rg . Only chains with between 50 and 100 residues are
used to encourage fast relaxation during the contrastive divergence simulations. All proteins
homologous to proteins in the benchmark folding set are eliminated from the training set.
Additionally, all proteins with backbone gaps, either missing residues due to diffuse electron
density or non-standard amino acids that Upside does not handle, are also excluded from
the training set.
The final training set of 456 proteins is divided into 38 groups of 12 proteins each. These
groups, called minibatches, called minibatches are an essential feature of algorithms derived
from stochastic gradient descent. The idea is that it is very wasteful to compute the gradient
direction in equation (3.6) using all 456 proteins. Using all of the proteins would give an
extremely precise estimate of the gradient that nonetheless could only be used to make
a small change in the objective because of the non-linearity of the Boltzmann ensemble
to changing parameters. Instead, it is better to use a few proteins at a time to obtain a
inexpensive, noisy estimate of the gradient then take a small step in that direction. As
long as the parameter step sizes decrease appropriately, the noise of the minibatch gradients
will average out and the iteration will converge. To avoid overfitting, the minibatches are
cycled through sequentially so that each minibatch is only re-used every 38 steps (a full pass
through the minibatches is termed an epoch). We use the Adam optimizer[14] to accelerate
convergence of the parameters.
68
Figure 3.12: Scatter plot of the first two principal components for the lowest temperature
replica of each trajectory, after discarding the first 1/3 of the simulation. The cluster of each
conformation is indicated by colors, using the same color scheme as Fig. 3.5, and the cluster
representative is indicated by a plus sign.
The test proteins are the same as those described in Section 2.12.
Figures 3.4 and 3.5 require a clustering of the low temperature structures as well the choice
of a representative conformation for each cluster.
To cluster each trajectory, the conformations are mapped to their Cα contact matrices,
defined so that the ij entry of the contact matrix is 1 if the position of the residue i Cα
is within 8Åof the residue j Cα . The contact matrix is unrolled into a long vector, so
that each conformation is represented by a 0/1-vector of length Nres (Nres+1 )/2. Principal
component analysis is used to make these long vectors to five dimensions, scaled by the
relative size of their variance components so that. This five-dimensional representation
typically shows well-defined clusters for the trajectories we study. The resulting principal
69
components representation of the conformations is partitioned into five clusters using the
k-means++[2] implemenation in scikit-learn[17]. The choice of five clusters is arbitrary and
may not be ideal to represent a natural clustering for any particular protein. The clustering
is merely intended to roughly map the low temperature structures.
Representative conformations are chosen for each cluster as regions of high density in the
first two principle components. For each cluster, the smoothed probability density in the
first two principle components is estimated using a kernel density estimate with bandwidth
given by 20% of the standard deviation in the cluster. The representative conformation for
the cluster is chosen to be the conformation with the high probability density of its first two
principle components according to this measure.
represents an arbitrary label to enumerate the proteins and Xa represents the configuration
of the protein (in our case, we are only interested in the backbone trace for Xa ). Note that
the “true” Boltzmann distribution is an unobservable idealization of the conformational
ensemble of a protein under physiological conditions, and we further idealize that the true
Boltzmann distribution is derived from from an extremely-complicated true potential Vstrue
a
by statistical mechanics,
exp(−Vstrue (Xa ))
ptrue
sa (Xa ) = a
true (3.7)
exp(−Gsa )
Z
true
Gsa = − log e−Vsa (X) dX.
true (3.8)
70
The subscript sa indicates that both the potential Vstrue
a
and free energy Gtrue
sa depend
on the sequence of the protein. We may think of this as an artifact of working in the
coarse-grained coordinates of the backbone trace, where the energy Vstrue
a
really represent
the free energy of the backbone coordinates after integrating away the solvent and side chain
degrees of freedom. An analogous situation occurs in parameterizing all-atom molecular
dynamics, where the “energy” of the system really represents the free energy of the system
after integrating over the electronic degrees of freedom. Our goal is to define a parametric
approx
Vs (X) that approximates the Vstrue for any sequence s. We drop the subscript s below
where there is no chance of confusion.
For an approximate potential V approx , such as the Upside model defined above, it is
almost certain that V approx does not have enough flexibility in its functional form to match
all of the Boltzmann distributions pa for any sequence sa . We must instead find a V approx
that is “close” to V approx . Defining the Boltzmann distribution of V approx in the same
manner as that of V true ,
approx
approx exp(−Vsa (Xa ))
psa (Xa ) = approx (3.9)
exp(−Gsa )
Z
approx approx
Gsa = − log e−Vsa (X) dX, (3.10)
KL(ptrue , papprox )
ptrue (X)
Z
= ptrue (X) log approx dX
p (X)
= h− log papprox (X) + log ptrue (X)itrue
71
In the last equation, we note that the KL divergence is simply the average energy difference
between the true and approximate potentials (after subtracting the free energies to normalize
the probabilities), where the average is taken over the true Boltzmann distribution. This
makes the KL divergence highly non-symmetric between the true and approximate potentials.
If the approximate potential is high (unfavorable) where the true potential is low, this will
make the KL divergence much larger. Regions of configuration space where the true potential
is high contribute little to the average since these regions will have low probability in the true
Boltzmann ensemble. The key fact when minimizing KL divergence is that if the approximate
distribution lacks the freedom to exactly match the true distribution, then the minimizing
distribution will be weaker than the true distribution (i.e. less sharp) to avoid assigning
highly unfavorable energy to configurations that are likely in the true distribution.
We unfortunately lack knowledge of the true energy ptrue , so that we are unable to
compute the expectation needed for the KL divergence for a concrete V approx . Instead of
minimizing the KL divergence, we can instead minimize
since the remaining term hV true (Xa ) − Gtrue itrue is independent of the approximating po-
tential. This expectation value is still intractable since we do not know ptrue . We instead
approximate
M
1 X
ptrue (X) ≈ pempirical (X) = δ(X − Xa ), (3.13)
M
a=1
where δ is the Dirac delta function and M is the number of proteins. We can approximate
72
minimizing the KL divergence by instead minimizing
Minimize the expression (3.14) is exactly the method of maximum likelihood. The derivation
given above illustrates two points via the connection to KL divergences. The first is what
happens when the approximating energy V true cannot capture the nuances of V approx . In
this case, the model will be overly broad so as not to assign high energy to any configurations
that have low energy under V true . The second salient point is that with only a finite number
of samples, pempirical may be a poor approximation to ptrue . Whether that is the case
depends on the ability of V approx to wrap itself tightly near the δ-functions associated with
each sample. This is the origin of overfitting, which is especially problematic with large
numbers of parameters.
We can now take the derivative with respect to an arbitrary forcefield parameter αi in
preparation to perform gradient descent to minimize (3.14). This gradient is given by
M
d 1 X approx
(V (Xa ) − Gapprox )
dαi M
a=1
M
1 X dV approx dGapprox
= (Xa ) −
M dαi dαi
a=1
M
!
dV approx dV approx
1 X
= (Xa ) − (X) , (3.15)
M dαi dαi approx
a=1
where we have used the standard statistical mechanics identity dG/dαi = hdV /dαi i. While
we have obtained a concrete expression for gradient descent in (3.15), we still have a ma-
jor stumbling block. Computing the expectation of the derivative of the potential at Xa is
73
straightforword given a functional form for V approx , but obtaining even a reliable approxi-
D approx E
mation for dV dα (X) is extraordinarily difficult. To approximate the expectation
i approx
value with a finite sample, we would need to obtain Boltzmann samples from our current
approximating potential. Even obtaining the single most likely configuration for our ap-
proximating potential is equivalent to finding the native state of the model, and this is very
difficult for realistic pairwise potentials. Making matters more difficult, we would need to
find the Boltzmann ensemble for all the proteins in our training set, and keep those Boltz-
mann ensembles up to date as we use gradient descent to optimize our potentials. This
represents an extreme expense and it is unrealistic to obtain good converged estimates for
D approx E
dV (X) over the whole training set for anything but the simplest models of
dα
i approx
proteins. Note also that we cannot simply construct a large list of structures at some time
and reweight those structures according to the potential, since the potential is constantly
changing. Reweighting ensembles is only valid over very small neighborhoods of parameter
space, and this procedure would depend on being able to generate an exhaustive survey of
candidate structures FOOTNOTE there are contrastive divergence variants that attempt
such a thing but they really work based on the driving effect of the changing potential as
seen in [25].
The contrastive divergence method[10] works based on a key insight. We do not need an
accurate approximation to (3.15), so long as the derivative points in direction of parameter
space that improves the potential accuracy (i.e. any direction is acceptable as long as it is
D approx E
not uphill). The authors propose replacing Boltzmann average dV dα (X) with
i approx
a finite-time Fokker-Planck average over a very short period of time for a simulation that
originates at the data point Xa . In the Monte Carlo (MC) dynamics that the authors use,
even one MC step is sufficient to produce decent optimization of the model. In our case, we
replace their small number of MC steps with a short time simulation using replica exchange
Langevin dynamics. As the duration of the simulation is increased, our derivative estimate
will converge to the true derivative (3.15).
74
3.17 References
[1] Aashish N Adhikari, Karl F Freed, and Tobin R Sosnick. De novo prediction of protein
folding pathways and structure using the principle of sequential stabilization. Proceed-
ings of the National Academy of Sciences, 109(43):17442–17447, 2012.
[2] David Arthur and Sergei Vassilvitskii. k-means++: The advantages of careful seeding.
In Proceedings of the eighteenth annual ACM-SIAM symposium on Discrete algorithms,
pages 1027–1035. Society for Industrial and Applied Mathematics, 2007.
[3] Wouter Boomsma, Kanti V Mardia, Charles C Taylor, Jesper Ferkinghoff-Borg, Anders
Krogh, and Thomas Hamelryck. A generative, probabilistic model of local protein
structure. Proceedings of the National Academy of Sciences, 105(26):8932–8937, 2008.
[5] Guillaume Desjardins, Aaron Courville, Yoshua Bengio, Pascal Vincent, and Olivier De-
lalleau. Parallel tempering for training of restricted boltzmann machines. In Proceedings
of the Thirteenth International Conference on Artificial Intelligence and Statistics, pages
145–152. MIT Press Cambridge, MA, 2010.
[6] David Duvenaud, Dougal Maclaurin, and Ryan P Adams. Early stopping as nonpara-
metric variational inference. In Proceedings of the 19th International Conference on
Artificial Intelligence and Statistics, pages 1070–1077, 2016.
[7] Martin A Fischler and Robert C Bolles. Random sample consensus: a paradigm for
model fitting with applications to image analysis and automated cartography. Commu-
nications of the ACM, 24(6):381–395, 1981.
[8] Jes Frellsen, Ole Winther, Zoubin Ghahramani, and Jesper Ferkinghoff-Borg. Bayesian
75
generalised ensemble markov chain monte carlo. In Proceedings of the 19th International
Conference on Artificial Intelligence and Statistics, pages 408–416, 2016.
[9] Thomas Hamelryck, Mikael Borg, Martin Paluszewski, Jonas Paulsen, Jes Frellsen,
Christian Andreetta, Wouter Boomsma, Sandro Bottaro, and Jesper Ferkinghoff-Borg.
Potentials of mean force for protein structure prediction vindicated, formalized and
generalized. PloS one, 5(11):e13714, 2010.
[11] Malene Ringkjøbing Jensen and Martin Blackledge. Testing the validity of ensemble
descriptions of intrinsically disordered proteins. Proceedings of the National Academy
of Sciences, 111(16):E1557–E1558, 2014.
[12] Abhishek K Jha, Andrés Colubri, Karl F Freed, and Tobin R Sosnick. Statistical coil
model of the unfolded state: resolving the reconciliation problem. Proceedings of the
National Academy of Sciences of the United States of America, 102(37):13099–13104,
2005.
[13] Hüseyin Kaya and Hue Sun Chan. Polymer principles of protein calorimetric two-state
cooperativity. Proteins: Structure, Function, and Bioinformatics, 40(4):637–661, 2000.
[14] Diederik Kingma and Jimmy Ba. Adam: A method for stochastic optimization. arXiv
preprint arXiv:1412.6980, 2014.
[15] Kresten Lindorff-Larsen, Stefano Piana, Ron O Dror, and David E Shaw. How fast-
folding proteins fold. Science, 334(6055):517–520, 2011.
[16] Adam Liwo, Matthew R Pincus, Ryszard J Wawak, Shelly Rackovsky, S Oldziej, and
Harold A Scheraga. A united-residue force field for off-lattice protein-structure simula-
tions. ii. parameterization of short-range interactions and determination of weights of
76
energy terms by z-score optimization. Journal of computational chemistry, 18(7):874–
887, 1997.
[18] Alexei A Podtelezhnikov, Zoubin Ghahramani, and David L Wild. Learning about
protein hydrogen bonding by minimizing contrastive divergence. Proteins: Structure,
Function, and Bioinformatics, 66(3):588–599, 2007.
[19] Alexei A Podtelezhnikov and David L Wild. Inferring knowledge based potentials using
contrastive divergence. In Bayesian Methods in Structural Bioinformatics, pages 135–
155. Springer, 2012.
[20] Ruslan R Salakhutdinov. Learning in markov random fields using tempered transitions.
In Advances in neural information processing systems, pages 1598–1606, 2009.
[21] Min-yi Shen and Andrej Sali. Statistical potential for assessment and prediction of
protein structures. Protein science, 15(11):2507–2524, 2006.
[22] Joseph P Simmons, Leif D Nelson, and Uri Simonsohn. False-positive psychology undis-
closed flexibility in data collection and analysis allows presenting anything as significant.
Psychological science, page 0956797611417632, 2011.
[23] John J Skinner, Wookyung Yu, Elizabeth K Gichana, Michael C Baxa, James R Hin-
shaw, Karl F Freed, and Tobin R Sosnick. Benchmarking all-atom simulations using
hydrogen exchange. Proceedings of the National Academy of Sciences, 111(45):15975–
15980, 2014.
77
[24] Theano Development Team. Theano: A Python framework for fast computation of
mathematical expressions. arXiv e-prints, abs/1605.02688, May 2016.
[25] Tijmen Tieleman. Training restricted boltzmann machines using approximations to the
likelihood gradient. In Proceedings of the 25th international conference on Machine
learning, pages 1064–1071. ACM, 2008.
[26] Daniel Ting, Guoli Wang, Maxim Shapovalov, Rajib Mitra, Michael I Jordan, and
Roland L Dunbrack Jr. Neighbor-dependent ramachandran probability distributions of
amino acids developed from a hierarchical dirichlet process model. PLoS Comput Biol,
6(4):e1000763, 2010.
[27] Csilla Várnai, Nikolas S Burkoff, and David L Wild. Efficient parameter estimation
of generalizable coarse-grained protein force fields using contrastive divergence: a max-
imum likelihood approach. Journal of chemical theory and computation, 9(12):5718–
5733, 2013.
[28] Fugao Wang and DP Landau. Determining the density of states for classical statistical
models: A random walk algorithm to produce a flat histogram. Physical Review E,
64(5):056101, 2001.
[29] Fugao Wang and DP Landau. Efficient, multiple-range random walk algorithm to cal-
culate the density of states. Physical review letters, 86(10):2050, 2001.
[30] Guoli Wang and Roland L Dunbrack. Pisces: a protein sequence culling server. Bioin-
formatics, 19(12):1589–1591, 2003.
[31] Bartlomiej Zaborowski, Dawid Jagiela, Cezary Czaplewski, Anna Halabis, Agnieszka
Lewandowska, Wioletta Z?mudzin?ska, Stanislaw Oldziej, Agnieszka Karczyn?ska,
Christian Omieczynski, Tomasz Wirecki, et al. A maximum-likelihood approach to
force-field calibration. Journal of chemical information and modeling, 55(9):2050–2070,
2015.
78
CHAPTER 4
PHYSICS
4.1 Introduction
The purpose of this chapter is to give a detailed account of the Upside model for protein
physics and to explain the myriad aspects of the potential energy function. Modelling choices
are made both for physical realism and for the ability to stably integrate Hamilton’s equa-
tions using the Verlet algorithm. Detail-averse readers may skip this chapter without com-
promising their understanding of the rest of the text, although they are encouraged to read
subsection 4.3 for a description of interesting effect when converting between Ramachandran
frequencies and Ramachandran potentials.
While this chapter does not describe the software implementation of the Upside model
in detail, descriptions of the components of the Upside potential will include a reference to
the class in the Upside program that implements them.
Harmonic potentials are used to restrain the N–Cα , Cα –C, and C–N distances to physical
values. Upside uses the standard form of a harmonic potential,
1
Vbond (x1 , x2 ) = k(|x1 − x2 | − r0 )2 , (4.1)
2
where the equilibrium distance r0 is chosen to be 1.453Å, 1.526Å, and 1.300Åfor the three
types of bond springs, respectively. The spring constant k is chosen to be 48. energy units
/ Å2 . While the bond distances are based on crystallographic data, the choice of spring
79
constant is made in an attempt to maximize the allowable Verlet time step. Since bond
vibrations are often the highest frequencies in a molecular system, the step size for molecular
simulation is often chosen to avoid instability of the Verlet integrator due to instabilities in
integrating the motion of bonded atoms. Hence, simulation efficiency may be increased by
increasing the integrator time step after reducing the spring constant for bonded atoms. A
limit on the extent to which the bond spring constant may be reduced is discussed in the
next section.
Bond springs are implemented in the DistSpring class.
Angle springs are used to restrain the bond angles for N–Cα –C, Cα –C–N, and C–N–Cα
groups. The form of the potential is
1 2
1 − x2 ) · (x3 − x1 ) − a0 ) .
Vangle (x1 , x2 , x3 ) = k((x\ \ (4.2)
2
Note that the dot product gives cos θ where θ is the bond angle. The integration stability
of using the more standard θ instead cos θ to be investigated, as angular springs can cause
integrator instability when the spring constant k is high (in Upside, k is chosen to be 175.).
Note that this angular term limits the allowable spring constant for the distance springs at a
fixed integration time step. When one of the flanking atoms in an angular spring approaches
the central atom, the force due to the angular potential increases proportional to 1/r. Since
these high forces are likely to push the system outside the stability conditions of the Verlet
integrator, a sufficiently high spring constant must be used for the distance restraints to
avoid near approach of atoms participating in an angular interaction.
80
The equilibrium value a0 is given by
α
aN–C –C = cos 109.5◦ (4.3)
0
α
aC –C–N = cos 120◦ (4.4)
0
aC–N–C
0
α
= cos 120◦ . (4.5)
Dihedral springs are used to restrain the ω dihedral angles. They use the functional form
1
Vdihedral (x1 , x2 , x3 , x4 ) = k(minimage(dihedral(x1 , x2 , x3 , x4 ) − θ0 ))2 , (4.6)
2
where minimage(φ) adds a multiple of 2π to φ to give an angle in the range [−π, π]. Simu-
lations employ only fixed cis or trans angles without interconversion since the derivative of
this potential is discontinuous at the antipodal point to the equilibrium value. While the
fixed-ω constraint is limiting for describing the Boltzmann ensemble of unfolded proteins, it
is natural for describing kinetics as cis/trans interconversion takes place over much longer
time scales than most small-protein folding. The dihedral springs could be modified to have
a continuous derivative using the potential
in analogy with the von Mises distribution on the circle. A pair of these potentials could
be used to implement cis/trans interconversion, but this would require changes to the Ra-
machandran potentials which distinguish cis- and trans-proline.
The equilibrium θ0 is chosen to 0 or π as appropriate. The spring constant k is chosen to
be 30. Note that the dihedral springs place constraints on the spring constants of the angular
81
springs. If either of the internal angles of the dihedral group approach π (collinear atoms),
the derivative of the dihedral potential varies rapidly as the dihedral angle is undefined for
collinear atoms. The angle potential spring constants must be sufficiently large to avoid
these integration instabilities.
Dihedral springs are implement in the DihedralSpring class.
There are a variety of physical effects and statistical correlations in protein structures that
can be represented using a Ramachandran potential, which requires careful care to capture
only those effects that at not subsequently captured by the side chain potential.
The conceptually simplest method to construct to construct a Ramachandran poten-
tial is to first create a statistical model of the backbone conformations of protein in the
Protein Databank, such as the TorusDBN[2] model. By taking the negative logarithm of
such a model, one may obtain a potential energy surface for sampling from the probabil-
ity model. While such a model is conceptually straightforward to construct, it is fraught
with problems for interpreting the physicality of such a potential. Separately parameteriz-
ing the Ramachandran potential forces all other potential terms to model only the terms
statistically independent to the potential described by the Ramachandran angles. Such a
division is simple for terms like the bond springs between backbone atoms. Physically, there
is little important correlation to be expected between bond fluctuations and Ramachandran
propensities. The situation for side chain and hydrogen bonding terms is quite a bit more
complicated. Consider determining the helical fraction for a short peptide sequence in solu-
tion. A sufficiently powerful, statistical Ramachandran probability model may capture the
helicity of this peptide correctly (i.e. the helical probability of the peptide matches that of
experiment). While this would be a success of probability modeling for Ramachandran po-
tential, it is a challenge for the hydrogen bond and side chain potentials of the model. Since
the helicity is correct using only the Ramachandran potential, the hydrogen bond potential
82
and side chain potential must contribute nothing in aggregate to the helical probability of
the peptide. This is obviously unphysical, as surely the hydrophobicity of the residues or the
strength of hydrogen bond interactions are largely responsible for determining the helicity of
the peptide. Indeed, the effects have been moved out of the physical side chain and hydrogen
bonding interactions to be put in the Ramachandran potential. While this may work for
an isolated peptide with no tertiary structure, it is extremely hard to make such a model
transferable to larger systems, where the hydrophobic effects that must be omitted for the
peptide may be driving the formation of tertiary structure. This problem is quite general
to statistical potentials. Given the dihedral angles of protein, it is possible to recover the
atomic positions of the backbone to high accuracy, especially if care is taken to avoid steric
clash. For this reason, a sufficiently sophisticated statistical Ramachandran potential is able
to describe arbitrary protein physics, yet the stated purpose of Ramachandran potentials is
only to capture sequence local interactions.
The above argument shows that it is imperative to limit the capacity of the Ramachan-
dran potential to avoid capturing protein physics that should be described in the hydrogen
bonding and side chain interactions. Yet, it is not clear that the appropriate course of action
is to use similar dihedral potentials to those used in all-atom molecular dynamics. These
potentials are parameterized from quantum calculations of the energies of backbone torsions
and depend heavily on the side chain interactions to set local conformational preferences.
Such direct energetic models put a great strain on the side chain interactions to describe
the secondary structure, such as the secondary structure preferences of β-branched residues.
While this is appropriate for an all-atom model, it may be difficult to describe these con-
formational preferences in a coarse-grained model. Unfortunately, we must anticipate which
effects are likely to be described by the remaining terms of the Upside potential when we
choose the form of the Ramachandran potential.
83
We use a simple form for the Ramachandran potential
X
VRama (φ1 , ψ1 , . . . , φN , ψN ) = Vaai−1 ,aai ,aai+1 (φi , ψi ). (4.8)
i
84
L
separately provides faa (φ2 , ψ2 ), which is the smoothed Ramachandran frequency for
1 ,aa2
residues of type aa2 given that there is a residue of type aa1 preceding it in sequence, and
R
faa (φ2 , ψ2 ), which is the smoothed Ramachandran frequency for residues of type aa2
2 ,aa3
given that there is a residue of type aa1 following it in sequence. To combine these to two
frequencies into an approximation of the triplet frequency, we use a simple mixture
L
faa R
(φ2 , ψ2 ) + faa (φ2 , ψ2 )
mixture (φ , ψ ) = 1 ,aa2 2 ,aa3
faa 1 ,aa2 ,aa3 2 2 . (4.9)
2
This mixture is more diffuse than that we would obtain using the stricter ratio rule recom-
mended in [7],
L
faa R
(φ2 , ψ2 ) ∗ faa (φ2 , ψ2 )
mixture (φ , ψ ) ∝ 1 ,aa2 2 ,aa3
faa1 ,aa2 ,aa3 2 2 1-residue (φ , ψ )
, (4.10)
faa2 2 2
where the 1-residue frequency is used to avoid double-counting effects that depend only one
residue. While either rule is reasonable, we have noted better results from the less peaky
distributions obtained with the mixture rule.
Since these Ramachandran models empirically tends to create very helical Upside models,
we add an additional parameter λ to increase the proportion of β-sheet angles in the model.
To do so, we have fit a simple gaussian kernel density estimate for angles classifed as sheet
in DSSP, as well as the fraction faa1 ,aa2 of residues of each dimer type that are in β-sheets in
a representative sample of the PDB. We combine the TCB Ramachandran frequencies and
the sheet frequencies according to weighted mixture,
final
TCB
(1 − g)faa 1 ,aa2 ,aa3
(φ2 , ψ2 ) + g ∗ e−λ faa
sheet
1 ,aa2 ,aa3
(φ2 , ψ2 )
faa 1 ,aa2 ,aa3
(φ2 , ψ2 ; λ) = , (4.11)
(1 − g) + g ∗ e−λ
where g is the average of the DSSP sheet frequencies of the dimers (aa1 , aa2 ) and (aa2 , aa3 ).
The purpose of this combining rule is to only add sheet frequencies for residues observed to
be in the sheet conformation in the PDB. The parameter λ gives the contrastive divergence
85
Figure 4.1: Probability density of Ramachandran angles for the central residue of a free
alanine chain with no Ramachandran potential for the central residue (colors scale gives the
probability density values in units of 1/degrees2 ). The chain experience only short-range
excluded volume interactions, and interactions between adjacent residues are excluded from
the potential. The depression near φ = ψ = 0 occurs due to a high probability of i − 1, i + 1
steric overlap when the i-th residue is near the origin of φ/ψ space.
optimization a way to tune the amount of added sheet bias to compensate partially for the
effect of other terms in the model. To obtain the final Ramachandran energy term, we
perform a naive Boltzmann inversion (with a small modification below) to give
final
X
Vrama (φ1 , ψ1 , . . . , φN , ψN ) = − log faa i−1 ,aai ,aai+1
(φi , ψi ; λ), (4.12)
i
The potential V in (4.12) is sufficient to ensure that the Ramachandran frequencies of the
Boltzmann distribution are equal to the f final functions in the case that the only other terms
are the bond, angle, and dihedral springs given above. The situation is more complicated
when even short-range steric interactions are present to give proper atomic radii to the N,
Cα , and C atoms of the backbone, interactions which are described in more detail below.
86
The bias of Ramachandran angles for a small peptide with zero Ramachandran potential is
depicted in Fig. 4.1, which shows a preference against angles near φ = ψ = 0. The edge
of the bias region is sufficient to significantly alter the helical probability of proteins. The
bias depends only weakly on the Ramachandran potential of the flanking residues. For this
reason, we correct the bias by added a reference state potential
Because the dependence on the Ramachandran distribution of the flanking residues is weak,
there is no need to iteratively correct the bias.
Since the configuration of the protein chain is represented only by N, Cα , and C, we lack
the amide hydrogen (H) and carbonyl oxygen (O) atoms necessary to represent the physics
of hydrogen bonding. We remedy this difficiency by noting that the position of these atoms
can be reconstructed to high accuracy from the position of the backbone atoms. To find the
position of H and O, we compute
xH = xN + (0.88Å)dH (4.14)
xO = xC + (1.24Å)dO (4.15)
dO = sp3(xCα , xC , xN ) (4.17)
The coordinate computation is implemented in the Infer H O class, and the resulting coor-
dinates are used heavily in subsequent Upside potential calculations. The directions dH and
87
dO are the N–H and C–O bond vectors respectively, which are used to assess angular com-
ponents of both hydrogen bonding and side chain interactions. In Upside, the N-terminal
and proline residues have no amide hydrogen, while the C-terminal residue has no carbonyl
oxygen. No capping is currently implemented in Upside.
Given the position and bond vectors for the hydrogen and oxygen atoms, a hydrogen
bond score for each possible hydrogen bond is computed according to
hij = radial(|xH O HO O H HO
i − xj |) angular(dij · di ) angular(−di · dij ) (4.20)
dHO H
ij = unitvec(xi − xj )
O (4.21)
r − (1.4Å) (2.5Å) − r
radial(r) = sigmoid sigmoid (4.22)
0.10Å 0.125Å
a − 0.682
angular(a) = sigmoid . (4.23)
0.05
The hydrogen bonding score hij is non-symmetric in i and j and quantifies the extent to
which residue i and residue j form a hydrogen bond as the donor and acceptor, respec-
tively. The hydrogen bond score is in the range [0, 1], where the hydrogen bond score is one
approximately when 1.4Å < r < 2.5Å, θHOC < 47◦ , and θOHN < 47◦ .
Given the hydrogen bond score of each possible hydrogen bond, a hydrogen bond score in
the range [0, 1] is assigned to each hydrogen or oxygen atom. This score can be interpreted
as the confidence that the atom is participating in some hydrogen bond. We compute the
atomic hydrogen bond score as
sH
Y
i =1− (1 − hij ) (4.24)
j6=i
sO
Y
j =1− (1 − hij ). (4.25)
i6=j
88
bond to residue j and assuming that all such events are independent, (4.24) represents the
probability that residue i is a hydrogen bond donor for any other residue. This score is
constrained to the range [0, 1] and the sum of these scores provides a differentiable count of
the number of hydrogen bonds.
The Upside energy for backbone-backbone hydrogen bonding is given by
(sH O
X
Vhbond = Ehbond i + sj ), (4.26)
i
where Ehbond is a scalar quantity that approximately represents half of the energy of forming
a single hydrogen bond. The hydrogen bond scores are further used in the side chain potential
energy functions.
The hydrogen bond score is implemented in the ProteinHBond class and the backbone-
backbone hydrogen bond energy is implemented in the HBondEnergy class.
For specified χ-angles, the side chain atomic positions are fixed relative to the position and
orientation of N, Cα , and C atoms. We want to define the position and orientation of our
interaction sites in a side chain reference frame with N, Cα , and C in defined positions, then
apply rigid body rotation and translation to move the side chain interaction sites to their
appropriate positions for a given backbone structure.
In Upside, we define the reference backbone positions (in Å) to be
xref
N = (−1.21, −0.26, 0.) (4.27)
xref
Cα = (−0.02, 0.56, 0.) (4.28)
xref
C = ( 1.23, −0.30, 0.). (4.29)
89
The transformation between the reference frame backbone and the current configuration of
residue i is given by a rotation matrix Ui and a translation i, which are chosen to minimize
the RMSD,
1
|xa − (U xref 2
X
Ui , ti = arg min a + t)| . (4.30)
U,t 3
a∈(N,Cα ,C)
The tranformation Ui and ti are computed independently for each residue i using the quater-
nion algorithm of [5]. The computation of the alignment is handled by the AffineAlignment
class.
Given data (scalars, vectors, or points) in the reference frame of residue i, we compute
equivalent data in the simulation frame of the residue using the usual rules for covariant
transformation of spatial data. Scalar values, such as the probabilities of individual χ rotamer
states, are invariant under transformation. Vectors, such as bond vectors, are rotated by
Ui . Points, quanties which define a position in space, are transformed by both Ui and i.
Summarizing,
si = sref
i scalar transformation (4.31)
di = Ui dref
i vector transformation (4.32)
xi = Ui xref
i + ti point transformation. (4.33)
90
integrator steps. While using rigid body coordinates as dynamical variables has not yet been
pursued due to the added technical challenge of rigid body integration and thermostatting, it
is a promising area to speed simulation. If handled properly, it is likely that the Boltzmann
ensemble of the protein would be hardly affected for quantities of interest.
where xia is the position of the atom a of residue i. The position of atoms are actually
placed using the rigid body placement described above (necessary for Cβ but also true for
the other atoms).
Backbone sterics are implemented in the BackbonePairs class.
As described in the previous chapters, the rotamer states of the side chains are determined
by self-consistent belief propagation on single and pair interactions. Below, we describe the
individual interactions that combine to determine the interaction energies of the side chain
states. The subscript χ will be used to indicate the rotamer state for each iteraction residues,
so that xχi represents the position of residue i corresponding to rotamer χi .
This section will avoid discussing in detail energy functions describe in previous chapters.
The Ramachandran-dependent rotamer prior probabilities are adapted from [6] by summing
the probabilities of fine-grained states within each coarse-grained state used in Upside. The
91
implementation as a scalar (non-rotated) residue potential is straightforward, except to note
that the probabilities are converted to 1-residue energies using a negative logarithm transform
X
Vprior ({χi }) = − log pχi (φi , ψi ). (4.35)
i
The purpose of this energy is to account for within-residue side chain-backbone inter-
actions that may be more difficult for Upside to model using only pair interactions. For
example, β-branched amino acids have strong χ preferences due to the need accommodate
the backbone.
This potential is implemented using a scalar component of a PlacementNode class. The
same placement class handles placing the location and direction for each of the side chain
beads.
There are two terms for side chain-backbone interactions, both of which uses the pair inter-
action form defined in section 2.5.
The first term controls the interaction with the hydrogen and oxygen atoms with the side
chains. The position and direction of these atoms are given in section 2.5. The interaction is
modulated by κi = (1 − si )2 , where the si is given by (4.24) or (4.25) as appropriate, so that
the side chain interaction only occurs for atoms not already participating in a backbone-
backbone interaction. This interaction is implemented in the HBondCoverage class.
The second term allows the interaction of side chain atoms with three sites (position and
direction) on the protein backbone. The form of the interaction is identical to the hydrogen
and oxygen interaction, except κi = 1 so that there is no modulation of this interaction.
The interaction sites are initialized at the N, Cα , and C locations but move somewhat as
they are optimized in the side chain maximum likelihood training. This interaction is also
implemented by the HBondCoverage class but with different parameters than above.
92
4.8 Side chain-side chain interactions
The side chain-side chain interactions also the follow the form of (2.24) and are not mod-
ulated. The side chain-side chain interactions are implemented in the RotamerSidechain
class. Though the capability is currently unused, the model supports having multiple beads
for each side chain. This may improve the modeling of residues such as lysine or threonine
that have a mixture of polar and hydrophobic interactions.
The belief propagation algorithm for computing the side chain free energy is detailed in 2.11.
An important detail is that the belief propagation is not a convex optimization, and so
the stationary point to which the algorithm converges may be dependent on the initialization
of the algorithm. We initialize the algorithm to the probabilities implied by the 1-rotamer
energies, including the prior frequency energy and all the side chain-backbone interactions.
Note that the latter interactions are 1-rotamer interactions despite representing the interac-
tions of two beads, since each involves only a single side chain. The strong accuracy of the
prior probability of each rotamer at predicting crystallographic rotamer states [6] suggests
that initializing the rotamer probabilities to the 1-body probabilities has a high likelihood
of ensuring convergence to the global minimum of the approximate side chain free energy.
The theoretical soundness of the belief propagation free energy would be enhanced by
monitoring the belief propagation to identify if the optimization converges to different basins
on successive steps of the Verlet integration of the backbone. Such basin hopping will be
noted in an anomalously large change in the free energy resulting from a discrete change in
basin to which the algorithm converged. There is no guarantee that the old and new basins
have equal free energies; rather they may be artifact of sliding down one hill or another by
chance. If a large upward jump in energy were noted, presumably it could be treated with
Monte Carlo accept or reject step to some accuracy (possibly by embedding the algorithm
93
in a hybrid Monte Carlo framework[3]). As such jumps are expected to be rare, though a
careful study has not been performed, this potential source of error is currently uncontrolled.
The solve for marginals method in the RotamerSidechain class computes the belief
propagation solution for the side chain ensemble.
The Upside model as described above has no explicit solvation energy, only those effects
implied by the attraction of hydrophobic side chains to the backbone and other hydrophobic
side chains. Most approximate solvation interactions are designed for atomic simulation and
are inappropriately expensive for Upside. Furthermore, many models need to be modified
to deal with the reduced level of detail inherent in Upside’s representation of the protein.
Instead, the solvation model is quite similar to the simple model presented in [1], except that
the solvation potential is modified to be differentiable and to be optimized by contrastive
divergence. We call the term an environment interaction because it reflects the energy of the
typical environment of a given residue type, with the intention that this reflects a reasonable
energy for solvation.
We define a count of surrounding residues in a similar manner to [1] but modify the con-
struction so that it is differentiable.
The main component of the environment interaction is to count the number of side chains
beads within a fixed radius of the Cβ atom in a hemisphere above the atom. We define
! Cβ !
X |xji | − (8Å) di · x̂ji + 0.1
bi = pχj (φj , ψj ) sigmoid sigmoid (4.36)
1Å 1.0
j
C
xji = xSC β
j − xi , (4.37)
94
Cβ Cβ
where xi is the position of the Cβ on residue i and di is the corresponding Cα –Cβ bond
vector. The sigmoid parameters are chosen to approximately maximize distinctiveness of
the bi burial distributions for different residue types. In a well-formed definition of burial,
the burial distribution for a hydrophobic residue like valine should be very distinct from the
burial distribution of a charged residue like aspartic acid.
The side chain probability pχj (φj , ψj ) appearing in the definition (4.36) is the prior
probability of the side chain bead, not the marginal probability from belief propagation.
The derivative of the marginal probabilities with respect to the side chain positions are
complex, much moreso than the derivative of the free energy with respect to the side chain
positions. Furthermore, for intellectual self-consistency, the side chain rotamers ensemble
should account for burial interactions. Unfortunately, belief propagation is not defined for
many-body interactions, such as would arise from a nonlinear function of the burial bi .
It would be possible to extend belief propagation to handle many-body terms at a lower
level of approximation, although it is not clear how accurate such an approximation would
be and whether it would be guaranteed to converge. Due to the derivative difficulties of
using the marginal probabilities without perturbing for burial and the intellectual difficulties
of extending belief propagation to handle many-body terms, we simply avoid the issues
by using only the prior (in the sense of equation (4.35), not the values of the previous
step) probabilities to define the environment interactions. The precise probabilities used to
define burial are unlikely to compromise the ability of the bi to distinguish hydrophobic and
hydrophillic residues.
The burial calculation is implemented in the EnvironmentCoverage class.
95
4.10.2 Computing the non-linear energy
Given the burial values bi , the environment energy is an arbitray smooth function of the
burials, parametrized as a natural cubic spline. The total burial energy is given by
X
Vburial = vaai (bi ), (4.38)
i
where the vaai function depends only on the residue type aai of residue i. This function
should be considered as a method to correct the deficiencies in the model of the hydrophobic
effect implied by the pairwise side chain energies. As a solvation model, it is very crude
but the limits of the model are partially compensated by the co-training of the environment
energy with the side chain energies, so that at least the energy terms are co-adapted.
The environment energy is implemented in the NonlinearCoupling class.
4.11 References
[1] Aashish N Adhikari, Karl F Freed, and Tobin R Sosnick. De novo prediction of protein
folding pathways and structure using the principle of sequential stabilization. Proceedings
of the National Academy of Sciences, 109(43):17442–17447, 2012.
[2] Wouter Boomsma, Kanti V Mardia, Charles C Taylor, Jesper Ferkinghoff-Borg, An-
ders Krogh, and Thomas Hamelryck. A generative, probabilistic model of local protein
structure. Proceedings of the National Academy of Sciences, 105(26):8932–8937, 2008.
[3] Simon Duane, Anthony D Kennedy, Brian J Pendleton, and Duncan Roweth. Hybrid
monte carlo. Physics letters B, 195(2):216–222, 1987.
[4] Abhishek K Jha, Andres Colubri, Muhammad H Zaman, Shohei Koide, Tobin R Sosnick,
and Karl F Freed. Helix, sheet, and polyproline ii frequencies and strong nearest neighbor
effects in a restricted coil library. Biochemistry, 44(28):9691–9702, 2005.
96
[5] Pu Liu, Dimitris K Agrafiotis, and Douglas L Theobald. Fast determination of the
optimal rotational matrix for macromolecular superpositions. Journal of computational
chemistry, 31(7):1561–1563, 2010.
[7] Daniel Ting, Guoli Wang, Maxim Shapovalov, Rajib Mitra, Michael I Jordan, and
Roland L Dunbrack Jr. Neighbor-dependent ramachandran probability distributions of
amino acids developed from a hierarchical dirichlet process model. PLoS Comput Biol,
6(4):e1000763, 2010.
97
CHAPTER 5
FUTURE DIRECTIONS
The probabilistic view of protein coarse-graining is most vibrant and important when we
take seriously that the potential energy encodes a Boltzmann distribution, and work directly
to understand and optimize the full ensemble. Given the growing scale of computational
resources as well as large experimental data sets, we expect the methods espoused in thesis
to increase in importance. Furthermore, the rigorous coarse-graining typified in our side chain
interaction methods allow us to cross the boundary between the sharp detail of individual
interactions and smoother landscape of slow-degrees of freedom. Such methods will allow
the field to simulataneously attack the sampling and parameterization challenges of protein
simulation. In the future, we expect a synergy between computational speed and simulation
accuracy to supplant their current adversarial relationship precisely due to the ability to
rigorously coarse-grain simulations.
This thesis has demonstrated a path forward to use statistical information for rigorous
coarse-grained simulation of protein physics. We expect that the Upside model and associ-
ated software will be a strong starting point both for de novo prediction of structure and
dynamics as well as an integration point to meld experimental observables (NMR chemi-
cal shifts, evolutionary features, mutational data, etc.) with a strong and sensible prior on
protein structure. We further expect that the side chain free energy approximations that
we have developed will expand to a rich vein of methods which blur the distinction be-
tween included and omitted degrees of freedom in coarse-grained methods. Finally, as the
total throughput of molecular simulation increases, we expect that the parameterization of
even all-atom molecular dynamics will come to represent a more equal hybrid of quantum
and statistical information, and that experiences in pure statistical models like Upside will
greatly inform practitioners about the strengths and limitations of statistical training to
inform protein dynamics.
98
5.1 Future improvements to the Upside model and training
The ability to both train and run the Upside model is limited by the time required to equi-
librate a simulation trajectory. While the Upside model is designed for fast equilibration, it
may still take days or longer to equilibrate the Boltzmann ensemble for some proteins, even
those with fewer than 100 amino acids. Currently, the only enhanced sampling techniques
used in Upside are replica exchange and pivot moves. Replica exchange, however, is ineffec-
tive at accelerating the sampling of first-order phase transitions[2], and most small protein
folding is approximately a first-order phase transition[3]. Instead of temperature-enhanced
sampling, energy-enhanced sampling using the Wang-Landau[6] and other methods should
provide faster decorrelation for the Upside model.
Currently, Upside makes no use of multi-core parallelism outside of replica exchange.
While this is reasonable for very small proteins where the overhead of parallelism can be
significant, there are many potential applications of Upside to larger systems for both protein
association and conformational change that are greatly limited by the single-threaded nature
of Upside. The Upside model is amenable to parallelization. Most of the computational
cost occurs in evaluating the spline interactions necessary for the side chain interactions, a
trivially-parallelizable task. Future work will address parallelism in the Upside model.
Finally, persistent contrastive divergence[5] may be used to enhance the decorrelation
of structures during the contrastive divergence training. In this scheme, the individual
contrastive divergence trajectories are not reset on each pass through the training set, and
new proteins simulations are initialized from the last structure of their previous simulations.
This allows much greater exploration of the energy landscape. As long as the parameters do
not change too quickly, as controlled by the decorrelation time of the protein simulations,
this method can converge to the true maximum likelihood potential energy rather than the
approximately fluctuation-optimal potential energy that terminates contrastive divergence
99
optimization.
Coarse-grained modeling is often seen as a way to extract and understand the essential
elements of protein structure and dynamics, but the success or failure of a single coarse-
grained model often provides only cryptic evidence for the importance of the interactions
that it describes. Upside can successfully fold many proteins without the solvation-like
environment term. It is unreasonable to conclude that the partial success of a no-solvent
model proves that solvation is unimportant to protein folding. Instead, it just shows the
ability of side chain pair interactions to mask the absence of the solvation terms. Only by
constraining the model to reproduce a large number of experimental observables, such as side
chain packing and unfolded-state physics in addition to structure prediction, may we start to
understand the truly essential elements of protein folding. The ability of the Upside model
100
to make a wide range of experimental predictions, both dynamic and ensemble-averaged,
will allow us to explore the important physics of protein folding using a variety of terms that
include or exclude different types of protein physics while re-training to achieve the optimal
parameters for each model. By systematic experimentation, we should be able to better
characterize the essential elements of protein folding.
Since Upside is a general method for protein simulations, there are a wide variety of potential
future applications. We highlight a few applications that are either novel or uniquely suited
to Upside’s strengths.
Since the contrastive divergence training of Upside inherently samples the near-native states
of proteins, we expect the model to be especially accurate at describing the subglobal fluc-
tuations of the native state commonly found through hydrogen exchange. Using replica
exchange or even simple constant temperature simulation, we can map the local unfoldings
and other conformational changes of proteins to compare our ensemble to the predictions
of hydrogen exchange. Such a comparison should have the dual benefits of enhancing our
understanding of the Upside model as well as helping to interpret the results of experiments
with local unfoldings. This would will likely require understanding the effects of denaturant
in the Upside model, a subject of future research.
Upside’s high accuracy and computational speed make it a natural candidate to explore
protein-protein association, especially in cases where flexibility is needed to form the binding
interface. While the interaction parameters derived from contrastive divergence on single-
101
domain proteins should be fairly transferable to protein-protein association, it is not clear
that these parameters would be sufficiently accurate to make a state of the art prediction
method for protein association. Instead, we may use crystal structures of protein complexes
to fine-tune the interaction parameters through contrastive divergence. This work is ongoing
with Nabil Faruk and early results are promising.
102
augmented by an energy term that increased the probability of a contact in proportion to
the number of other proteins that also formed that contact for corresponding residues in a
multiple sequence alignment. In the case of a single protein, this represent ordinary structure
prediction through simulation. In the case of multiple proteins, each protein feels a force
encouraging it to adopt (a superset of) the average contact matrix of all of the homologous
sequences. The key idea is that homologous proteins will each have a low free energy for
the true native structure, but each protein may also have low free energy for incorrect
structures. So long as the incorrect structures are different for different homologues, they
will add incoherently to the potential, and thus every structure except the true native will
be weakened by the additional potential.
The advantages of the proposed technique are significant. This technique is far more data-
efficient than statistical analysis. Even using two sequences is likely to greatly enhance the
accuracy of structure prediction. Additionally, this cooperative folding incorporates strong,
well-calibrated information about protein structures from the Upside model, an independent
source of information from the statistical correlations. Finally, this method does not assume
that the contact matrices of homologous proteins are extremely similar; each protein is free
to have its own structure and contact matrix, which may be helpful in case a minority of
proteins are strongly incompatible with the structures of the majority of the collection.
Given the high efficiency of Upside simulation, we can directly simulate the large collection
of mutants studied in a typical mutational scan. A typical method to study these proteins
is an alanine-scan to compute φ-values at each residue. Instead of predicting the results
through careful study of a wild-type protein simulations[1], we may perform an em in silico
alanine-scan by running each of the mutant in Upside at reasonable computational cost.
Such studies are likely to provide much greater insight both into the nature of the observed
fractional φ-values as well as the trustworthiness of Upside simulation for the protein in
103
question.
5.3 References
[1] Robert B Best and Gerhard Hummer. Microscopic interpretation of folding φ-values
using the transition path ensemble. Proceedings of the National Academy of Sciences,
113(12):3263–3268, 2016.
[2] Jes Frellsen, Ole Winther, Zoubin Ghahramani, and Jesper Ferkinghoff-Borg. Bayesian
generalised ensemble markov chain monte carlo. In Proceedings of the 19th International
Conference on Artificial Intelligence and Statistics, pages 408–416, 2016.
[3] Hüseyin Kaya and Hue Sun Chan. Polymer principles of protein calorimetric two-state
cooperativity. Proteins: Structure, Function, and Bioinformatics, 40(4):637–661, 2000.
[4] Debora S Marks, Thomas A Hopf, and Chris Sander. Protein structure prediction from
sequence variation. Nature biotechnology, 30(11):1072–1080, 2012.
[5] Tijmen Tieleman. Training restricted boltzmann machines using approximations to the
likelihood gradient. In Proceedings of the 25th international conference on Machine learn-
ing, pages 1064–1071. ACM, 2008.
104
[6] Fugao Wang and DP Landau. Determining the density of states for classical statistical
models: A random walk algorithm to produce a flat histogram. Physical Review E,
64(5):056101, 2001.
[7] Sheng Wang, Siqi Sun, Zhen Li, Renyu Zhang, and Jinbo Xu. Accurate de novo prediction
of protein contact map by ultra-deep learning model. arXiv preprint arXiv:1609.00680,
2016.
105