ArticlePDF Available

Molecular Dynamics Simulations of Shockwave Affected STMV Virus to Measure the Frequencies of the Oscillatory Response

Authors:

Abstract and Figures

Acoustic shockwaves are of interest as a possible means of the selective inactivation of viruses. It has been proposed that such inactivation may be enhanced by driving the virus particles at frequencies matching the characteristic frequency corresponding to acoustic modes of the viral structures, setting up a resonant response. Characteristic frequencies of viruses have been previously studied through opto-mechanical techniques. In contrast to optical excitation, shockwaves may be able to probe acoustic modes without the limitation of optical selection rules. This work explores molecular dynamics simulations of shockwaves interacting with a single STMV virus structure, in full atomistic detail, in order to measure the frequency of the response of the overall structure. Shockwaves of varying energy were set up in a water box containing the STMV structure by assigning water molecules at the edge of the box with an elevated velocity inward—in the direction of the virus. It was found that the structure compressed and stretched in a periodic oscillation of frequency 65 ± 6.5 GHz. This measured frequency did not show strong dependency on the energy of the shockwave perturbing the structure, suggesting the frequency is a characteristic of the structure. The measured frequency is also consistent with values predicted from elastic theory. Additionally, it was found that subjecting the virus to repeated shockwaves led to further deformation of the structure and the magnitude of the overall deformation could be altered by varying the time delay between repeated shockwave pulses.
Content may be subject to copyright.


Citation: Burkhartsmeyer, J.;
Wong, K.S. Molecular Dynamics
Simulations of Shockwave Affected
STMV Virus to Measure the
Frequencies of the Oscillatory
Response. Acoustics 2022,4, 268–275.
https://doi.org/10.3390/
acoustics4010016
Academic Editors: Fardin Khalili,
AmirtahàTaebi and Hansen
A. Mansy
Received: 10 December 2021
Accepted: 14 March 2022
Published: 18 March 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
acoustics
Article
Molecular Dynamics Simulations of Shockwave Affected
STMV Virus to Measure the Frequencies of the
Oscillatory Response
Jeffrey Burkhartsmeyer * and Kam Sing Wong
Department of Physics, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon,
Hong Kong, China; phkswong@ust.hk
*Correspondence: jmb@connect.ust.hk
Abstract:
Acoustic shockwaves are of interest as a possible means of the selective inactivation of
viruses. It has been proposed that such inactivation may be enhanced by driving the virus particles
at frequencies matching the characteristic frequency corresponding to acoustic modes of the viral
structures, setting up a resonant response. Characteristic frequencies of viruses have been previously
studied through opto-mechanical techniques. In contrast to optical excitation, shockwaves may be
able to probe acoustic modes without the limitation of optical selection rules. This work explores
molecular dynamics simulations of shockwaves interacting with a single STMV virus structure,
in full atomistic detail, in order to measure the frequency of the response of the overall structure.
Shockwaves of varying energy were set up in a water box containing the STMV structure by assigning
water molecules at the edge of the box with an elevated velocity inward—in the direction of the
virus. It was found that the structure compressed and stretched in a periodic oscillation of frequency
65 ±6.5 GHz
. This measured frequency did not show strong dependency on the energy of the
shockwave perturbing the structure, suggesting the frequency is a characteristic of the structure. The
measured frequency is also consistent with values predicted from elastic theory. Additionally, it was
found that subjecting the virus to repeated shockwaves led to further deformation of the structure
and the magnitude of the overall deformation could be altered by varying the time delay between
repeated shockwave pulses.
Keywords: molecular dynamics; acoustic shockwave; virus biophysics; harmonic modes
1. Introduction
Laser-generated acoustic shockwaves or “transients” have been previously used to
break apart viruses in order to study their structural bonding forces [
1
,
2
]. Elsewhere,
selective inactivation of viruses by pulsed laser irradiation has been observed [
3
], with
impulsive stimulated Raman scattering as the suggested mechanism. Recent work proposes
acoustic shockwaves may play a role in pulsed laser inactivation of viruses with plasmonic
enhancement of the laser absorption, from the addition of metal nanoparticles [
4
]. Other
works have proposed the inactivation of viruses via resonant absorption of ultrasonic [
5
]
or microwave energy [
6
,
7
]. Resonant absorption would involve driving the virus at its
characteristic frequencies, such that the vibrational response matches the frequency of
driving energy, leading to increasing amplitude of oscillations and eventual failure of the
structure. For this purpose it is useful to have an estimate of the characteristic frequencies
of a target virus.
Early analytical models used elastic theory to predict vibrational frequencies of
viruses [
8
12
]. More recent computational models have been developed to study such fre-
quencies using elastic network normal mode analysis [
13
15
]. Another study [
16
] models
the subunits of the structure at the atomic level, applying constraints due to the inherent
symmetries of the icosahedral structure, to predict the vibrational frequencies of a virus
Acoustics 2022,4, 268–275. https://doi.org/10.3390/acoustics4010016 https://www.mdpi.com/journal/acoustics
Acoustics 2022,4269
in the absence of explicit water. In contrast to these, the present work treats the entire
structure of a fully solvated virus in full atomistic detail, by manipulating the surrounding
water molecules to perturb the system with an acoustic shock wave, and measures the
resulting oscillations in the structure.
Opto-mechanical studies involving Brillouin and Raman scattering techniques have
previously been conducted to measure the optically active modes of viruses [
17
23
]. Opti-
cally active modes in these cases are those which obey selection rules for light coupling
with the acoustic modes of the material [
24
]. In contrast to this, shock wave perturbation is
a direct mechanical displacement of molecules via Coulombic and other near-field forces
between the water molecules and those of the virus structure. Thus, the modes available
for probing by shockwave perturbation are not restricted by the optical selection rules.
The Satellite Tobacco Mosaic Virus (STMV) has been studied extensively, as a small-
sized virus~17 nm in diameter. Its structure is well-known from crystallography studies—
its geometry is icosahedral as are many other viruses, and thus it is a representative case
study for such viruses. The capsid is comprised of 60 identical copies of a single protein with
molecular weight of 14,500. The structure has been studied via molecular dynamics (MD)
simulations [
25
,
26
]. In fact, because it is a relatively large structure for such MD simulations,
comprised of millions of molecules, it is often used as a benchmark to test the performance
of computing systems [
27
,
28
]. Shockwaves and their effects on biological systems have
been simulated in previous work [
29
33
], using different methods to set up the shockwaves
in the simulation. For example Surdutovich [
29
31
] initiates the shock waves using a “hot
cylinder” of atoms in a specified radius with average velocities higher than those of the
surrounding media, which in turn expand and interact with neighboring atoms and thus
propagate outward. These simulated shockwaves then passed through DNA structures,
in order to study whether such waves would damage DNA. Since molecular dynamics
simulations treat covalent bonds as harmonic oscillators, and do not take into account
such bonds breaking, Surdutovich considered any bonds exhibiting vibrational energies
above a certain threshold as being broken during the simulation. Other methods for setting
up shockwaves such as that of [
32
,
33
] also involve selecting a region of water molecules
surrounding the biological system of interest, and giving these water molecules a velocity
in a particular direction. This is also the approach used in the present work, described in
more detail below.
2. Materials and Methods
The structure of the Satellite Tobacco Mosaic Virus (STMV), solvated in a cube of
water of cell dimensions 22 nm, was prepared based on files available from the website
of the Theoretical and Computational Biophysics group, NIH Center for Macromolecular
Modeling and Bioinformatics, at the Beckman Institute, University of Illinois at Urbana-
Champaign. The Protein Structure File (PSF) contains the structure of the virus capsid
and RNA, along with water molecules of the solution. The Protein Data Bank (PDB) file
contains information about the coordinates of the atoms from the PSF file, for the start of the
simulation. Using Nanoscale Molecular Dynamics (NAMD) simulation software [
34
], the
system was then allowed to equilibrate under a constant pressure of 1 atm, and a constant
temperature of 298 K for 15 ps. The output from this step was used as the input for the
subsequent step, in which a shockwave was introduced in the following way. Along the
edge of the water box, perpendicular to the Z-axis, atoms falling within a certain distance
of the edge of the box (arbitrarily chosen to be ~3 nm) had their velocities altered. The
X and Y velocities were left unchanged, but the Z velocity was changed to be a fixed value
for these specific atoms. This method for producing simulated shockwaves is similar to
the one employed in [
33
]. The forces on all atoms in the system are then computed via the
force field resulting from interaction with neighboring atoms within the cutoff distance—in
this case 12 Å. For a given time-step, the velocity and new position of each atom are then
numerically computed from the velocity and position of the prior time-step and the overall
force on the atom from the current time-step. This process is repeated for each time-step,
Acoustics 2022,4270
of duration 1 fs in this case, and a trajectory for all atoms is computed. Since the selected
water molecules were given an elevated velocity inward, towards the rest of the water
box containing the virus, they move closer to other water molecules, resulting in forces
acting on these neighboring atoms, so that they too gain an elevated velocity. In this
manner, in subsequent time-steps the shockwave propagates through water molecules until
it eventually encounters the molecules of the virus structure.
Perhaps a more realistic model of the initial velocities which would occur in an actual
physical shockwave would instead have a distribution of velocities centered around this
chosen value, but this method of using a uniform value was simply chosen for convenience,
since it made identifying the affected atoms simple for debugging purposes and to easily
vary the parameter of energy input. To justify the use of this simplistic approach to creating
shockwaves in the water box, it should be noted that the water molecules which have
been subjected to this artificial instantaneous Z-direction velocity, will quickly (by the next
processing step) encounter other water molecules outside the affected region. As they
interact with these unaltered water molecules, there will be an exchange of momentum
through the force field, and the random distribution of the velocities of those unaffected
water molecules will once again smooth out and randomize the velocities of those molecules
which underwent the artificial velocity change. Further processing steps will lead to more
interactions with randomly moving water molecules and the strong pulse of velocities
in the Z-direction will disperse into other directions, but the overall momentum will be
biased in the Z-direction and the pulse will propagate through the rest of the system. Thus,
provided the molecules of the biological system (the virus in this case) are far enough
away from the edge of the box, by the time this pulse reaches the molecules of interest, it
should resemble a planar acoustic wave transient, such as that which could be produced by
an ultrashort laser pulse. Indeed, measuring the distribution of Z-direction velocities at
various times shows that after a few time-steps, the velocities of the water molecules fall in a
skewed Gaussian distribution. In addition to the aforementioned convenience of being able
to quickly pick out the atoms which were subject to the artificial velocity change (since they
are all assigned the same Z-direction value for the initial step), there is the added benefit of
being able to inject a very high initial Z-directed energy pulse, while avoiding the velocity
limits of the NAMD simulation system. If more energy is still needed, this can be achieved
by simply choosing a larger slice of atoms along the boundary. Another approach would be
to instead simply reverse the Z-direction velocities of these same atoms, since they should
already have a realistic distribution of velocities after the equilibrium step; this approach is
known as a “momentum mirror.” [
35
] It has the advantage of not altering the overall kinetic
energy of the system—however, this would not be representative of a real laser-generated
shock wave as in this case kinetic energy is added to the system via absorption of the laser
energy. In another, similar approach, a slowly moving boundary could be applied, and
atoms reaching this boundary would have their velocities reversed—like a “hard edge.”
This is the “piston” approach of [
32
], and it has the advantage of avoiding clashes of atoms
which may arise when the sudden change in velocities causes atoms to come too close
together. Such clashes can cause the simulation to fail, since the calculated forces become so
large that they are beyond the ability for the computer to store and compute. However, if a
sufficiently small time-step is chosen, clashes can be avoided since the atoms have time to
react to the forces generated when they approach one another, thus altering their direction
and avoiding unmanageably large force values. For this reason, for some simulations
performed in this work—those involving the highest energies—a very short timestep of 1
fs was used for parts of the simulation.
3. Results
Using NAMD, the STMV virus structure in a water box was subjected to a single pulse
of a simulated acoustic shockwave using the method described above, at various pulse
energies, and the response of the system studied. Using Visual Molecular Dynamics (VMD)
software [
36
], conformational changes were studied. The Root Mean Squared Displacement
Acoustics 2022,4271
(RMSD) for the entire protein backbone of the virus capsid structure was calculated using
the built-in plug-in of this software. A higher value indicates a larger overall conformational
change to the entire structure throughout the trajectory. Shown in Figure 1a are the RMSD
curves for simulated shockwaves at various energies. Shown in Figure 1b are RMSD curves
for repeated shockwaves compared to a single pulse. For each pulse, the initial velocity
of water molecules is 39.999 Å/ps—a relatively gentle pulse. By varying the time delay
between pulses, the amplitude of the overall deformation of the structure is seen to vary.
Shown in Figure 2are screenshots from the VMD visualization of the virus subject to a
shockwave of initial water speed 99.999 Å/ps (the maximum allowable velocity of the
software—chosen to show the deformation most clearly), at time points corresponding to
extrema in the RMSD curve. Figure 2a shows an illustration of a planar cross-section of a
sphere with the L3 mode at maximum deformation for comparison. Figure 2b shows the
virus before encountering the shockwave—at time 0 and 0 RMSD, with the spherical cross-
section from Figure 2a overlaid for comparison. Figure 2c shows the virus at maximum
deformation, with the cross-section of L3 mode at max deformation overlaid.
Acoustics 2022, 4 FOR PEER REVIEW 4
3. Results
Using NAMD, the STMV virus structure in a water box was subjected to a single
pulse of a simulated acoustic shockwave using the method described above, at various
pulse energies, and the response of the system studied. Using Visual Molecular Dynam-
ics (VMD) software [36], conformational changes were studied. The Root Mean Squared
Displacement (RMSD) for the entire protein backbone of the virus capsid structure was
calculated using the built-in plug-in of this software. A higher value indicates a larger
overall conformational change to the entire structure throughout the trajectory. Shown in
Figure 1a are the RMSD curves for simulated shockwaves at various energies. Shown in
Figure 1b are RMSD curves for repeated shockwaves compared to a single pulse. For
each pulse, the initial velocity of water molecules is 39.999 Å /psa relatively gentle
pulse. By varying the time delay between pulses, the amplitude of the overall defor-
mation of the structure is seen to vary. Shown in Figure 2 are screenshots from the VMD
visualization of the virus subject to a shockwave of initial water speed 99.999 Å /ps (the
maximum allowable velocity of the softwarechosen to show the deformation most
clearly), at time points corresponding to extrema in the RMSD curve. Figure 2a shows an
illustration of a planar cross-section of a sphere with the L3 mode at maximum defor-
mation for comparison. Figure 2b shows the virus before encountering the shock-
waveat time 0 and 0 RMSD, with the spherical cross-section from Figure 2a overlaid for
comparison. Figure 2c shows the virus at maximum deformation, with the cross-section
of L3 mode at max deformation overlaid.
(a)
0
2
4
6
8
10
0246810 12 14
RMSD (Å)
Time (ps)
RMSD Virus Backbone
for Varying Initial Water Velocities
99.999 (Å/ps)
89.999 ( Å/ps)
79.999 ( Å/ps)
69.999 (Å/ps)
59.999 (Å/ps)
49.999 (Å/ps)
39.999 (Å/ps)
Acoustics 2022, 4 FOR PEER REVIEW 5
(b)
Figure 1. Response of virus structure to shockwave pulses (a) RMSD curves exhibiting the oscilla-
tory response of the virus to single shockwave pulseseach curve represents the time evolution of
the deformation of the virus structure for a given initial velocity value assigned to water molecules
in the water slice to initialize the shockwave. (b) RMSD curves showing response of the virus to
repeated shockwaves pulses compared to a single shockwave pulsea varying delay between
pulses leads to variation in the overall amplitude of deformation of the structure.
(a)
(b)
(c)
Figure 2. Output from VMD visualization software for the shockwave with initial water velocities
of 99.999 Å /ps showing: (a) Illustration of cross-sectional shape of L3 spheroidal mode at maximum
deformation (solid line), compared to initial spherical shape (dashed line) (b) The virus in its
equilibrated state (time = 0.0 ps)before encountering the shockwave, with spherical cross-section
overlaid (dashed line) (c) The virus in the maximum compression state (time = 1.4 ps), corre-
sponding to the first peak in the RMSD value, with cross-section of L3 mode of a sphere at maxi-
mum deformation overlaid (solid line). The color scheme shows the different types of mole-
culesthe purple molecules are the nucleic acid of the core while the green molecules are the pro-
tein capsid.
As seen in Figures 1 and 2, as well as video SV1, when the simulated shockwave
front begins to hit the virus at time 200 fs, the structure begins to deform and the mag-
nitude of this deformation increases as the wave propagates through the structure and
the surrounding water medium, peaking at time 1.4 ps, when the structure is at maxi-
mum compression. As the wave reaches the end of the structure, it is stretched out, re-
0
2
4
6
8
10
0 5 10 15 20 25 30
RMSD (Å)
Time (ps)
RMSD Virus Backbone
for Varying Repetition Rate
Single Pulse
Repeated Pulses (6 ps delay)
Repeated Pulses (8 ps delay)
Repeated Pulses (9 ps delay)
Figure 1.
Response of virus structure to shockwave pulses (
a
) RMSD curves exhibiting the oscillatory
response of the virus to single shockwave pulses—each curve represents the time evolution of the
deformation of the virus structure for a given initial velocity value assigned to water molecules in the
water slice to initialize the shockwave. (
b
) RMSD curves showing response of the virus to repeated
shockwaves pulses compared to a single shockwave pulse—a varying delay between pulses leads to
variation in the overall amplitude of deformation of the structure.
Acoustics 2022,4272
(b)
(a)
(b)
(c)
0
2
4
6
8
10
0 5 10 15 20 25 30
RMSD (Å)
Time (ps)
RMSD Virus Backbone
for Varying Repetition Rate
Single Pulse
Repeated Pulses (6 ps delay)
Repeated Pulses (8 ps delay)
Repeated Pulses (9 ps delay)
Figure 2.
Output from VMD visualization software for the shockwave with initial water velocities of
99.999 Å/ps showing: (
a
) Illustration of cross-sectional shape of L3 spheroidal mode at maximum
deformation (solid line), compared to initial spherical shape (dashed line) (
b
) The virus in its equili-
brated state (time = 0.0 ps)—before encountering the shockwave, with spherical cross-section overlaid
(dashed line) (
c
) The virus in the maximum compression state (time = 1.4 ps), corresponding to the
first peak in the RMSD value, with cross-section of L3 mode of a sphere at maximum deformation
overlaid (solid line). The color scheme shows the different types of molecules—the purple molecules
are the nucleic acid of the core while the green molecules are the protein capsid.
As seen in Figures 1and 2, as well as video SV1, when the simulated shockwave front
begins to hit the virus at time 200 fs, the structure begins to deform and the magnitude of this
deformation increases as the wave propagates through the structure and the surrounding
water medium, peaking at time 1.4 ps, when the structure is at maximum compression.
As the wave reaches the end of the structure, it is stretched out, resulting in a secondary
peak in the RMSD at time (9 ps). Finally, there is a recoiling of the overall structure as
the structure is compressed from the recoil. This is reminiscent of a damped harmonic
oscillator type motion, better described in the model of [
8
], with spherical Bessel functions.
The time in between the first and second peaks ranges from 7.0 ps to 8.4 ps, for pulses with
average initial velocity ranging from 99.999 Å/ps and 29.999 Å/ps. Since the primary and
secondary peaks represent the times of maximal compression and stretching, respectively,
the overall time for the complete oscillatory motion is double the time between first and
second peak. These times correspond to 59.5 to 71.4 GHz. As seen in Table 1, there is little
correlation between this measured frequency and the initial velocities of the water slice
molecules used to set up the simulated shockwave.
Table 1.
Initial velocity of water molecules in shockwave initialization water slice vs. measured
frequency of the virus structure oscillation.
Initial Velocity Period of Oscillation (ps) Frequency (GHz)
Water Slice (Å/ps)
99.999 15.2 65.79
89.999 16 62.5
79.999 16.4 60.98
69.999 16.4 60.98
59.999 16.8 59.52
49.999 14 71.43
39.999 14 71.43
29.999 14.8 67.57
4. Discussion
The simulated structural response of STMV exposed to a shockwave exhibited oscilla-
tory behavior, the periodicity of which did not show high dependency on the initial energy
Acoustics 2022,4273
of the shockwave. This suggests that the frequency may be an intrinsic property of the
structure. This frequency was determined to be 65 ±6.5 GHz.
For a comparison with values predicted from elastic theory, the icosahedral virus can
be approximated as a sphere, in order to employ Lamb’s theory [
9
,
37
,
38
] for estimating
vibrational frequencies. Taking values of longitudinal wave speed and Poisson’s ratio to be
1800 m/s and 0.30, respectively (corresponding to a transverse wave speed of 900 m/s),
consistent with values used in prior works [
6
,
11
,
18
,
22
], and with a radius of 17 nm, the
L1 and L3 modes are predicted to be ~60.6 GHz and 66.6 GHz, respectively, which are
in good agreement with the measured value. It should not be surprising that these odd-
numbered modes would be activated by a pulse energy which asymmetrically compresses
the structure from one side, leaving the other side initially unperturbed. Indeed from
visual inspection of the oscillatory motion using the VMD software, the shape of the
compressed state appears to most resemble that of the L3 mode, as seen in Figure 2c. Thus,
it seems reasonable that the predicted L3 mode frequency (66.6 GHz) should be a good
approximation of the measured values (65
±
6.5 GHz.) Note, however that the error in this
measurement is quite significant (10%), so the L1 mode could be considered an equally
good numeric fit. Table A1 shows the first few spheroidal modes, for comparison. Further,
the real motion of the virus oscillations is more complex than that of a simple sphere, as
the composition of the structure is very complex and the shockwave perturbation can
activate more than one mode. Moreover, the parameters (Poisson’s ratio, wave speeds of
the material) used in the Lamb’s theory approximation are not entirely well-established
in the literature, so these should be considered estimations as well. Given all the sources
of error, it is reassuring that the predicted values match the measured values so well, but
this should be taken with some care. For example if a Poisson’s ratio of 0.40 were instead
used, this would result in a predicted value of 53.6 GHz for the frequency of the L3 mode.
It is also worth noting that, although it is reassuring that elastic theory gives similar values
for vibrational frequencies, the full atomistic molecular dynamics model gives additional
information by taking into account atomic interactions of the biological components with
one another and with the water molecules. These could be approximated by introducing a
damping term to the elastic theory model, but molecular dynamics simulation gives a more
comprehensive understanding of the dynamics of the system. For example, it appears that
the vibrations quickly dissipate and begin to level off after the first couple of peaks in the
RMSD curve, as seen in Figure 1a.
It was further observed that an enhancement of the amplitude of deformation could
be achieved by subjecting the virus to repeated shock-wave pulses. By choosing a time
delay between pulses, such that the deformation of the structure is at a maximum as a
subsequent shockwave pulse arrives at the structure, the overall deformation increased,
as seen in Figure 1b. This supports the plausibility of achieving a resonant response by
driving the virus structure at the proper frequency, via acoustic waves such as shockwaves.
Future research may explore lower energy perturbations—for example, using water
speeds which do not exceed the wave speed of water, in order to see if the oscillatory mo-
tions are still observable and whether the frequencies of such motions would be consistent
with those measured using these super-sonic shockwave pulses. If oscillatory frequencies
are ubiquitous, irrespective of excitation energy, it may suggest the possibility of driving
the virus structure at such frequencies, in order to set up a resonant response.
Another interesting direction for future study would be to examine the possibility
of virus inactivation, by investigating bond-breaking events which may occur during
shockwave excitation. The results described here treated bonds as a simple harmonic
potential, which is unrealistic as bond energies simply increase unbounded with bond-
length. More interesting would be to simulate shockwave interaction with viruses using
more advanced MD simulation techniques such as ab initio or reactive force field methods
in order to study bond-breaking and re-formation. This would give a more realistic view as
to the plausibility of viral inactivation via resonant excitation of the structure.
Acoustics 2022,4274
In spite of the relative simplicity of this model, some useful observations can be made.
To wit: frequencies of oscillation due to shockwave perturbation seem to correspond to
the odd-numbered spheroidal modes (L1 and L3), and the frequencies are around 65 GHz,
irrespective of shockwave energy. With these observations, it may be possible to tailor a
scheme whereby selective inactivation of viruses could be enhanced by driving the system
at such a characteristic frequency—either via shockwaves or otherwise.
Supplementary Materials:
The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/acoustics4010016/s1, Video S1: VMD Visualization of Virus
Response to Shockwave. File S1: NAMD instruction file.
Author Contributions:
Conceptualization, J.B. and K.S.W.; software and simulation, J.B.; supervision,
K.S.W.; writing—original draft preparation, J.B.; writing—review and editing, K.S.W. All authors
have read and agreed to the published version of the manuscript.
Funding:
This research was funded by the Research Grants Council of Hong Kong through grant
numbers AoE/P-02/12 and C7035-20G.
Data Availability Statement:
Datasets and NAMD instruction files can be found at the following:
https://figshare.com/articles/software/stmv_shockwave_zip/19380299 (accessed on
8 December 2021
).
Conflicts of Interest: The authors declare no conflict of interest.
Appendix A
Shockwave simulations were performed with Charm22 force field [
39
], rigid water
bonds and periodic boundary conditions. For further simulation details, the NAMD
simulation instruction file is provided in the supplementary materials section.
Table A1.
Frequencies of the first few modes of a sphere, with longitudinal wave speed 1800 m/s
and Poisson’s ratio 0.30.
Spheroidal Mode Number Frequency (GHz)
L0 92.47
L1 60.60
L2 44.67
L3 66.57
References
1. Hamrick, P.E.; Cleary, S.F. Laser-Induced Acoustic Breakage of Tobacco Mosaic Virus. Nature 1968,220, 909–910. [CrossRef]
2.
Smith, R.J.; Cleary, S.F. Investigation of Structural Bonding Forces in Bacteriophage T2. J. Acoust. Soc. Am.
1974
,56, 1883–1889.
[CrossRef]
3.
Tsen, S.-W.D.; Kingsley, D.H.; Poweleit, C.; Achilefu, S.; Soroka, D.S.; Wu, T.C.; Tsen, K.-T. Studies of Inactivation Mechanism of
Non-Enveloped Icosahedral Virus by a Visible Ultrashort Pulsed Laser. Virol. J. 2014,11, 20. [CrossRef]
4.
Nazari, M.; Xi, M.; Lerch, S.; Alizadeh, M.; Ettinger, C.; Akiyama, H.; Gillespie, C.; Gummuluru, S.; Erramilli, S.; Reinhard, B.M.
Plasmonic Enhancement of Selective Photonic Virus Inactivation. Sci. Rep. 2017,7, 11951. [CrossRef]
5.
Babincová, M.; Sourivong, P.; Babinec, P. Resonant absorption of ultrasound energy as a method of HIV destruction. Med.
Hypotheses 2000,55, 450. [CrossRef]
6.
Yang, S.-C.; Lin, H.-C.; Liu, T.-M.; Lu, J.-T.; Hung, W.-T.; Huang, Y.-R.; Tsai, Y.-C.; Kao, C.-L.; Chen, S.-Y.; Sun, C.-K. Efficient
Structure Resonance Energy Transfer from Microwaves to Confined Acoustic Vibrations in Viruses. Sci. Rep.
2015
,5, 18030.
[CrossRef]
7.
Hempel, E. Irradiation Device for Influencing a Biological Structure in a Subject with Electromagnetic Radiation 2010.
U.S. Patent US7648498B2, 19 January 2010.
8. Ford, L. Estimate of the Vibrational Frequencies of Spherical Virus Particles. Phys. Rev. E 2003,67, 051924. [CrossRef]
9.
Saviot, L.; Murray, D.B.; Mermet, A.; Duval, E. Comment on “Estimate of the Vibrational Frequencies of Spherical Virus Particles”.
Phys. Rev. E 2004,69, 023901. [CrossRef]
10.
Fonoberov, V.A.; Balandin, A.A. Low-Frequency Vibrational Modes of Viruses Used for Nanoelectronic Self-Assemblies. Phys.
Status Solidi (B) 2004,241, R67–R69. [CrossRef]
11.
Talati, M.; Jha, P.K. Acoustic Phonon Quantization and Low-Frequency Raman Spectra of Spherical Viruses. Phys. Rev. E
2006
,73,
011901. [CrossRef]
Acoustics 2022,4275
12.
Yang, Z.; Bahar, I.; Widom, M. Vibrational Dynamics of Icosahedrally Symmetric Biomolecular Assemblies Compared with
Predictions Based on Continuum Elasticity. Biophys. J. 2009,96, 4438–4448. [CrossRef]
13.
Tama, F.; Brooks III, C.L. Diversity and Identity of Mechanical Properties of Icosahedral Viral Capsids Studied with Elastic
Network Normal Mode Analysis. J. Mol. Biol. 2005,345, 299–314. [CrossRef]
14.
Dykeman, E.C.; Sankey, O.F. Normal Mode Analysis and Applications in Biological Physics. J. Phys. Condens. Matter
2010
,22,
423202. [CrossRef]
15.
Lee, B.H.; Jo, S.; Choi, M.; Kim, M.H.; Choi, J.B.; Kim, M.K. Normal Mode Analysis of Zika Virus. Comput. Biol. Chem.
2018
,72,
53–61. [CrossRef]
16.
Dykeman, E.C.; Sankey, O.F. Atomistic Modeling of the Low-Frequency Mechanical Modes and Raman Spectra of Icosahedral
Virus Capsids. Phys. Rev. E 2010,81, 021918. [CrossRef] [PubMed]
17.
Tsen, K.-T.; Dykeman, E.C.; Sankey, O.F.; Tsen, S.-W.D.; Lin, N.-T.; Kiang, J.G. Raman Scattering Studies of the Low-Frequency
Vibrational Modes of Bacteriophage M13 in Water—Observation of an Axial Torsion Mode. Nanotechnology
2006
,17, 5474.
[CrossRef]
18.
Stephanidis, B.; Adichtchev, S.; Gouet, P.; McPherson, A.; Mermet, A. Elastic Properties of Viruses. Biophys. J.
2007
,93, 1354–1359.
[CrossRef] [PubMed]
19.
Hartschuh, R.; Wargacki, S.; Xiong, H.; Neiswinger, J.; Kisliuk, A.; Sihn, S.; Ward, V.; Vaia, R.; Sokolov, A. How Rigid Are Viruses.
Phys. Rev. E 2008,78, 021907. [CrossRef]
20.
Tcherniega, N.; Pershin, S.; Bunkin, A.; Donchenko, E.; Karpova, O.; Kudryavtseva, A.; Lednev, V.; Mironova, T.; Shevchenko, M.;
Strokov, M.; et al. Laser Excitation of Gigahertz Vibrations in Cauliflower Mosaic Viruses’ Suspension. Laser Phys. Lett.
2018
,15,
095603. [CrossRef]
21.
Sirotkin, S.; Mermet, A.; Bergoin, M.; Ward, V.; Van Etten, J.L. Viruses as Nanoparticles: Structure versus Collective Dynamics.
Phys. Rev. E 2014,90, 022718. [CrossRef]
22.
Burkhartsmeyer, J.; Wang, Y.; Wong, K.S.; Gordon, R. Optical Trapping, Sizing, and Probing Acoustic Modes of a Small Virus.
Appl. Sci. 2020,10, 394. [CrossRef]
23.
Mattarelli, M.; Vassalli, M.; Caponi, S. Relevant Length Scales in Brillouin Imaging of Biomaterials: The Interplay between
Phonons Propagation and Light Focalization. ACS Photonics 2020,7, 2319–2328. [CrossRef]
24.
Duval, E. Far-Infrared and Raman Vibrational Transitions of a Solid Sphere: Selection Rules. Phys. Rev. B
1992
,46, 5795. [CrossRef]
[PubMed]
25.
Zeng, Y.; Larson, S.B.; Heitsch, C.E.; McPherson, A.; Harvey, S.C. A Model for the Structure of Satellite Tobacco Mosaic Virus.
J. Struct. Biol. 2012,180, 110–116. [CrossRef] [PubMed]
26.
Freddolino, P.L.; Arkhipov, A.S.; Larson, S.B.; McPherson, A.; Schulten, K. Molecular Dynamics Simulations of the Complete
Satellite Tobacco Mosaic Virus. Structure 2006,14, 437–449. [CrossRef]
27.
Jung, J.; Mori, T.; Kobayashi, C.; Matsunaga, Y.; Yoda, T.; Feig, M.; Sugita, Y. Wiley Interdiscip. Rev. Comput. Mol. Sci.
2015
,5, 310.
[CrossRef]
28.
Kumar, S.; Sun, Y.; Kalé, L.V. Acceleration of an Asynchronous Message Driven Programming Paradigm on IBM Blue Gene/Q.
In Proceedings of the 2013 IEEE 27th International Symposium on Parallel and Distributed Processing, Cambridge, MA, USA, 20–24 May
2013; IEEE: Piscataway, NJ, USA, 2013; pp. 689–699.
29.
Solov’yov, A.V.; Surdutovich, E.; Scifoni, E.; Mishustin, I.; Greiner, W. Physics of Ion Beam Cancer Therapy: A Multiscale
Approach. Phys. Rev. E 2009,79, 011909. [CrossRef]
30.
Surdutovich, E.; Yakubovich, A.V.; Solov’yov, A.V. Biodamage via Shock Waves Initiated by Irradiation with Ions. Sci. Rep.
2013
,
3, 1289. [CrossRef]
31.
Surdutovich, E.; Solov’yov, A. Multiscale Physics of Ion-Beam Cancer Therapy. In Proceedings of the Journal of Physics: Conference
Series; IOP Publishing: Bristol, UK, 2012; Volume 373, p. 012001.
32.
Adhikari, U.; Goliaei, A.; Berkowitz, M.L. Mechanism of Membrane Poration by Shock Wave Induced Nanobubble Collapse:
A Molecular Dynamics Study. J. Phys. Chem. B 2015,119, 6225–6234. [CrossRef]
33.
Koshiyama, K.; Kodama, T.; Yano, T.; Fujikawa, S. Molecular Dynamics Simulation of Structural Changes of Lipid Bilayers
Induced by Shock Waves: Effects of Incident Angles. Biochim. Biophys. Acta (BBA)-Biomembr. 2008,1778, 1423–1428. [CrossRef]
34.
Phillips, J.C.; Hardy, D.J.; Maia, J.D.C.; Stone, J.E.; Ribeiro, J.V.; Bernardi, R.C.; Buch, R.; Fiorin, G.; Hénin, J.; Jiang, W.; et al.
Scalable Molecular Dynamics on CPU and GPU Architectures with NAMD. J. Chem. Phys. 2020,153, 044130. [CrossRef]
35.
Holian, B.L.; Lomdahl, P.S. Plasticity Induced by Shock Waves in Nonequilibrium Molecular-Dynamics Simulations. Science
1998
,
280, 2085–2088. [CrossRef] [PubMed]
36. Humphrey, W.; Dalke, A.; Schulten, K. VMD–Visual Molecular Dynamics. J. Mol. Graph. 1996,14, 33–38. [CrossRef]
37. Lamb, H. On the Vibrations of an Elastic Sphere. Proc. Lond. Math. Soc. 1881,1, 189–212. [CrossRef]
38.
Murray, D.B.; Saviot, L. Phonons in an Inhomogeneous Continuum: Vibrations of an Embedded Nanoparticle. Phys. Rev. B
2004
,
69, 094305. [CrossRef]
39.
MacKerell Jr, A.D.; Bashford, D.; Bellott, M.; Dunbrack Jr, R.L.; Evanseck, J.D.; Field, M.J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; et al.
All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins. J. Phys. Chem. B
1998
,102, 3586–3616.
[CrossRef]
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Prior opto-mechanical techniques to measure vibrational frequencies of viruses work on large ensembles of particles, whereas, in this work, individually trapped viral particles were studied. Double nanohole (DNH) apertures in a gold film were used to achieve optical trapping of one of the smallest virus particles yet reported, PhiX174, which has a diameter of 25 nm. When a laser was focused onto these DNH apertures, it created high local fields due to plasmonic enhancement, which allowed stable trapping of small particles for prolonged periods at low powers. Two techniques were performed to characterize the virus particles. The particles were sized via an established autocorrelation analysis technique, and the acoustic modes were probed using the extraordinary acoustic Raman (EAR) method. The size of the trapped particle was determined to be 25 ± 3.8 nm, which is in good agreement with the established diameter of PhiX174. A peak in the EAR signal was observed at 32 GHz, which fits well with the predicted value from elastic theory.
Article
Full-text available
Femtosecond (fs) pulsed laser irradiation techniques have attracted interest as a photonic approach for the selective inactivation of virus contaminations in biological samples. Conventional pulsed laser approaches require, however, relatively long irradiation times to achieve a significant inactivation of virus. In this study, we investigate the enhancement of the photonic inactivation of Murine Leukemia Virus (MLV) via 805 nm femtosecond pulses through gold nanorods whose localized surface plasmon resonance overlaps with the excitation laser. We report a plasmonically enhanced virus inactivation, with greater than 3.7-log reduction measured by virus infectivity assays. Reliable virus inactivation was obtained for 10 s laser exposure with incident laser powers ≥0.3 W. Importantly, the fs-pulse induced inactivation was selective to the virus and did not induce any measurable damage to co-incubated antibodies. The loss in viral infection was associated with reduced viral fusion, linking the loss in infectivity with a perturbation of the viral envelope. Based on the observations that physical contact between nanorods and virus particles was not required for viral inactivation and that reactive oxygen species (ROS) did not participate in the detected viral inactivation, a model of virus inactivation based on plasmon enhanced shockwave generation is proposed.
Article
Full-text available
Virus is known to resonate in the confined-acoustic dipolar mode with microwave of the same frequency. However this effect was not considered in previous virus-microwave interaction studies and microwave-based virus epidemic prevention. Here we show that this structure-resonant energy transfer effect from microwaves to virus can be efficient enough so that airborne virus was inactivated with reasonable microwave power density safe for the open public. We demonstrate this effect by measuring the residual viral infectivity of influenza A virus after illuminating microwaves with different frequencies and powers. We also established a theoretical model to estimate the microwaves power threshold for virus inactivation and good agreement with experiments was obtained. Such structure-resonant energy transfer induced inactivation is mainly through physically fracturing the virus structure, which was confirmed by real-time reverse transcription polymerase chain reaction. These results provide a pathway toward establishing a new epidemic prevention strategy in open public for airborne virus.
Article
Full-text available
GENESIS (Generalized‐Ensemble Simulation System) is a new software package for molecular dynamics ( MD ) simulations of macromolecules. It has two MD simulators, called ATDYN and SPDYN . ATDYN is parallelized based on an atomic decomposition algorithm for the simulations of all‐atom force‐field models as well as coarse‐grained Go‐like models. SPDYN is highly parallelized based on a domain decomposition scheme, allowing large‐scale MD simulations on supercomputers. Hybrid schemes combining OpenMP and MPI are used in both simulators to target modern multicore computer architectures. Key advantages of GENESIS are (1) the highly parallel performance of SPDYN for very large biological systems consisting of more than one million atoms and (2) the availability of various REMD algorithms (T‐ REMD , REUS , multi‐dimensional REMD for both all‐atom and Go‐like models under the NVT , NPT , NPAT , and NPγT ensembles). The former is achieved by a combination of the midpoint cell method and the efficient three‐dimensional Fast Fourier Transform algorithm, where the domain decomposition space is shared in real‐space and reciprocal‐space calculations. Other features in SPDYN , such as avoiding concurrent memory access, reducing communication times, and usage of parallel input/output files, also contribute to the performance. We show the REMD simulation results of a mixed ( POPC / DMPC ) lipid bilayer as a real application using GENESIS . GENESIS is released as free software under the GPLv2 licence and can be easily modified for the development of new algorithms and molecular models. WIREs Comput Mol Sci 2015, 5:310–323. doi: 10.1002/wcms.1220 This article is categorized under: Structure and Mechanism > Computational Biochemistry and Biophysics Computer and Information Science > Computer Algorithms and Programming Molecular and Statistical Mechanics > Molecular Dynamics and Monte-Carlo Methods
Article
Recent advances in photonics technologies pushed optical microscopy towards new horizons in materials characterization. In this framework, Brillouin microscopy emerged as an innovative method to provide images of materials with mechanical contrast without any physical contact, but exploiting the light-matter interaction. Brillouin imaging holds a great promise; to allow mechanical analysis inside soft and heterogeneous materials, addressing the pivotal role played by viscoelastic properties in the physiology and pathology of living tissues and cells. Nevertheless, extending the approach of Brillouin imaging to characterize elastic heterogeneities of micro and nanostructured samples is especially challenging, and it poses a critical question about the actual spatial resolution reachable in the mechanical characterization. We focus this critical review on the key quantities that define the spatial resolution in the Brillouin scattering process, and we highlight that not only the optical focalization of the light, but also the acoustic excitations present in the material, influence the information collected from a sample by Brillouin imaging. Referring to the body of knowledge gained in the field of material science, we review new results and recently obtained progresses in the more unexplored context of life science. In future developments, a comprehensive strategy to tackle both the acoustic and optical aspects of the measurement will be required to maximize the efficacy of the technique.
Article
NAMD is a molecular dynamics program designed for high-performance simulations of very large biological objects on CPU- and GPU-based architectures. NAMD offers scalable performance on petascale parallel supercomputers consisting of hundreds of thousands of cores, as well as on inexpensive commodity clusters commonly found in academic environments. It is written in C++ and leans on Charm++ parallel objects for optimal performance on low-latency architectures. NAMD is a versatile, multipurpose code that gathers state-of-the-art algorithms to carry out simulations in apt thermodynamic ensembles, using the widely popular CHARMM, AMBER, OPLS, and GROMOS biomolecular force fields. Here, we review the main features of NAMD that allow both equilibrium and enhanced-sampling molecular dynamics simulations with numerical efficiency. We describe the underlying concepts utilized by NAMD and their implementation, most notably for handling long-range electrostatics; controlling the temperature, pressure, and pH; applying external potentials on tailored grids; leveraging massively parallel resources in multiple-copy simulations; and hybrid quantum-mechanical/molecular-mechanical descriptions. We detail the variety of options offered by NAMD for enhanced-sampling simulations aimed at determining free-energy differences of either alchemical or geometrical transformations and outline their applicability to specific problems. Last, we discuss the roadmap for the development of NAMD and our current efforts toward achieving optimal performance on GPU-based architectures, for pushing back the limitations that have prevented biologically realistic billion-atom objects to be fruitfully simulated, and for making large-scale simulations less expensive and easier to set up, run, and analyze. NAMD is distributed free of charge with its source code at www.ks.uiuc.edu.
Article
The interaction of laser pulses with the Cauliflower mosaic virus (CaMV) in a Tris-HCl pH7.5 buffer is investigated. 20 ns ruby laser pulses are used for the excitation. Spectra of the light passing through the sample and reflected from it are registered with the help of a Fabri-Perot interferometer. Stimulated low-frequency Raman scattering (SLFRS) in a CaMV suspension is registered. The SLFRS frequency shift, conversion efficiency and threshold are measured for the first time, to the best of our knowledge.
Article
In recent years, Zika virus (ZIKV) caused a new pandemic due to its rapid spread and close relationship with microcephaly. As a result, ZIKV has become an obvious global health concern. Information about the fundamental viral features or the biological process of infection remains limited, despite considerable efforts. Meanwhile, the icosahedral shell structure of the mature ZIKV was recently revealed by cryo-electron microscopy. This structural information enabled us to simulate ZIKV. In this study, we analyzed the dynamic properties of ZIKV through simulation from the mechanical viewpoint. We performed normal mode analysis (NMA) for a dimeric structure ZIKV consisting of the envelope proteins and the membrane proteins as a unit structure. By analyzing low-frequency normal modes, we captured intrinsic vibrational motions and defined basic vibrational properties of the unit structure. Moreover, we also simulated the entire shell structure of ZIKV at the reduced computational cost, similar to the case of the unit structure, by utilizing its icosahedral symmetry. From the NMA results, we can not only comprehend the putative dynamic fluctuations of ZIKV but also verify previous inference such that highly mobile glycosylation sites would play an important role in ZIKV. Consequently, this theoretical study is expected to give us an insight on the underlying biological functions and infection mechanism of ZIKV.
Article
We performed coarse-grained molecular dynamics simulations in order to understand the mechanism of membrane poration by shock wave induced nanobubble collapse. Pressure profiles obtained from the simulations show that the shock wave initially hits the membrane and is followed by a nanojet produced by the nanobubble collapse. While in the absence of the nanobubble, the shock wave with an impulse of up to 18 mPa*s does not create a pore in the membrane, in the presence of a nanobubble even a smaller impulse leads to the poration of the membrane. Two-dimensional pressure maps depicting the pressure distributed over the lateral area of the membrane reveal the differences between these two cases. In the absence of a nanobubble, shock pressure is evenly distributed along the lateral area of the membrane, while in the presence of a nanobubble an unequal distribution of pressure on the membrane is created, leading to the membrane poration. The size of the pore formed depends on both shock wave velocity and shock wave duration. The results obtained here show that these two properties can be tuned to make pores of various sizes.