logo

Molecular rotation-caused autocorrelation

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Molecular rotation-caused autocorrelation

Yu-Wei Guo
Jing-Yu Qin
Jian-Hua Hu
Ji-Hua Cao
Zhi Zhu
Chun-Lei Wang
Nuclear Science and TechniquesVol.31, No.6Article number 53Published in print 01 Jun 2020Available online 14 May 2020
37500

The finite autocorrelation time of thermal noise is crucial to unidirectional transportation on the molecular scale. Therefore, it is important to understand the cause of the intrinsic picosecond autocorrelation time of thermal noise in water. In this work, we use molecular dynamics simulations to compare the autocorrelation behaviors of the thermal noise, hydrogen bonds, and molecular rotations found in water. We found that the intrinsic picosecond autocorrelation time for thermal noise is caused by finite molecular rotation relaxation, in which hydrogen bonds play the role of a bridge. Furthermore, the simulation results show that our method of calculating the autocorrelation of thermal noise, by observing the fluctuating force on an oxygen atom of water, provides additional information about molecular rotations. Our findings may advance the understanding of the anomalous dynamic nanoscale behavior of particles, and the applications of terahertz technology in measuring the structural and dynamical information of molecules in solutions.

Key words: Thermal noiseHydrogen bondRotationMolecular dynamics simulation

behaviors of thermal noise in water

Introduction

Thermal noise, also referred to as thermal fluctuation, is ubiquitous in nature. This noise is significant in almost all physical-biochemical processes, including the movement of molecular motors1-5, adsorption of molecules6,7, molecular transport 8-11, and water evaporation12,13. Thermal noise plays a key role in microscopic measurements14-16. For example, it can be used to measure the spring constant of atomic force microscope cantilevers17. Furthermore, the noises associated with biological processes are important when performing single-particle tracking18. In traditional theories, thermal noise is modeled as Gaussian white noise with a constant power spectrum density (PSD ~ 1/f α, α = 0)19. This treatment may be impractical on the nanoscale, because nanoscale phenomena typically differ from those on the macroscopic and mesoscopic scales. Gong et al. and Aluru et al. discovered the unidirectional transportation of water molecules across a water nanochannel in the presence of an asymmetric potential, in 2007 and 2008, respectively20,21. In 2012, Detcheverry et al. demonstrated that the center of mass for a confined fluid performs a non-Markovian random walk in nanoporous structures, which is inconsistent with its behavior in macrosystems22. Initially, the existence of unidirectional transportation in nanosystems was controversial23-25. However, Wan et al. proved that the unidirectional transportation of molecules at the nanoscale was feasible in the presence of an appropriate asymmetric potential and a finite autocorrelation time of thermal noise26. They further suggested that thermal noise is crucial to unidirectional transportation in asymmetric nanosystems, and that in water it has an autocorrelation time of the order of picoseconds26. In other words, thermal noise in water can generate molecular driving forces in thermally fluctuating conditions. Previous theoretical and experimental studieshave also shown that thermal noise cannot be treated simply as white noise on the molecular scale, it may be colored (PSD ~ 1/f α, α ≠ 0)27-29. In 2014, using simulations modeling different thermostat coupling methods and parameters, we discovered that the autocorrelation time of thermal noise is intrinsic. This means it depends only on the system temperature, and does not need to be evaluated through calculation, with a duration of ~10 ps in water27. In 2016, we proposed (using power spectrum analysis) that the thermal noise of polar solute molecules (such as water, methanol, etc.) in water resembles a 1/f noise distribution (a typical non-white noise) on the nanoscale28. In 2019, Mukherjee et al. further confirmed that the thermal noise of protein, which is sensitive to energy fluctuations, behaves in water as 1/f noise29. Non-white thermal noise appears to be crucial to understanding the structural and dynamical characteristics of molecules or polymers in water. However, our previous work only suggested that the picosecond autocorrelation time of thermal noise may be due to the alteration of the water’s hydrogen bond network27. Detailed research should clarify the cause of the picosecond autocorrelation time of thermal noise.

Rotation is the most basic form of molecular motion. The rotation of water molecules in the hydration layer of solutes influences various biochemical processes, including ion transport, protein folding, and chemical reactions in the solution. For example, by calculating molecular dynamics for a model Cl + CH3ClSN2 reaction in water, Gertner et al. observed that rotations induced the rearrangement of water molecules’ hydrogen bonds and generated the conditions required for reactants to enter a transition state30. Bernal et al. proposed that the rotations of water molecules determine proton transport in water, they found that the motion (including rotation) of neighboring water molecules was a key feature in the necessary solvent reorganization31. Our earlier research (2018) also demonstrated that limiting the rotations of the outermost water molecules can modulate the evaporation of water on a uniformly and completely wetted surface32.

In this study, we initially employ a molecular dynamic simulation33-39 to compare the autocorrelation behavior of thermal noise in bulk water with those of the hydrogen bonds and molecular rotations of water molecules. We discover that the molecular rotations of water molecules have an autocorrelation time of ~10 ps, which is equal to the time of thermal noise autocorrelation in water. Moreover, the rotational autocorrelation function of water molecules closely resembles that of their thermal noise and hydrogen bonds. The variational trends in characteristic time for the autocorrelation functions of molecular rotations, hydrogen bonds, and thermal noise in water correspond to the changes in environment temperature are the same. We then study the effects of molecular rotation on the hydrogen bond network. In further simulations, we partly constrain the rotations of water molecules, by applying a uniform electric field of varying strengths. The simulation results show that restricting the rotations of water molecules can increase the lifetime of hydrogen bonds in water, indicating that molecular rotation determines the stability of the hydrogen bond network. The rotations of water molecules can alter the water’s hydrogen bond network; thus, the picosecond molecular relaxation time determines the finite lifetime of hydrogen bonds in water. In addition, the intrinsic ~10 ps autocorrelation time of thermal noise is thought to originate from alterations to the water’s hydrogen bond network27. Therefore, the detailed cause of the picosecond autocorrelation time of thermal noise can be attributed to the finite rotation relaxation of water molecules in bulk water, where the hydrogen bonds appear to function as a bridge. Moreover, we find that our calculation method (choosing the oxygen atom of water to consider the autocorrelation characteristics of thermal noise) describes the rotational behavior information of molecules. As predicted, when the rotational freedom of the single-file water chain is restricted in a nanotube or nanopore, the thermal noise has a longer autocorrelation time and unidirectional transportation becomes significant20,21,40,41. Finally, we analyze the thermal noise power spectrum in a broad frequency domain for water molecules and found that thermal noise can be modeled as white noise below the terahertz frequency domain (<0.1 THz); however, it must be regarded as non-white noise in the terahertz and higher frequency domains. The dynamic nanoscale processes usually occur on picosecond timescales (corresponding to the terahertz band). Our findings here are significant for the understanding of anomalous nanoscale behaviors, and have implications for the application of terahertz technologies in observing and detecting the structural and dynamical information of molecules in solutions.

Methods

As shown in Figure 1, we placed water molecules into a simulation box (dimensions LX = LY = LZ = 3.0 nm) then began the simulations. All simulations were performed with a canonical ensemble using the classical molecular dynamics programming package GROMACS (version 4.5.5)42. The simulation system used 895 TIP3P rigid water molecules43 because the TIP3P water model is widely used for studying the liquidity and transport properties of water. The interaction parameters of the water molecules were modeled with an OPLS/AA force field. The atomic charges of oxygen and hydrogen were −0.8476 e and 0.4238 e, respectively. The Van der Waals radius (σO) and well depth (εo) for an oxygen atom were 0.316557 nm and 0.605194 kJ/mol, respectively; the Van der Waals radius and well depth for a hydrogen atom were 0 nm and 0 kJ/mol, respectively. Under the OPLS/AA force field, the Lennard–Jones parameters between different atoms were calculated using the geometric average for σij=(σiσj)1/2 and εij=(εiεj)1/2; thus, the non-bonded interaction Enb=ij[qiqje2/rij+4εij(σij12/rij12σij6/rij6)]. The bond length between a hydrogen and oxygen atom in a water molecule was 0.09572 nm. The bond angle of a water molecule (denoted as H−O−H) was 104.52º. Periodic boundary conditions were applied in all directions. Nosé–Hoover thermostats44,45 were used to maintain the system at a set of reference temperatures, which in this study were 285, 300, 315, 330, 345, and 360 K. Nosé–Hoover thermostats depend on two significant coupling parameters: the time constant for coupling τT and the period for coupling tcouple for a particle are replaced by d2ridt2=Fimiζdridt. Here, ri is the coordinate of the i-th atom, Fi is the force exerted on the i-th atom, and mi denotes the mass of the i-th atom. The heat-bath parameter ζ can be expressed as dζdt=4π2τT2T0(TT0), where τT is the coupling time constant, which determines the relaxation time with which the system changes from the instantaneous temperature T to the reference temperature T0. The coupling period tcouple is the time interval between two applications of the temperature-coupling method. Under the Nosé–Hoover approach, these two coupling parameters cause an oscillatory relaxation in the system, resulting in a correct canonical ensemble. The particle mesh Ewald46 integration was used to treat the long-range electrostatic interactions. The simulation time for each system was set as 20 ns with a time step of 2 fs. The trajectory was measured every 0.2 ps (simulation results using different recording-time resolutions can be found in the ESI†), and data from the previous 10 ns were extracted to analyze the characteristics of thermal noise.

Figure 1.
(Color online) The initial simulation state with 895 TIP3P water molecules in a 3.0 × 3.0 × 3.0 nm3 cube (simulation results with different box sizes can be seen in the ESI†). The red and gray spheres represent the oxygen and hydrogen atoms of the water molecules, respectively. The blue lines represent the boundaries of the periodic box in the simulation system.
pic

Results and Discussion

Our simulations (Figure 2) show that the dynamical behaviors (including the thermal noise, hydrogen bonds, and rotations) of water molecules in bulk water have an intrinsic picosecond autocorrelation time, and that the autocorrelation time of thermal noise is closely related to the autocorrelation characteristics of the hydrogen bonds and rotations of the water molecules. Typically, thermal noise, which is also referred to as thermal agitation, is caused by the random motions of particles such as electrons and ions. In molecular dynamics simulations, the thermal noise can be characterized as the randomly fluctuating force that a molecule or atom receives from its surrounding environment. Thus, we used the fluctuating force (f) exerted on the oxygen atoms of water by surrounding water molecules to represent thermal noise in the simulations. The autocorrelation function of thermal noise is

Figure 2.
(Color online) (a) The autocorrelation functions  C(t)  for different quantities of water molecules in bulk water. The solid black, solid blue, and dotted red curves represent the autocorrelation functions of the thermal noise (Cnoise(t)), hydrogen bonds (CHbond(t)), and molecular rotations (Corientation(t)) of water molecules in bulk water, respectively. The inset displays the function using the expression of W(t)=t·C(t), in which the autocorrelation time t (with respect to the maximum value of  W(t)) is the characteristic time of the autocorrelation function. The solid black, solid blue, and dotted red curves represent the functions for the thermal noise, hydrogen bonds, and molecular rotations, respectively. (b) The characteristic time of the autocorrelation function with different temperatures (T), which were fitted using the autocorrelation function described by a single exponential function y=etτ. The solid black, blue, and red curves represent the characteristic times of the thermal noise τnoise, hydrogen bonds τHbond (also referred to as the lifetime of the hydrogen bond), and water molecular rotation τrotation in water at different temperatures, respectively. The error bars of characteristic times for thermal noise, hydrogen bonds, and molecular rotations (with respect to the temperature (285 K ≤ T ≤ 360 K)) did not exceed 0.02 ps.
pic
Cnoise(t)=f(0)f(t)f(0)f(0). (1)

Here, f(0) and f(t) represent the forces exerted on the oxygen atom of a water molecule at times 0 and t, respectively. We first calculated the autocorrelation of thermal noise in bulk water at the reference temperature of 300 K. It can be seen from Fig. 2(a) that Cnoise(t) decays exponentially from 1 to 0 as the autocorrelation time t is increased. The maximum autocorrelation time (the minimum time in which Cnoise(t)≈0) of thermal noise in bulk water was approximately 10 ps. In our previous work, we speculated that the finite correlation time of thermal noise in water was related to the dynamical behavior of the water’s hydrogen bond network27, because water molecules in bulk water maintain their tetrahedral coordination shells by forming hydrogen bonds with other water molecules in the vicinity. Hydrogen bonding is not only the main component of the interaction between a water molecule and its neighboring water molecules, it also strongly influences the physical and chemical properties of water47-50. Hydrogen bonds are likely to form between neighboring water molecules if the O-H bond of one molecule faces the oxygen of the next. In this study, we adopted a commonly used geometric definition of hydrogen bonding, to determine whether or not a hydrogen bond is formed. Specifically, two water molecules were considered to be hydrogen-bonded if the distance between the oxygen atoms of the molecules (O…O) was less than 3.5 Å and, at the same time, the angle between the H-O bond of the water molecule and the connecting line of two oxygen atoms (H―O…O) was less than 30° 51-53. The autocorrelation function expression of a hydrogen bond is

CHbond(t)=h(0)h(t)h(0)h(0), (2)

where h(0) = 1 indicates that two water molecules initially form a hydrogen bond, and h(t) = 1 indicates that the these two water molecules are still bonded at time t; otherwise, h(t) = 0. CH-bond(t) is indeed similar to Cnoise(t), resulting in a comparable autocorrelation time (~ 10 ps) (see also Figure 2(a)). It is worth noting that the intrinsic dynamics of water molecules—for instance, rotations and translations—profoundly affect the lifetime of their hydrogen bonds. Thus, we further studied the autocorrelation of rotations for molecules in bulk water. A water molecule can be modeled as a set of three points, corresponding to a negatively polarized oxygen O and two positively polarized hydrogen H1 and H2. The angle between, and the respective lengths of, the vectors OH1 and OH2 can fluctuate. The water dipole moment vector and the vector identified from the two hydrogen atoms of each water molecule are considered here. Such molecular orientations can be conveniently described with a choice of vectors, defined as d=OH1+OH2|OH1+OH2|. Likewise, the autocorrelation function of molecular orientation is

Corientation(t)=d(0)d(t)d(0)d(0), (3)

where d(0) and d(t) are the dipoles of tagged water molecules at times 0 and t, respectively. As shown in Figure 2(a), it can be seen that the autocorrelation function of thermal noise for water molecules is almost identical with that of water’s molecular orientation. To facilitate a quantitative comparison, we used the single exponential functions y(t)=etτnoise,  y(t)=etτHbond, and y(t)=etτorientation to fit the autocorrelation functions of the thermal noise, the hydrogen bonds, and the orientations of the water molecules, respectively. Here, τnoise  is the characteristic time of the thermal noise autocorrelation,  τHbond  is the characteristic time of the water hydrogen bond autocorrelation (also referred to as the lifetime of the hydrogen bond), and  τorientation  is the characteristic time of the molecular orientation autocorrelation. The values were 1.90, 1.38, and 2.10 ps for the τnoise, τHbond , and τorientation of water molecules in bulk water at 300 K, respectively. It is worth noting that the characteristic time of hydrogen bond autocorrelation was smaller than that of thermal noise or water molecule rotation; this was a result of the hydrogen bond definition (i.e., the O–O distance ≤ 3.5   Å and hydrogen-bond angle ≤ 30°)51-53. Because the interaction between a hydrogen donor and oxygen acceptor remains after breaking of the hydrogen bond, the average lifetime of the hydrogen bond should be smaller than the characteristic time for thermal noise or water molecule rotation. Furthermore, we used W(t)=t·C(t) to model the characteristic time τ, because W(t) has a maximum when t=τ. The inset of Figure 2(a) illustrates that the value of t (with respect to the maximum positions of W(t)) for the thermal noise, hydrogen bonds, and water orientations are almost identical, which further confirms that their respective characteristic times are also almost identical.

The thermal kinetic behavior of molecules is determined by temperature, because the reference temperature of the system determines the velocity distribution of the water molecules undergoing Brownian motion. Molecular velocity affects the collisions between the water molecules and their neighbors. At a certain temperature, the duration of the interaction between neighboring molecules is intrinsic. Therefore, the thermal noise correlation time of water molecules may also depend on the temperature, if it is also intrinsic. Thus, we proceeded to study the effect of bulk water temperature on the characteristic times of the thermal noise τnoise, the hydrogen bonds τHbond, and the molecular rotations τrotation. It can be seen from Figure 2(b) that τnoise decreases significantly under an increase in temperature, which confirms that the autocorrelation time of thermal noise in bulk water depends on the environment temperature. Interestingly, τrotation  and τHbond also decrease as the temperature increases. This phenomenon may be attributable to fact that an increase in the reference temperature can accelerate the molecular rotational velocity of water, which causes the finite relaxation time of molecular rotation to decease. Meanwhile, the molecular rotation can affect the neighboring hydrogen bond network; thus, the hydrogen bond lifetime should decrease as the temperature increases. Notably, the changing trends of τnoise with respect to temperature are highly consistent. These facts indicate that the cause of the intrinsic picosecond autocorrelation time of thermal noise in water is the finite molecular rotation relaxation, in which the hydrogen bonds act as a bridge.

The hydrogen bond is the bridge between molecular rotation and thermal noise, which suggests that limiting the molecular rotation will prolong the dynamic adjustment time of the hydrogen bond network. To study the effects of molecular rotation on the alteration of the neighboring hydrogen bond network, we partly constrain the rotations of water molecules by applying a uniform electric field. Thus, we conducted several molecular dynamics simulations with an electric field intensity of 1, 2, 3, 4, and 5 V/nm. As shown in Fig. 3, the value of τHbond is 1.38 ± 0.01 ps in the absence of an external electric field (i.e., E = 0 V/nm). In contrast, the lifetime of the hydrogen bonds in water increases with the electric field increase, its value is 1.39 ± 0.02 ps, 1.43 ± 0.02 ps, 1.48 ± 0.02 ps, 1.61 ± 0.02 ps, and 1.78 ± 0.03 ps for E = 1, 2, 3, 4, and 5 V/nm in the presence of an external electric field, respectively. This confirms that the lifetime of a hydrogen bond increases with the constraint strength under an increase of molecular rotation. Because the water molecule dipole is constrained by the electric field, the molecular rotation relaxation time increases as the constraint strength of molecular rotation increases. Combined with the fact that the intrinsic picosecond autocorrelation time of thermal noise stems directly from the alteration of the hydrogen bond network27, we can conclude that the origin of the intrinsic picosecond autocorrelation time of thermal noise is the finite relaxation time of molecular rotations in water.

Figure 3.
The black squares with error bars represent the lifetimes of hydrogen bonds formed by water molecules in bulk water when applying different uniform electric fields to partly constrain the molecular rotation.
pic

To confirm the relationship between the water molecule rotations and the thermal noise further, we analyzed the dynamic trajectories of water molecules in bulk water and calculated the angle θ between the molecular dipole and thermal noise vectors. Figure 4(a) shows the angle between the thermal noise and water molecule dipole vectors with respect to the simulation time t, where we rotated the water dipole vector in a clockwise direction until it aligned with the thermal noise vector. The angle can thus be changed from 0º to 360º. From the statistics of the time-dependent angle in Fig. 4(a), we found that the angle θ between the molecular dipole vector and the noise vector predominantly fluctuated around 180º, as indicated by the red line in Fig. 4(a). Our subsequent calculations (Fig. 4(b)), based on the probability distribution of the angle θ between the molecular dipole and noise vectors, show two peaks located at 156.9º and 213.1º. This implies that the orientation of thermal noise is almost identical with that of the molecular dipole. In our previous work27, we speculated that the intrinsic picosecond autocorrelation time of thermal noise in water is directly derivable from the hydrogen bond network alteration. Moreover, the molecular rotation of water can induce the generation and fracture of hydrogen bonds and the water dipole vector is aligned with the thermal noise vector. These facts further confirm that the thermal noise of water molecules depends on the water molecule’s rotations.

Figure 4.
The change and distribution of the angle θ between the molecular dipole and noise vectors, in the clockwise direction; here, we used the molecular dipole as the initial vector. (a) The angle θ between the molecular dipole and noise vectors with respect to the simulation time t. The dashed gray lines represent the upper and lower limits of error for the value of the time-dependent angles. (b) The probability distribution of the angle θ between the molecular dipole and noise vectors. Angles are measured in degrees.
pic

Thermal noise is a vector, which implies that the autocorrelation of thermal noise contains magnitude and direction information. Because the autocorrelation function is normalized, the magnitude of autocorrelation is averaged, thus the orientation plays a crucial role. However, the thermal noise vector is aligned with that of molecular rotation, which entails that the thermal noise and water dipole have near-identical autocorrelation functions. This means that our method of computing the fluctuating force on oxygen atoms in water26-28 contains information pertaining to the water’s dynamical behavior. Specifically, it characterizes the autocorrelation behavior of the water dipole. It is worth noting that molecules are treated as spherical particles in the traditional theory of Brownian motion; thus, the traditional theory only treats the randomly fluctuating force (that of the surrounding environment on the particle) as thermal noise. As a result, it can only represent translational information of the dynamical molecular behavior. However, the structure of a molecule is directly related to its rotation, and the rotations of the molecule directly affects its dynamic properties. Therefore, molecules cannot be simply treated as spherical particles on the nanoscale, and considering the rotations of molecules can better explain the dynamical behavior of molecules at these scales.

In traditional theory, the thermal fluctuation (thermal noise) is treated or modeled as white noise. White noise is a specific type of noise that describes a random signal or process whose autocorrelation time is zero and whose power spectral density is constant. We have detailed the cause of the intrinsic thermal noise autocorrelation time in bulk water and know that it is on the picosecond scale, indicating that thermal noise is not white noise on the nanoscale. Here, we further study the non-white behavior of thermal noise. The power spectrum function is widely used to analyze fluctuation properties in a random process54; it describes the distribution of noise power with respect to frequency, and with this we can indicate the periodicities that dominate the whole process of fluctuation and distinguish the colors of the noise55-57. The spectral power of stationary stochastic processes is the Fourier transform of the autocorrelation function. Because the autocorrelation function of thermal noise Cnoise(t) is time-reversible (in other words, Cnoise(t) is an even function), the power spectrum of thermal noise can be expressed as

S(ω)=1tmtmtmCnoise(t)eiωtdt=2tm0tmCnoise(t)cos(ωt)dt. (4)

Here, tm is the maximum time considered for Cnoise(t), which was here set as 1 ns. Figure 5(a) shows the autocorrelation function of thermal noise in water, the thermal noise autocorrelation time is ~ 10 ps. We used the exponential function R(t)=e-t/1.90 to fit the autocorrelation function of thermal noise. It can further be seen from Fig. 5(a) that R(t) fits the autocorrelation function of thermal noise well, which means that the characteristic time of thermal noise is 1.90 ps. Figure 5(b) shows the spectral density of the thermal noise power spectrum, which was derived from the data in Fig. 5(a) using Eq.(4). The spectral density is constant when the frequency ω is less than 1 × 1011 Hz, it then decreases as frequency increases from 1 × 1011 Hz to 2.5 × 1012 Hz. Because the Fourier transform of y(t)=e-at is S(ω)=2aa2+ω2, S(ω) should describe the power spectrum of thermal noise well. From Fig. 5(b), it can be seen that the spectral power of thermal noise indeed approximates to the function of S(ω)=2aa2+ω2, where a=1τnoise = 5.26 × 1011 with the frequency domain ω ranging from 1 × 109 Hz to 2.5 × 1012 Hz (see the inset of Figure 5(b) for the spectral density of thermal noise in a wider frequency range, obtained from S(ω)=2aa2+ω2). According to the expression for S(ω), we know that S(ω) approximates to a constant of 2/a when ω< 0.1 THz, because ω is considerably smaller than a in this range. When ω > 0.1 THz, ω is comparable to a, and S(ω) is no longer constant. This suggests that thermal noise in bulk water takes the form of white noise in the lower frequency domain and non-white noise at terahertz and higher frequency domains (see the ESI†for further discussion). Because the particle dynamics we observed on macroscopic or mesoscopic scales are low frequency, it may be feasible to treat the thermal noise as white noise on these scales. This treatment may not be appropriate at the nanoscale, because the energy of a nanoscale system is expressed by several kBT ranges (kB represents the Boltzmann constant); thus, the magnitude of the thermal fluctuations is comparable to the thermodynamic quantities to which the system is subjected. The molecular or atomic collisions occur within a timescale of picoseconds or faster; this corresponds to high-frequency dynamics. Consequently, thermal noise is non-white at nanoscales. Thus, the biased motion of particles resulting from thermal noise may be possible in nanoscale systems if (a) spatial inversion asymmetry holds58 and (b) we can prevent distorted measurement in high time-resolution experimental techniques by distinguishing the noise color18.

Figure 5.
(Color online) The autocorrelation function of thermal noise in the time and frequency domains, respectively. (a) The solid black curve represents the autocorrelation function of thermal noise within time t from 0 ps to 40 ps. The dashed red curve represents the fitting function R(t)=etτnoise of the thermal noise autocorrelation function, where τnoise = 1.90 ps is the characteristic time of thermal noise. (b) The solid black curve represents the spectral density of thermal noise in bulk water with respect to the frequency ω, which was calculated from the thermal noise autocorrelation function within the frequency domain from 1×109 Hz to 2.5×1012 Hz. The red curve is the fitting function of S(ω)=2aa2+ω2, where a=11.90×1012= 5.26 × 1011; the inset illustrates the fitting function of S(ω)with a wider frequency domain (from 1 × 103 Hz to 1 × 1016 Hz).
pic

Conclusion

We used molecular dynamic simulations to investigate the relationship between the autocorrelation behaviors of thermal noise, hydrogen bonds, and the molecular rotations of water molecules in bulk water. The simulation results showed that the thermal noise autocorrelation function of water molecules in bulk water is very similar to the autocorrelation functions of the hydrogen bonds and molecular rotations of water molecules. Moreover, the characteristic autocorrelation times of the thermal noise, hydrogen bonds, and water rotations exhibited the same behaviors under changes in environment temperature. Further study on the thermal noise power spectrum of water molecules shows that thermal noise on the nanoscale cannot be regarded as white noise. We found that the cause of the intrinsic picosecond autocorrelation time for thermal noise is the finite (picosecond) molecular rotation relaxation in water. The underlying mechanism arises because the rotation of water molecules induces the alteration of their hydrogen bond network, indicating that the finite molecular rotation relaxation is the origin of the finite autocorrelation time of hydrogen bonds in water. Meanwhile, the thermal fluctuating force occurs from the alteration of the hydrogen bond network in water27, meaning the hydrogen bond plays the role of a bridge. Moreover, we found that the orientation of the molecular dipole is nearly opposite that of the thermal fluctuating force exerted on the oxygen atom of the water molecule. This result indicates that our calculation method for the thermal noise, that of choosing the oxygen atom of the water molecule to consider the autocorrelation time of thermal noise in water, provides additional information about the molecular rotation behavior.

The finite water molecular rotation relaxation time causes the finite autocorrelation time of thermal noise in water. Therefore, the thermal noise under confined conditions (such as a water channel of nanometer size) might have a longer autocorrelation time when the rotational freedom is restricted. Therefore, the unique dynamical behavior of a single-file water chain confined in a nanotube or other porous material of similar size, in the presence of an appropriate asymmetric potential, may be made more understandable20,21. It should also be noted that many physical and biochemical processes occur on the picosecond timescale, and that this timescale corresponds to the terahertz 58 frequency domain. Accordingly, our findings are expected to further understanding by using terahertz technology to observe and detect the unique physical and biochemical processes occurring on the nanoscale, and to promote the applications of terahertz technology in biomedical science.

References
1. A. Efremov and Z. Wang,

Universal optimal working cycles of molecular motors

. Phys. Chem. Chem. Phys. 13, 6223-6233 (2011). https://doi.org/10.1039/C0CP02118K
Baidu ScholarGoogle Scholar
2. M. Alvarez-Pérez, S. M. Goldup, D. A. Leigh et al.,

A chemically-driven molecular information ratchet

. J. Am. Chem. Soc. 130, 1836-1838 (2008). https://doi.org/10.1021/ja7102394
Baidu ScholarGoogle Scholar
3. A. Yildiz and P. R. Selvin,

Fluorescence imaging with one nanometer accuracy:  Application to molecular motors

. Acc. Chem. Res. 38, 574-582 (2005). https://doi.org/10.1021/ar040136s
Baidu ScholarGoogle Scholar
4. T. E. Kelly, R. A. Silva, H. D. Silva, et al.,

A rationally designed prototype of a molecular motor

. J. Am. Chem. Soc. 122, 6935-6949 (2000). https://doi.org/10.1021/ja001048f
Baidu ScholarGoogle Scholar
5. B. Wand, L. Vuković and P. Král,

Nanoscale rotary motors driven by electron tunneling

. Phys. Rev. Lett. 101, 186808 (2008). https://doi.org/10.1103/PhysRevLett.101.186808
Baidu ScholarGoogle Scholar
6. J. Liu, G. Shi, P. Guo, et al.,

Blockage of water flow in carbon nanotubes by ions due to interactions between cations and aromatic rings

. Phys. Rev. Lett. 115, 164502 (2015). https://doi.org/10.1103/PhysRevLett.115.164502
Baidu ScholarGoogle Scholar
7. F. Ganazzoli and G. Raffaini,

Computer simulation of polypeptideadsorption on model biomaterials

. Phys. Chem. Chem. Phys. 7, 3651-3663 (2005). https://doi.org/10.1039/B506813D
Baidu ScholarGoogle Scholar
8. J. Su, K. Yang and H. Guo,

Asymmetric transport of water molecules through a hydrophobic conical channel

. RSC Adv. 4, 40193-40198 (2014). https://doi.org/10.1039/C4RA07034H
Baidu ScholarGoogle Scholar
9. E. R. Cruz-Chu, A. Aksimentiev and K. Schulten,

Ionic current rectification through silica nanopores

. J. Phys. Chem. C 113, 1850-1862 (2009). https://doi.org/10.1021/jp804724p
Baidu ScholarGoogle Scholar
10. Y. Wand, Y. Zhao and J. Huang,

Giant pumping of single-file water molecules in a carbon nanotube

. J. Phys. Chem. B 115, 13275-13279 (2011). https://doi.org/10.1021/jp2069557
Baidu ScholarGoogle Scholar
11. C. Zhu, H. Li and S. Meng,

Transport behavior of water molecules through two-dimensional nanopores

. J. Chem. Phys. 141, 18C528 (2014). https://doi.org/10.1063/1.4898075
Baidu ScholarGoogle Scholar
12. X. Nan, Y. W. Guo, R. Z. Wan.

Effect of Na and Cl ions on water evaporation on graphene oxide

. Nuc. Sci. Tech. 30, 122 (2019). https://doi.org/10.1007/s41365-019-0646-7
Baidu ScholarGoogle Scholar
13. C. Caleman and D. van der Spoel,

Evaporation from water clusters containing singly charged ions

. Phys. Chem. Chem. Phys. 9, 5105-5111 (2007). https://doi.org/10.1039/B706243E
Baidu ScholarGoogle Scholar
14. R. Levy and M. Maaloum,

Measuring the spring constant of atomic force microscope cantilevers: thermal fluctuations and other methods

. Nanotechmology 13, 33 (2002). https://doi.org/10.1088/0957-4484/13/1/307
Baidu ScholarGoogle Scholar
15. A. Burtzlaff, A. Weismann, M. Brandbyge et al.,

Shot noise as a probe of spin-polarized transport through single atoms

. Phys. Rev. Lett. 114, 016602 (2015). https://doi.org/10.1103/PhysRevLett.114.016602
Baidu ScholarGoogle Scholar
16. Y. Hovav, I. Kaminker, D. Shimon, et al.,

The electron depolarization during dynamic nuclear polarization: measurements and simulations

. Phys. Chem. Chem. Phys. 17, 226-244 (2015). https://doi.org/10.1039/C4CP03825H
Baidu ScholarGoogle Scholar
17. H. J. Butt and M. Jaschke,

Calculation of thermal noise in atomic force microscopy

. Nanotechmology 6, 1 (1995). https://doi.org/10.1088/0957-4484/6/1/001
Baidu ScholarGoogle Scholar
18. J. Siódmiak and P. Beldowski,

Hyaluronic acid dynamics and its interaction with synovial fluid components as a source of the color noise

. Fluct. Noise Lett. 18, 1940013 (2019). https://doi.org/10.1142/S0219477519400133
Baidu ScholarGoogle Scholar
19. P. Hanggi and P. Jung,

Colored noise in dynamical systems

. Adv. Chem. Phys. 89, 239-326 (1995). https://doi.org/10.1002/9780470141489.ch4
Baidu ScholarGoogle Scholar
20. X. J. Gong, J. Y. Li, H. J. Lu, et al.,

A charge-driven molecular water pump

. Nat. Nanotechnol. 2, 709-712 (2007). https://doi.org/10.1038/nnano.2007.320
Baidu ScholarGoogle Scholar
21. S. Joseph and N. R. Aluru,

Pumping of confined water in carbon nanotubes by rotation-translation coupling

. Phys. Rev. Lett. 101, 064502 (2008). https://doi.org/10.1103/PhysRevLett.101.064502
Baidu ScholarGoogle Scholar
22. F. Detcheverry and L. Bocquet,

Thermal fluctuations in nanofluidic transport

. Phys. Rev. Lett. 109, 024501 (2012). https://doi.org/10.1103/PhysRevLett.109.024501
Baidu ScholarGoogle Scholar
23. D. J. Bonthuis, K. Falk, C. N. Kaplan, et al.,

Comment on “pumping of confined water in carbon nanotubes by rotation-translation coupling”

. Phys. Rev. Lett. 105, 209401 (2010). https://doi.org/10.1103/PhysRevLett.105.209401
Baidu ScholarGoogle Scholar
24. D. J. Bonthuis, D. Horinek, L. Bocquet et al.,

Electrokinetics at aqueous interfaces without mobile charges

. Langmuir 26, 12614-12625 (2010). https://doi.org/10.1021/la9034535
Baidu ScholarGoogle Scholar
25. M. E. Suk and N. R. Aluru,

Suk and Aluru Reply

. Phys. Rev. Lett. 105, 209402 (2010). https://doi.org/10.1103/PhysRevLett.105.209402
Baidu ScholarGoogle Scholar
26. R. Z. Wan, J. Hu and H. P. Fang,

Asymmetric transportation induced by thermal noise at the nanoscale

. Sci. China. Phys. Mech. 55, 751-756 (2012). https://doi.org/10.1007/s11433-012-4695-8
Baidu ScholarGoogle Scholar
27. Z. Zhu, N. Sheng, R. Z. Wan et al.,

Intrinsic autocorrelation time of picoseconds for thermal noise in water

. J. Phys. Chem. A 118, 8936-8941 (2014). https://doi.org/10.1021/jp5009785
Baidu ScholarGoogle Scholar
28. Z. Zhu, N. Sheng, H. P. Fang, et al.,

Colored spectrum characteristics of thermal noise on the molecular scale

. Phys. Chem. Chem. Phys. 18, 30189-30195 (2016). https://doi.org/10.1039/C6CP04433F
Baidu ScholarGoogle Scholar
29. S. Mukherjee, S. Mondal and B. Bagchi,

Mechanism of solvent control of protein dynamics

. Phys. Rev. Lett. 122, 058101 (2019). https://doi.org/10.1103/PhysRevLett.122.058101
Baidu ScholarGoogle Scholar
30. B. J. Gertner, R. M. Whitnell, K. R. Wilson et al.,

Activation to the transition state: reactant and solvent energy flow for a model SN2 reaction in water

. J. Am. Chem. Soc. 113, 74-87 (1991). https://doi.org/10.1021/ja00001a014
Baidu ScholarGoogle Scholar
31. J. D. Bernal and R. H. Fowler,

A Theory of Water and Ionic Solution, with Particular Reference to Hydrogen and Hydroxyl Ions

. J. Chem. Phys. 1, 515-548 (1933). https://doi.org/10.1063/1.1749327
Baidu ScholarGoogle Scholar
32. Y. W. Guo and R. Z. Wan,

Evaporation of nanoscale water on a uniformly complete wetting surface at different temperatures

. Phys. Chem. Chem. Phys. 20, 12272-12277 (2018). https://doi.org/10.1039/C8CP00037A
Baidu ScholarGoogle Scholar
33. C. N. Peng,

The effects of hydrogen on the helium behavior in palladium

. Nucl. Sci. Tech. 27, 106 (2016). https://doi.org/10.1007/s41365-016-0115-5
Baidu ScholarGoogle Scholar
34. W. Xu, Y. S. Tu, C. L. Wang et al.,

Water transport through T-shaped carbon nanotubes

. Nucl. Sci. Tech. 22, 307-310 (2011). https://doi.org/10.13538/j.1001-8042/nst.22.307-310
Baidu ScholarGoogle Scholar
35. X. Ren, B. Zhou and C. Wang,

Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer

. Nucl. Sci. Tech. 23, 252-256 (2012). https://doi.org/10.13538/j.1001-8042/nst.23.252-256
Baidu ScholarGoogle Scholar
36. L. W. Gao, X. B. Xia, X. Q. Xu, et al.,

Immobilization of radioactive fluoride waste in aluminophosphate glass: a molecular dynamics simulation

. Nuc. Sci. Tech. 29, 92 (2018). https://doi.org/10.1007/s41365-018-0443-8
Baidu ScholarGoogle Scholar
37. X. C. Nie, B. Zhou, C. L. Wang, et al.,

Wetting behaviors of methanol, ethanol, and propanol on hydroxylated SiO2 substrate

. Nucl. Sci. Tech. 29, 18 (2018). https://doi.org/10.1007/s41365-018-0364-6
Baidu ScholarGoogle Scholar
38. C. L. Zhao, W. Z. Sun, X. D. Lv, et al.,

Incident angle effect on F+ ions interaction with β-SiC: Molecular dynamics simulation

. Nuclear Techniques 34, 1 (2011). https://d.old.wanfangdata.com.cn/Periodical/hjs201101012 (in Chinese)
Baidu ScholarGoogle Scholar
39. Y. F. Ding, Z. B. Zhang, X. Z. Ke, et al.,

Investigation on single carbon atom transporting through the single-walled carbon nanotube by MD simulation

. Nuclear Techniques 28, 4 (2005). https://doi.org/10.3321/j.issn:0253-3219.2005.04.012 (in Chinese)
Baidu ScholarGoogle Scholar
40. N. Sheng, Y. S. Tu, P. Guo, et al.,

Asymmetrical free diffusion with orientation-dependence of molecules in finite timescales

. Sci. China. Phys. Mech. 56, 1047-1052 (2013). https://doi.org/10.1007/s11433-013-5081-x
Baidu ScholarGoogle Scholar
41. J. Y. Su and H. X. Guo,

Control of Unidirectional Transport of Single-File Water Molecules through Carbon Nanotubes in an Electric Field

. Acs Nano 5, 351-359 (2011). https://doi.org/10.1021/nn1014616
Baidu ScholarGoogle Scholar
42. S. Pronk, S. Pall, R. Schulz, et al.,

GROMACS 4.5: a high-throughput and highly parallel open source molecular simulation toolkit

. Bioinformatics 29, 845-854 (2013). https://doi.org/10.1093/bioinformatics/btt055
Baidu ScholarGoogle Scholar
43. W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, et al.,

Comparison of simple potential functions for simulating liquid water

. J. Chem. Phys. 79, 926-935 (1983). https://doi.org/10.1063/1.445869
Baidu ScholarGoogle Scholar
44. S. Nose,

A molecular dynamics method for simulations in the canonical ensemble

. Mol. Phys. 52, 255-268 (1984). https://doi.org/10.1080/00268978400101201
Baidu ScholarGoogle Scholar
45. W. G. Hoover,

Canonical dynamics: Equilibrium phase-space distributions

. Phys. Rev. A 31, 1695-1697 (1985). https://doi.org/10.1103/PhysRevA.31.1695
Baidu ScholarGoogle Scholar
46. T. Darden, D. York and L. Pedersen,

Particle mesh Ewald: An N⋅log(N) method for Ewald sums in large systems

. J. Chem. Phys. 98, 10089-10092 (1993). https://doi.org/10.1063/1.464397
Baidu ScholarGoogle Scholar
47. W. Nadler and T. Krausche,

Universality in hydrogen-bond networks

. Phys. Rev. A 44, 7888-7890 (1991). https://doi.org/10.1103/PhysRevA.44.R7888
Baidu ScholarGoogle Scholar
48. L. J. Zhang, W. Jian, L. Yi, et al.,

A novel water layer structure inside nanobubbles at room temperature

. Nucl. Sci. Tech. 25, 060503 (2014). https://doi.org/10.13538/j.1001-8042/nst.25.060503
Baidu ScholarGoogle Scholar
49. M. X. He, M. Li, Z. Tian, et al.,

Terahertz spectral properties of melamine and its deuterated isotope, melamine-d_6

. Nucl. Sci. Tech. 23, 209-214 (2012). https://doi.org/10.13538/j.1001-8042/nst.23.209-214
Baidu ScholarGoogle Scholar
50. S. J. Shao, P. Guo, L. Zhao et al.,

Ordered water monolayer on ionic model substrates studied by molecular dynamics simulations

. Nucl. Sci. Tech. 25, 020502 (2014). https://doi.org/10.13538/j.1001-8042/nst.25.020502
Baidu ScholarGoogle Scholar
51. Z. Zhu, H. K. Guo, X. K. Jiang, et al.,

Reversible hydrophobicity-hydrophilicity transition modulated by surface curvature

. J. Phys. Chem. Lett. 9, 2346-2352 (2018). https://doi.org/10.1021/acs.jpclett.8b00749
Baidu ScholarGoogle Scholar
52. Z. Zhu, C. Chang, Y. S. Shu, et al.,

Transition to a superpermeation phase of confined water induced by a terahertz electromagnetic wave

. J. Phys. Chem. Lett. 11, 256-262 (2019). https://doi.org/10.1021/acs.jpclett.9b03228
Baidu ScholarGoogle Scholar
53. G. Hummer, J. C. Rasaiah and J. P. Noworyta,

Water conduction through the hydrophobic channel of a carbon nanotube

. Nature 414, 188-190 (2001). https://doi.org/10.1038/35102535
Baidu ScholarGoogle Scholar
54. M. Thomas, M. Brehm, R. Fligg, et al.,

Computing vibrational spectra from ab initio molecular dynamics

. Phys. Chem. Chem. Phys. 15, 6608-6622 (2013). https://doi.org/10.1039/C3CP44302G
Baidu ScholarGoogle Scholar
55. D. L. Gilden,

Cognitive emissions of 1/f noise

. Psychol. Rev. 108, 33-56 (2001). https://doi.org/10.1037/0033-295X.108.1.33
Baidu ScholarGoogle Scholar
56. P. Stoica and R. L. Moses, Spectral analysis of signals, (Pearson Prentice Hall, Upper Saddle River, NJ, 2005)
57. D. L. Gilman, T. Thornton and M. W. Mallon,

1/f noise in human cognition

. Science 267, 1837-1839 (1995). https://doi.org/10.1126/science.7892611
Baidu ScholarGoogle Scholar
58. T. Ji, H. W. Zhao, P. Y. Han, et al.,

Terahertz identification and quantification of penicillamine enantiomers

. Nucl. Sci. Tech. 24, 010201 (2013). https://doi.org/10.13538/j.1001-8042/nst.2013.01.005
Baidu ScholarGoogle Scholar
Footnote

The online version of this article (https://doi.org/10.1007/s41365-020-00767-w) contains supplementary material, which is available to authorized users.