This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Molecular dynamics simulations of 1H NMR relaxation in Gd3+–aqua

Philip M. Singer ps41@rice.edu Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Arjun Valiya Parambathu Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Thiago J. Pinheiro dos Santos Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Yunke Liu Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Lawrence B. Alemany Shared Equipment Authority and Department of Chemistry, Rice University, 6100 Main St., Houston, TX 77005, USA    George J. Hirasaki Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Walter G. Chapman Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA    Dilip Asthagiri dna6@rice.edu Department of Chemical and Biomolecular Engineering, Rice University, 6100 Main St., Houston, TX 77005, USA
Abstract

Atomistic molecular dynamics simulations are used to investigate 1H NMR T1T_{1} relaxation of water from paramagnetic Gd3+ ions in solution at 25C. Simulations of the T1T_{1} relaxivity dispersion function r1r_{1} computed from the Gd3+1H dipole–dipole autocorrelation function agree within 8\simeq 8% of measurements in the range f0f_{0}\simeq 5 \leftrightarrow 500 MHz, without any adjustable parameters in the interpretation of the simulations, and without any relaxation models. The simulation results are discussed in the context of the Solomon-Bloembergen-Morgan inner-sphere relaxation model, and the Hwang-Freed outer-sphere relaxation model. Below f0f_{0}\lesssim 5 MHz, the simulation overestimates r1r_{1} compared to measurements, which is used to estimate the zero-field electron-spin relaxation time. The simulations show potential for predicting r1r_{1} at high frequencies in chelated Gd3+ contrast-agents used for clinical MRI.

I Introduction

The traditional theory of enhanced 1H NMR (nuclear magnetic resonance) relaxation of water due to paramagnetic transition-metal ions and lanthanide ions in aqueous solutions originates from Solomon [1], Bloembergen and Morgan [2, 3], a.k.a. the Solomon-Bloembergen-Morgan (SBM) model. The extended SBM model [4] accounts for paramagnetic relaxation of inner-sphere water in the paramagnetic-ion complex [1, 2, 3], outer-sphere water [5, 6], a.k.a. the Hwang-Freed (HF) model, the contact term [2, 7, 8, 3], the Curie term [9, 10], and the electron-spin relaxation [3, 11, 12, 9, 13, 10, 14]. The extended SBM model is most widely used in the interpretation of paramagnetic enhanced 1H relaxation of water due to Gd3+-based contrast agents [15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36] used in clinical MRI (magnetic resonance imaging). The SBM model also forms the basis for the interpretation of paramagnetic relaxation in water-saturated sedimentary rocks [37, 38, 39, 40, 41, 42, 43, 44].

The inner-sphere water constitutes the ligands of the Gd3+ complex. The SBM inner-sphere model assumes a rigid Gd3+1H dipole-dipole pair undergoing rotational diffusion, which according to the Debye model results in mono-exponential decay of the autocorrelation function. The Debye model was previously used in the Bloembergen-Purcell-Pound (BPP) model [45] for like spins (e.g. 1H–1H pairs), and then adopted in the SBM model for unlike spins (e.g. Gd3+1H pairs). The SBM inner-sphere model also takes the electron-spin relaxation time into account, resulting in 6 free parameters for the 1H NMR relaxivity r1r_{1}. The inner-sphere relaxation is generally considered the largest contribution to relaxivity.

The outer-sphere water are less tightly bound than the inner-sphere water. The HF outer-sphere model assumes that the Gd3+ ion and H2O are two force-free hard-spheres undergoing translational diffusion, which results in stretched-exponential decay of the autocorrelation function [5, 6]. The HF outer-sphere model adds an additional 2 free parameters, bringing the total to 8 free parameters for the extended SBM model. Relaxation from the contact term is negligible for 1H NMR in Gd3+–aqua [22, 29], and is therefore neglected. Likewise, the Curie term is negligible in the present case [9, 10].

The application of the SBM and HF models to Gd3+–aqua therefore requires fitting 8 free parameters over a large frequency range in measured r1r_{1} dispersion (NMRD). In chelated Gd3+ complexes, an order parameter plus a shorter correlation time [46, 47] are added to the rotational motion of the complex [34, 33], or to the electron-spin relaxation [21, 28, 31], taking the total to 10 free parameters. This over-parameterized inversion problem often requires guidance in setting a range of values for the free parameters. It has also often speculated that the model for electron-spin relaxation is inadequate [4].

Atomistic molecular dynamics (MD) simulations can help elucidate some of these issues. MD simulation were previously used for like-spins, e.g. 1H–1H dipole-dipole pairs, such as liquid-state alkanes, aromatics, and water [48, 49, 50, 51], as well as methane over a large range of densities [52]. In all these cases, good agreement was found between simulated and measured 1H NMR relaxation and diffusion, without any adjustable parameters in the interpretation of the simulations. With the simulations thus validated against measurements, simulations can then be used to separate the intramolecular (i.e. rotational) from intermolecular (i.e. translational) contributions to relaxation, and to explore the corresponding 1H–1H dipole-dipole autocorrelation functions in detail. For instance, MD simulations revealed for the first time ever that water and alkanes do not conform to the BPP model of a mono-exponential decay in the rotational autocorrelation function, except for highly symmetric molecules such as neopentane. More complex systems such as viscous polymers [53] and heptane confined in a polymer matrix [54] have also been simulated, which again saw good agreement with measurements, and which lead to insights into the distribution in dynamic molecular modes.

In this report, we extend these MD simulation techniques to Gd3+1H dipole-dipole pairs, i.e. unlike spins, in a Gd3+–aqua complex. The simulations show good agreement with measurements in the range f0f_{0}\simeq 5 \leftrightarrow 500 MHz, without any adjustable parameters in the interpretation of the simulations, and without any relaxation models. These findings show potential for predicting r1r_{1} at high frequencies in chelated Gd3+ contrast agents used for clinical MRI. At the very least, the simulations could reduce the number of free parameters in the SBM and HF models, and help put constraints on its inherent assumptions for clinical MRI applications.

II Methodology

II.1 Molecular simulation

To model the Gd3+–aqua system, we use the AMOEBA polarizable force field to describe solvent water [55] and the ion [56]. The Gd3+ ion has an incomplete set of 4f4f orbitals, but this incomplete shell lies under the filled 5s5s and 5p5p orbitals. Thus ligand-field effects [57] are not expected to play a part in hydration and a spherical model of the ion ought to be adequate. However, polarization effects are expected to be important given the large charge on the cation.

Experimental NMR studies on Gd3+–aqua use concentrations of about 0.3 mM. This amounts to having one GdCl3 molecule in a solvent bath in a cubic cell about 176 Å in size. Such large simulations are computationally rather demanding with the AMOEBA polarizable forcefield. Further, at such dilutions, the ions essentially “see” only the water around them, and since understanding the behavior of water around the ion is of first interest, here we study a single Gd3+ ion in a bath of 512 water molecules. (Note that within the Ewald formulation for electrostatic interactions, there is an implicit neutralizing background. This background does not impact the forces between the ion and the water molecules that are of interest here.) The partial molar volume of Gd3+ in water at 298.15 K has been estimated by Marcus (1985) to be 59.6-59.6 cc/mol. We use this to fix the length of the cubic simulation cell to 24.80524.805 Å. (From constant pressure simulations in a 2006 water system we find a Gd3+ partial molar volume of about -63 cc/mol, in good agreement with the value suggested by Marcus. However, in this work, we will use the value suggested by Marcus.)

All the simulations were performed using the OpenMM-7.5.1 package [58]. The van der Waals forces were switched to zero from 11 Å to 12 Å. The real space electrostatic interactions were cutoff at 9 Å and the long-range electrostatic interactions were accounted using the particle mesh Ewald method with a relative error tolerance of 10-5 for the forces. In the polarization calculations the (mutually) induced dipoles were converged to 10-5 Debyes.

The equations of motion were integrated using the “middle” leapfrog propagation algorithm with a time step of 1 fs coupled; this combination of method and time step provides excellent energy conservation in constant energy simulations. Exploratory work shows that the NMR relaxation in the Gd3+-aqua system is sensitive to the system temperature. So we additionally use with a Nosé-Hoover chain [59, 60] with three thermostats [61] to simulate the system at 298.15 K. The collision frequency of the thermostat was set to 100 fs to ensure canonical sampling. We carried out extensive tests with and without thermostats to ensure that the Nosé-Hoover thermostat does not affect dynamical properties. Our conclusions are consistent with an earlier study on good practices for calculating transport properties in simulations [62].

Initially, the system was equilibrated under NVTNVT conditions at 298.15 K for over 200 ps. We then propagated the system under the same NVTNVT conditions for 8 ns, saving frames every 0.1 ps for analysis. We used the last 6.5536 ns of simulation (equal to 65536(=216=2^{16}) frames) for analysis. The mean temperature in the production phase was 298.15298.15 K with a standard error of the mean being 0.03 K.

II.2 Structure and dynamics

Fig. 1 shows the ion-water radial distribution function. The location and magnitude of the peak is in good agreement with earlier studies founded on either ab initio [29] or empirical force field-based [31] simulations. Consistent with those studies, we find that the first sphere (i.e. inner sphere) contains between q=89q=8\leftrightarrow 9 water molecules, with a mean of qq\simeq 8.5 consistent with published values [15, 17, 22, 21, 24, 31, 32].

Refer to caption
Figure 1: Radial distribution function g(r)g(r) of water oxygen around the Gd3+ ion; the function n(r)n(r) gives the coordination number.

II.3 Residence time analysis

To estimate the residence time of water molecules in the inner sphere, we need to keep track of the water molecule as it moves into/out of the inner sphere. To this end, we follow the residence time of a defined water molecule ww using an indicator function χw\chi_{w} that equals 1 if the water molecule is in the inner sphere and zero otherwise. The inner sphere is defined as a sphere of radius r3.3r\leq 3.3 Å around the Gd3+ ion; this corresponds to the first hydration of the ion (Fig. 1). We perform this analysis for all the water molecules that visit the inner sphere at least once during the simulation. (We emphasize that the time here is discrete because configurations are saved only every 100 fs.)

Fig. 2 shows the trace of the indicator function for a particular water molecule.

Refer to caption
Figure 2: The trace of the indicator function χw\chi_{w} for part of the simulation trajectory (red curve). Note the several transient escapes of the water molecule out of the inner sphere. The dashed (blue) curve is obtained by “windowing” the raw data as noted in the text. τ\tau is the length of time the water molecule spends continuously inside the inner sphere (one such domain, τi\tau_{i}, is shown).

Note that during the approximately 4 ns window (from the 8\approx 8 ns trajectory), the water molecule makes several excursions out of the inner sphere before permanently leaving the inner sphere around 4.5 ns. The length of time that the water molecule spends continuously inside the inner sphere is denoted by τi\tau_{i}. As Fig. 2 shows, there can be several such islands of continuous occupation.

To test if the transient excursion is a bona fide escape from the inner sphere, we window average the data as follows: we consider a transient escape as a bona fide escape only if it persists for a defined number of consecutive time points. Fig. 2 shows the trace (blue curve) for such a “windowing” for a window length of 200 fs, i.e. two consecutive frames in the trajectory of configurations. As can be expected, windowing extends the length of time that the molecule is defined to be inside the inner sphere.

For all the water molecules that visit the inner sphere, we accumulate the set {τi}\{\tau_{i}\} and then construct the histogram h(τ)h(\tau). Figure 3 shows the h(τ)h(\tau) for the raw data and the data with window length of 200 fs.

Refer to caption
Figure 3: Distribution of continuous occupancy times τ\tau, h(τ)h(\tau), for raw data and 200 fs window.

The h(τ)h(\tau) curve shows that there are many cases where water only transiently resides in the inner sphere. But we also find several water molecules that spend upwards of 0.5 ns continuously within the inner sphere. Specifically, for the data that has not been smoothed by “windowing”, we find three instances of water molecules continuously spending between 1 to 1.5 ns inside the inner sphere (and these also happen to be three distinct water molecules [data not shown]); with windowing using a 200 fs window, the upper limit is extended to just over 2 ns. Thus residence time, τm\tau_{m}, between 11\leftrightarrow 2 ns are predicted for a complete rejuvenation of the inner sphere population. This time-scale is in accord with the range of published 17O NMR data that suggest residence times τm1.01.5\tau_{m}\simeq 1.0\leftrightarrow 1.5 ns [16, 63, 20, 64, 26].

II.4 1H NMR relaxivity

The enhanced 1H NMR relaxation rate 1/T11/T_{1} for water is given by the average over the N=1024N=1024 1H nuclei in the L=24.805L=24.805 Å box containing one paramagnetic Gd3+ ion:

1T1\displaystyle\frac{1}{T_{1}} =1Ni=1N1T1i,\displaystyle=\frac{1}{N}\sum_{i=1}^{N}\frac{1}{T_{1i}}, (1)
1T2\displaystyle\frac{1}{T_{2}} =1Ni=1N1T2i,\displaystyle=\frac{1}{N}\sum_{i=1}^{N}\frac{1}{T_{2i}}, (2)

where T1iT_{1i} is the T1T_{1} relaxation for the ii’th 1H nucleus. The gadolinium molar concentration is given by [Gd]=[H]/N[Gd]=[H]/N, where [H][H] is the molar concentration of 1H in the simulation box. Equivalently, [H]=2[W][H]=2[W] where [W]=[W]= 55,705 mM is the molar concentration of H2O at 25C. This leads to the following expression for the relaxivity in units of (mM-1s-1):

r1\displaystyle r_{1} =1[Gd]1T1=1[H]i=1N1T1i,\displaystyle=\frac{1}{[Gd]}\frac{1}{T_{1}}=\frac{1}{[H]}\sum_{i=1}^{N}\frac{1}{T_{1i}}, (3)
r2\displaystyle r_{2} =1[Gd]1T2=1[H]i=1N1T2i.\displaystyle=\frac{1}{[Gd]}\frac{1}{T_{2}}=\frac{1}{[H]}\sum_{i=1}^{N}\frac{1}{T_{2i}}. (4)

Note that r1,2r_{1,2} are independent of NN (or box size LL, equivalently). The “fast-exchange” regime (T1,2τmT_{1,2}\gg\tau_{m}) is assumed [11], and the chemical shift term in r2r_{2} [10] is neglected for simplicity. Note also that the 1H–1H dipole-dipole relaxation [48] is not considered in these simulations.

The computation of T1i,2iT_{1i,2i} originates from the Gd3+1H dipole-dipole autocorrelation function G(t)G(t) [45, 5, 65, 66, 67, 68] shown in Fig. 4(a), where tt is the lag time. This autocorrelation is well suited for computation using MD simulations [69, 70]. Using the convention in the text by McConnell [66], Gi(t)G_{i}(t) in units of (s-2) is determined by:

Gi(t)=14(μ04π)22γI2γS2S(S+1)×(3cos2θi(t+t)1)ri3(t+t)(3cos2θi(t)1)ri3(t)t,G_{i}(t)=\frac{1}{4}\!\left(\frac{\mu_{0}}{4\pi}\right)^{2}\!\hbar^{2}\gamma_{I}^{2}\gamma_{S}^{2}S(S+1)\times\\ \left<\frac{(3\cos^{2}\!\theta_{i}\!(t+t^{\prime})-1)}{r_{i}^{3}\!\left(t+t^{\prime}\right)}\frac{(3\cos^{2}\!\theta_{i}\!(t^{\prime})-1)}{r_{i}^{3}\!(t^{\prime})}\right>_{\!\!t^{\prime}}, (5)

for the ii’th 1H nucleus. θi\theta_{i} is the angle between the Gd3+1H vector 𝐫i{\bf r}_{i} and the applied magnetic field 𝐁0{\bf B}_{0}. μ0\mu_{0} is the vacuum permeability, \hbar is the reduced Planck constant. γI/2π=42.58\gamma_{I}/2\pi=42.58 MHz/T is the nuclear gyro-magnetic ratio for 1H with spin I=1/2I=1/2, and γS=658γI\gamma_{S}=658\,\gamma_{I} is the electron gyro-magnetic ratio for Gd3+ with spin S=7/2S=7/2.

Note that in Eq. 5 we assume a spherically symmetric (i.e. isotropic) system, and therefore Gim(t)=Gi(t)G_{i}^{m}(t)=G_{i}(t) is independent of the order mm, which amounts to saying that the direction of the applied magnetic field 𝐁0=B0𝐳{\bf B}_{0}=B_{0}\bf z is arbitrary. This assumption was verified in Ref. [31]. For simplicity, we therefore use the m=0m=0 harmonic Y20(θ,ϕ)=5/16π(3cos2θ1)Y_{2}^{0}(\theta,\phi)=\sqrt{5/16\pi}\,(3\cos^{2}\theta-1) for the MD simulations [48].

Refer to caption
Figure 4: (a) MD simulations of the autocorrelation function G(t)G(t), where 1 in 10 data points are shown for clarity. Also shown is the modes expansion (Eq. 16) to G(t)G(t), and SBM model (Eq. 25) with τR=τ\tau_{R}=\left<\tau\right> = 30 ps. (b) Spectral density functions J(ω)J(\omega) from FFT (fast Fourier transform) (Eq. 8), including the f=0f=0 data point represented as a horizontal line placed at low frequency. Also shown is the modes expansion (Eq. 21), and the f2f^{-2} power-law expected in the SBM model at high frequencies.

The second-moment Δωi2\Delta\omega_{i}^{2} (i.e. strength) of the dipole-dipole interaction is defined as such [67]:

Gi(0)\displaystyle G_{i}(0) =13Δωi2.\displaystyle=\frac{1}{3}\Delta\omega_{i}^{2}. (6)

Assuming the angular term in Eq. 5 is uncorrelated with the distance term at t=0t=0, the relation (3cos2θi(τ)1)2τ=4/5\left<(3\cos^{2}\!\theta_{i}\!(\tau)-1)^{2}\right>_{\tau}=4/5 (which is independent of ii) reduces the second moment to:

Δωi2=35(μ04π)22γI2γS2S(S+1)1ri6(t)t.\displaystyle\Delta\omega_{i}^{2}=\frac{3}{5}\!\left(\frac{\mu_{0}}{4\pi}\right)^{2}\!\hbar^{2}\gamma_{I}^{2}\gamma_{S}^{2}S(S+1)\left<\!\frac{1}{r_{i}^{6}(t^{\prime})}\!\right>_{\!\!t^{\prime}}. (7)

The next step is to take the (two-sided even) fast Fourier transform (FFT) of the Gi(t)G_{i}(t) to obtain the spectral density function:

Ji(ω)\displaystyle J_{i}(\omega) =20Gi(t)cos(ωt)𝑑t.\displaystyle=2\int_{0}^{\infty}\!G_{i}(t)\cos(\omega t)\,dt. (8)

The relaxation rates are then determined for unlike spins [66]:

1T1i\displaystyle\frac{1}{T_{1i}} =Ji(ω0)+73Ji(ωe),\displaystyle=J_{i}(\omega_{0})+\frac{7}{3}J_{i}(\omega_{e}), (9)
1T2i\displaystyle\frac{1}{T_{2i}} =23Ji(0)+12Ji(ω0)+136Ji(ωe),\displaystyle=\frac{2}{3}J_{i}(0)+\frac{1}{2}J_{i}(\omega_{0})+\frac{13}{6}J_{i}(\omega_{e}), (10)

assuming ωeω0\omega_{e}\gg\omega_{0}, where ω0=γIB0=2πf0\omega_{0}=\gamma_{I}B_{0}=2\pi f_{0} is the 1H NMR resonance frequency, and ωe=658ω0\omega_{e}=658\,\omega_{0} is the electron resonance frequency.

The expressions for T1i,2iT_{1i,2i} are then summed in Eqs. 3 and 4 to compute r1r_{1} and r2r_{2}, respectively. We also define the following quantities summed over the N=1024N=1024 1H nuclei:

G(t)\displaystyle G(t) =i=1NGi(t),\displaystyle=\sum\limits_{i=1}^{N}\!G_{i}(t), (11)
Δω2\displaystyle\Delta\omega^{2} =i=1NΔωi2\displaystyle=\sum\limits_{i=1}^{N}\Delta\omega_{i}^{2} (12)
J(ω)\displaystyle J(\omega) =i=1NJi(ω).\displaystyle=\sum\limits_{i=1}^{N}J_{i}(\omega). (13)

Note that the summed quantities G(t)G(t), Δω2\Delta\omega^{2}, and J(ω)J(\omega) are independent of NN (or box size LL, equivalently). The G(t)G(t) simulation data is plotted in Fig 4(a), while the J(ω)J(\omega) simulation data is plotted in Fig 4(b).

We also define the average correlation time τ\left<\tau\right> as the normalized integral [67]:

τ\displaystyle\left<\tau\right> =1G(0)0G(t)𝑑t.\displaystyle=\frac{1}{G(0)}\int_{0}^{\infty}\!G(t)\,dt. (14)

The low frequency (i.e. extreme narrowing) limit r1(0)=r2(0)=r1,2(0)r_{1}(0)=r_{2}(0)=r_{1,2}(0) can then be expressed as:

r1,2(0)=1[H]209Δω2τ.\displaystyle r_{1,2}(0)=\frac{1}{[H]}\frac{20}{9}\Delta\omega^{2}\!\left<\tau\right>. (15)

Note how r1,2(0)r_{1,2}(0) (for unlike Gd3+1H spin pairs) is a factor 2/3 less than the equivalent expression for like 1H–1H spin pairs [48].

We assume the “fast-exchange” regime in the above formulation, which is discussed in more detail in Section III.2. The fast-exchange regime can be inferred directly from measurements since r1r_{1} increases with decreasing temperature [15]. Investigations are underway to extend the simulations to the slow-exchange regime.

The above analysis also assumes the electron spin is a point-dipole centered at the Gd3+ ion [4]. Given that the simulation agrees with measurements in the range f0f_{0}\simeq 5 \leftrightarrow 500 MHz, the point-dipole approximation is considered valid in the present case for 1H [10].

II.5 Expansion of G(t)G(t) in terms of molecular modes

The FFT result for J(ω)J(\omega) in Fig. 4(b) is sparse. Besides the ff = 0 data point, the lowest frequency FFT data point is given by the resolution Δf=1/2tmax=\Delta f=1/2t_{max}= 103 MHz, where tmax=500t_{max}=500 ps is the longest lag time in G(t)G(t).

As an alternative to zero-padding the FFT, we see to model G(t)G(t) in terms of molecular modes. To this end, we expand G(t)G(t) as

G(t)\displaystyle G(t) =0P(τ)exp(tτ)𝑑τ,\displaystyle=\int_{0}^{\infty}\!P(\tau)\exp\left(-\frac{t}{\tau}\right)d\tau, (16)

where P(τ)P(\tau) is the underlying distribution in molecular correlation times, τ\tau. We solve this Fredholm integral equation of the first kind to recover the P(τ)P(\tau) distribution. Since G(t)G(t) is available only at discrete time intervals, the inversion is an ill-posed problem. We address this by using Tikhonov regularization [50, 71], with the vector P being one for which

𝐏=argmin𝐏0𝐆K𝐏2+α𝐏2\displaystyle{\bf P}=\underset{{\bf P}\geq 0}{\mathrm{arg\,min}}\,\,||{\bf G}-K\,{\bf P}||^{2}+\alpha||{\bf P}||^{2} (17)

is a minimum. Here G is the column vector representation of the autocorrelation function G(t)G(t), P is the column vector representation of the distribution function P(τ)P(\tau), α\alpha is the regularization parameter, and KK is the kernel matrix:

K=Kij=exp(tiτj).\displaystyle K=K_{ij}=\exp\left(-\frac{t_{i}}{\tau_{j}}\right). (18)

The results for P(τ)P(\tau) are shown in Fig. 5, from which the following are determined:

τ\displaystyle\left<\tau\right> =1G(0)0P(τ)τ𝑑τ,\displaystyle=\frac{1}{G(0)}\int_{0}^{\infty}\!P(\tau)\,\tau\,d\tau, (19)
G(0)\displaystyle G(0) =0P(τ)𝑑τ=13Δω2.\displaystyle=\int_{0}^{\infty}\!P(\tau)d\tau=\frac{1}{3}\Delta\omega^{2}. (20)

The spectral density J(ω)J(\omega) is then determined from the Fourier transform (Eq. 8) of G(t)G(t) (Eq. 16):

J(ω)\displaystyle J(\omega) =02τ1+(ωτ)2P(τ)𝑑τ,\displaystyle=\int_{0}^{\infty}\!\frac{2\tau}{1+(\omega\tau)^{2}}P(\tau)d\tau\,, (21)

from which T1,2T_{1,2} at any desired f0f_{0} can be determined (Eqs. 9 and 10). More in-depth discussions of the above procedure, loosely termed as “inverse Laplace transform”, can be found in Refs. [54, 53, 52, 51, 72, 73] and the supplementary material in Refs. [50, 71]. We hasten to add that inverting Eq. 16 to recover PP is formally not a Laplace inversion [74], but this terminology is common in the literature. Possible alternatives to the above formulation are also discussed in Ref. [51].

The P(τ)P(\tau) is binned from τmin=0.05\tau_{min}=0.05 ps to τmax=500\tau_{max}=500 ps using 200 logarithmically spaced bins. In the present case of low-viscosity fluids (η\eta\simeq 1 cP), the choice of τmax\tau_{max} does not impact J(ω)J(\omega), and is therefore not a free parameter in the analysis in terms of molecular modes. The constant “div” in Fig. 5 is a “division” on a log-scale. More specifically, div =log10(τi+1)log10(τi)=\log_{10}(\tau_{i+1})-\log_{10}(\tau_{i}) is independent of the bin index ii, and ensures unit area when P(τ)P(\tau) is of unit height and a decade wide [71].

Refer to caption
Figure 5: Probability density function P(τ)P(\tau) obtained from the expansion of molecular modes (Eq. 16) of G(t)G(t). The average correlation time τ\left<\tau\right> (Table 1) and predicted translational correlation time τT\tau_{T} (Table 2) are shown.

As shown in Fig. 4(a), the residual between the G(t)G(t) data and the fit using molecular modes is not dominated by Gaussian noise. As such, the regularization parameter is fixed to α=\alpha= 10-1 in accordance with previous studies [50, 52, 53, 54, 51]. As shown in Fig. 4(b), we find that α=\alpha= 10-1 gives excellent agreement with the parameter-free J(ω)J(\omega) from FFT, which validates the analysis in terms of molecular modes. The results in Fig. 4(b) further emphasize the following advantages: (1) the expansion (Eq. 16) filters out the noise while still honoring the FFT data (including the f=0f=0 data point), (2) Eq. 21 provides J(ω)J(\omega) for any desired ff value, and (3) the expansion in terms of molecular modes leads to physical insight into the distribution P(τ)P(\tau) of molecular correlation times τ\tau.

II.6 Diffusion

An independent computation of translational diffusion DTD_{T} was performed from MD simulations. We calculate the mean square displacement Δr2\left<\!\Delta r^{2}\right> of the water oxygen and Gd3+ ion as a function of the diffusion evolution time tt (<< 10 ps), and additionally average over a sample of 50 molecules to ensure adequate statistical convergence.

Refer to caption
Figure 6: MD simulations of mean-square displacement Δr2\left<\!\Delta r^{2}\right> versus time tt for bulk water, water in Gd3+–aqua, and Gd3+ in Gd3+–aqua. Solid lines show fitting region used to obtain translational diffusion coefficient DsimD_{sim} from Eq. 22 for t>2t>2 ps, and dashed lines show early time regime not used in the fit. Legend indicates DTD_{T} values which include the correction term (Eq. 23).

As shown in Fig.  6, at long-times (tt) the slope of the linear diffusive regime gives the translational self-diffusion coefficient DsimD_{sim} according to Einstein’s relation:

Dsim=16δΔr2δtD_{sim}=\frac{1}{6}\frac{\delta\!\left<\!\Delta r^{2}\right>}{\delta t}\, (22)

where DsimD_{sim} is the diffusion coefficient obtained in the simulation using periodic boundary conditions in a cubic cell of length LL. In the linear regression procedure, the early ballistic regime and part of the linear regime is excluded to obtain a robust estimate of DsimD_{sim}.

Following Yeh and Hummer [75] (see also Dünweg and Kremer [76]), we obtain the diffusion coefficient for an infinite system DTD_{T} from DsimD_{sim} using

DT=Dsim+kBT6πηξLD_{T}=D_{sim}+\frac{k_{B}T}{6\pi\eta}\frac{\xi}{L} (23)

where η\eta is the shear viscosity and ξ=2.837297\xi=2.837297 [75] is the same quantity that arises in the calculation of the Madelung constant in electrostatics. (In the electrostatic analog of the hydrodynamic problem, ξ/L\xi/L is the potential at the charge site in a Wigner lattice.) For the system sizes considered in this study, L25L\simeq 25 Å, the correction factor constituted 13%\simeq 13\% of D0D_{0}, and 16%\simeq 16\% of DWD_{W}. The correction factor was not applied to DGdD_{Gd}.

II.7 Measurements

We prepared a Gd3+–aqua solution in de-ionized water at [Gd][Gd] = 0.3 mM and measured T1,measT_{1,meas} at a controlled temperature of 25C, using static fields at f0=2.3f_{0}=2.3 MHz with an Oxford Instruments GeoSpec2, at f0=20f_{0}=20 MHz with a Bruker Minispec, and at f0=500f_{0}=500 MHz with a Bruker 500 MHz Spectrometer. The measured relaxivity r1r_{1} was determined as such:

r1=1[Gd](1T1,meas1T1,bulk).\displaystyle r_{1}=\frac{1}{[Gd]}\left(\!\frac{1}{T_{1,meas}}-\frac{1}{T_{1,bulk}}\!\right). (24)

where T1,bulk=T_{1,bulk}= 3.13 s was found for bulk water (not de-oxygenated [77]) at 2.3 MHz and 500 MHz. Field cycling r1r_{1} data at [Gd][Gd] = 1 mM and 25C were taken from Luchinat et al. [32] (supplementary material) using a Stelar SpinMaster 1T. The field cycling results agreed with our measurements at f0=f_{0}= 2.3 MHz and 20 MHz (within \simeq 5%), while our f0=500f_{0}=500 MHz significantly extends the frequency range of the measurements.

III Results and discussions

In this section we compare the simulated relaxivity r1r_{1} with measurements. The simulated relaxivity is then discussed in the context of the Solomon-Bloembergen-Morgan (SBM) model, and Hwang-Freed (HF) model. The zero-field electron-spin relaxation time is then determined from r1r_{1} at low frequencies.

III.1 Comparison with measurements

The results of simulated relaxivity r1r_{1} (Eq. 3) are shown in Fig. 7, alongside corresponding measurements of Gd3+–aqua solution at 25C using field cycling by Luchinat et al. [32] (supplementary material), and static fields from this work. A cross-plot of simulated versus measured r1r_{1} results are also shown in Fig. 8. The simulation is within \simeq 8% of measurements in the range f0f_{0}\simeq 5 \leftrightarrow 500 MHz. Given that there are no adjustable parameters in the interpretation of the simulations, this agreement validates our simulations at high frequencies. We note that a similar degree of agreement (within \simeq 7%) was found in previous studies of liquid alkanes and water [48].

For convenience, Table 1 lists the average correlation time τ\left<\tau\right> (Eq. 19), the square-root of the second moment (i.e. strength) of Δω\Delta\omega (Eq. 20), and the residence time τm\tau_{m} (Fig. 3). Note that these three quantities are model free.

τ\left<\tau\right> Δω/2π\Delta\omega/2\pi τm\tau_{m} qq rinr_{in} Δωin/2π\Delta\omega_{in}/2\pi T1in(0)T_{1in}(0) Te0T_{e0}
(ps) (MHz) (ns) (Å) (MHz) (μ\mus) (ps)
30 38.4 1 \leftrightarrow 2 8.5 2.97 9.3 4.4 180
Table 1: Analysis of the simulation results including; (left) mean correlation time τ\left<\tau\right> (Eq. 19), square-root of second moment Δω\Delta\omega (Eq. 20), and residence time τm\tau_{m} (Fig. 3); (middle) approximate inner-sphere quantities including coordination number qq (Fig. 1), Gd3+1H distance rinr_{in} (Eq. 26), square-root of second moment Δωin\Delta\omega_{in} (Eq. 27), and relaxation time T1in(0)T_{1in}(0) at f0=0f_{0}=0 (Eq. 28); (right) zero-field electron-spin relaxation time Te0T_{e0} (Eq. 33).
Refer to caption
Figure 7: Simulated 1H NMR relaxivity r1r_{1} of Gd3+–aqua solution at 25C, compared with static-field measurements (this work) and field-cycling measurements (Luchinat et al. [32], supplementary material). Also shown is the average of the low-frequency (f0<0.5f_{0}<0.5 MHz) measurements.
Refer to caption
Figure 8: Cross-plot of measured r1r_{1} (including static-field and field-cycling measurements) versus simulations taken from Fig. 7, all at 25C. Dashed line is the 1-1 unity line.

III.2 SBM inner-sphere model

The SBM inner-sphere model assumes a rigid Gd3+1H pair undergoing rotational diffusion, leading to the following mono-exponential decay in the autocorrelation function:

GSBM(t)=GSBM(0)exp(tτR).\displaystyle G_{SBM}(t)=G_{SBM}(0)\exp\left(-\frac{t}{\tau_{R}}\right). (25)

This functional form is identical to the BPP model [45] which is based on the Debye model, where the rotational-diffusion correlation time τR\tau_{R} is defined as the average time it takes the rigid pair to rotate by 1 radian. GSBM(t)G_{SBM}(t) is plotted in Fig. 4(a) assuming τR=τ\tau_{R}=\left<\tau\right> = 30 ps.

As shown in Fig. 4(a), the mono-exponential decay in GSBM(t)G_{SBM}(t) is not consistent with the multi-exponential (i.e., stretched) decay in G(t)G(t). Equivalently, J(ω)J(\omega) in Fig. 4(b) does not follow the f2f^{-2} power-law behavior at large ff. This is expected since the simulations implicitly include both inner-sphere and outer-sphere (see below) contributions. Currently the simulations do not separate inner-sphere from outer-sphere contributions, therefore the simulations do not clarify the origin of the multi-exponential decay in G(t)G(t).

It is also possible that the inner-sphere G(t)G(t) itself has a multi-exponential decay of its own. The functional form of GSBM(t)G_{SBM}(t) is based on the Debye model, which was previously shown to be inaccurate when used in the BPP model for liquid alkanes and water [48]. It would therefore not be surprising if the inner-sphere G(t)G(t) was also multi-exponential in nature.

Assuming that inner-sphere relaxation dominates, and therefore that the multi-exponential decay in G(t)G(t) is due to inner-sphere dynamics alone, the correlation time τ\left<\tau\right>\simeq 30 ps is consistent with published values from the SBM inner-sphere model, where a range of τR2345\tau_{R}\simeq 23\leftrightarrow 45 ps is reported [15, 17, 22, 21, 24, 31, 32]. Note that τ\left<\tau\right>\simeq 30 ps is a factor 10\simeq 10 larger than that of bulk water (τR2.7\tau_{R}\simeq 2.7 ps [48]), which is expected given the hindered rotational motion of the Gd3+–aqua complex.

Continuing with the assumption that inner-sphere relaxation dominates, one can approximate the following inner-sphere quantities:

1rin6\displaystyle\frac{1}{r_{in}^{6}} 12qi=1N1ri6(t)t\displaystyle\simeq\frac{1}{2q}\sum_{i=1}^{N}\left<\!\frac{1}{r_{i}^{6}(t^{\prime})}\!\right>_{\!\!t^{\prime}} (26)
Δωin2\displaystyle\Delta\omega^{2}_{in} 12qi=1NΔωi2\displaystyle\simeq\frac{1}{2q}\sum_{i=1}^{N}\Delta\omega_{i}^{2} (27)
1T1in\displaystyle\frac{1}{T_{1in}} 12qi=1N1T1i.\displaystyle\simeq\frac{1}{2q}\sum_{i=1}^{N}\frac{1}{T_{1i}}. (28)

The expressions are averaged over the 2q2q 1H nuclei in the inner sphere, where q=8.5q=8.5 is the H2O inner-sphere coordination number determined from n(r)n(r) (Fig. 1) at r3.5r\simeq 3.5 Å. Note again that these equations neglect the outer-sphere contribution, implying that Δωin2\Delta\omega_{in}^{2} is an upper bound, while rinr_{in} and T1in(0)T_{1in}(0) are lower bounds.

According to Table 1, the resulting rinr_{in}\simeq 2.97 Å is consistent with published values from the inner-sphere SBM model, where a range of rin3.03.2r_{in}\simeq 3.0\leftrightarrow 3.2 Å is reported [15, 17, 22, 21, 24, 31, 32]. The product Δωinτ0.002\Delta\omega_{in}\left<\tau\right>\simeq 0.002 indicates that the Redfield-Bloch-Wagness condition (Δωτ1\Delta\omega\left<\tau\right>\ll 1) is satisfied [65, 66, 67], which justifies the relaxivity analysis used here. The fast-exchange regime [11] can also be verified by noting that T1in(0)T_{1in}(0)\simeq 4.4 μ\mus at f0=0f_{0}=0 is three orders of magnitude larger than τm1\tau_{m}\simeq 1 ns, i.e. (T1in+τm)T1in(T_{1in}+\tau_{m})\simeq T_{1in} can be assumed.

III.3 HF outer-sphere model

We now discuss the outer-sphere contribution to relaxivity, although it is generally believed (though not proven) to be smaller than inner-sphere relaxivity [15, 18, 23]. The outer-sphere contribution is expected to follow the Hwang-Freed (HF) model for the relative translational diffusion between Gd3+ and H2O assuming two force-free hard-spheres [6]:

GHF(t)=GHF(0)54π0x281+9x22x4+x6×exp(x2tτD)dx.G_{HF}(t)=G_{HF}(0)\frac{54}{\pi}\int\limits_{0}^{\infty}\frac{x^{2}}{81+9x^{2}-2x^{4}+x^{6}}\times\\ \exp\left(-x^{2}\frac{t}{\tau_{D}}\right)dx. (30)

The translational-diffusion correlation time τD\tau_{D} is defined as the average time it takes the molecule to diffuse by one hard-core diameter dd. GHF(t)G_{HF}(t) is a multi-exponential decay by nature, and therefore one expects the total G(t)G(t) to be stretched, the extent of which depends on the relative contributions of outer-sphere to inner-sphere.

The correlation time τD\tau_{D} can be predicted as such:

τD\displaystyle\tau_{D} =d2DW+DGd=94τT,\displaystyle=\frac{d^{2}}{D_{W}+D_{Gd}}=\frac{9}{4}\tau_{T}, (31)

where the simulated diffusion coefficients of H2O (DWD_{W}) and Gd3+ (DGdD_{Gd}) in Gd3+–aqua are taken from Fig. 6, the results of which are listed in Table 2. The hard-core diameter dd is taken from the local maximum at r4.6r\simeq 4.6 Å in the pair-correlation function g(r)g(r) in Fig. 1 (which is attributed to outer-sphere water). Note that the resulting τD99\tau_{D}\simeq 99 ps is a factor 10\simeq 10 larger than that of bulk water (τD9.0\tau_{D}\simeq 9.0 ps [48]), which is expected given the larger hard-core distance d4.6d\simeq 4.6 Å than bulk water (dd\simeq 2.0 Å [48]).

DWD_{W} DGdD_{Gd} dd τD\tau_{D} τT\tau_{T}
(Å2{\rm\AA}^{2}/ps) (Å2{\rm\AA}^{2}/ps) (Å{\rm\AA}) (ps) (ps)
0.19 0.03 4.6 99 44
Table 2: Diffusion coefficients of H2O (DWD_{W}) and Gd3+ (DGdD_{Gd}) in Gd3+–aqua, distance of closest approach (dd) between Gd3+ and outer-sphere H2O according to g(r)g(r) (Fig. 1), and, resulting translational-diffusion correlation time τD(=9/4τT)\tau_{D}\,(=9/4\,\tau_{T}) from Eq. 31.

The value τD99\tau_{D}\simeq 99 ps is compared to the distribution P(τ)P(\tau) in Fig. 5. More specifically, the translational correlation time τT44\tau_{T}\simeq 44 ps is plotted in Fig. 5, where the relation τD=9/4τT\tau_{D}=9/4\,\tau_{T} and the origin of the factor 9/4 is explained in [67, 50]. τT\tau_{T} lies within the P(τ)P(\tau) distribution, indicating that the outer-sphere may contribute to relaxivity. Further investigations are underway to separate inner-sphere from outer-sphere contributions in the simulations, without assuming any models.

We note that P(τ)P(\tau) in Fig. 5 has a small contribution at short τ101\tau\simeq 10^{-1} ps, which is a result of the sharp drop in G(t)G(t) over the initial t0.2t\simeq 0.2 ps (Fig. 4(a)). This molecular mode is also present for intramolecular relaxation in liquid alkanes and water, while it is absent for intermolecular relaxation. In the case of alkanes, the ubiquitous intramolecular mode at τ101\tau\simeq 10^{-1} ps is attributed to the fast spinning methyl end-groups [50]. Investigations are underway to better understand the origin of this mode in Gd3+–aqua, which can perhaps be explained using a two rotational-diffusion model such as found in bulk water [78]. The origin of other modes in P(τ)P(\tau) at τ100\tau\simeq 10^{0} ps and 101\simeq 10^{1} ps are also being investigated.

Finally, we note that the r1r_{1} dispersion in Fig. 7 results in a mild increase in the ratio T1/T2=r2/r17/6T_{1}/T_{2}=r_{2}/r_{1}\simeq 7/6 above f010f_{0}\gtrsim 10 MHz, until f06f_{0}\gtrsim 6 GHz where T1/T2T_{1}/T_{2} increases further. Combining Eqs. 9 and 10 with J(ωe)=0J(\omega_{e})=0 (i.e., slow-motion regime) and J(ω0)=J(0)J(\omega_{0})=J(0) (i.e., fast-motion regime) accounts for the factor T1/T2=T_{1}/T_{2}= 7/6 within the frequency range f0f_{0}\simeq 10 MHz \leftrightarrow 6 GHz. The ratio T1/T2=7/6T_{1}/T_{2}=7/6 was also used to explain water-saturated sandstones [38].

III.4 Electron-spin relaxation

At low-frequencies (f00.5f_{0}\lesssim 0.5 MHz), the difference between r1r_{1} measurements (\simeq 29.7 mM-1s-1) and simulation (\simeq 34.6 mM-1s-1) can be reconciled by taking the zero-field electron-spin relaxation time Te0=T1e(0)=T2e(0)T_{e0}=T_{1e}(0)=T_{2e}(0) into account. Assuming that the correlation times P(τ)P(\tau) are uncorrelated with the electron-spin relaxation times, the following expression results [3]:

r1,2(0)\displaystyle r_{1,2}^{\prime}(0) =1[H]209Δω2τ,\displaystyle=\frac{1}{[H]}\frac{20}{9}\Delta\omega^{2}\!\left<\tau^{\prime}\right>, (32)
1τ\displaystyle\frac{1}{\left<\tau^{\prime}\right>} =1τ+1Te0.\displaystyle=\frac{1}{\left<\tau\right>}+\frac{1}{T_{e0}}. (33)

This is equivalent to introducing an exponential decay term exp(t/Te0)\exp(-t/T_{e0}) inside the FFT integral of Eq. 8. The fitted value of Te0180T_{e0}\simeq 180 ps is determined by matching r1(0)r_{1}^{\prime}(0) to the average low-frequency (f00.5f_{0}\lesssim 0.5 MHz) measurement (Fig. 7). The resulting Te0180T_{e0}\simeq 180 ps is consistent with the published range of Te096160T_{e0}\simeq 96\leftrightarrow 160 ps [15, 17, 22, 21, 24, 31, 32]. Investigations are underway to incorporate T1e(ωe)T_{1e}(\omega_{e}) and T2e(ωe)T_{2e}(\omega_{e}) dispersion [4, 24, 26, 31] for predicting r1r_{1} at low frequencies.

IV Conclusions

Atomistic MD simulations of 1H NMR relaxivity r1r_{1} for water in Gd3+–aqua complex at 25C show good agreement (within \simeq 8%) with measurements in the range f0f_{0}\simeq 5 \leftrightarrow 500 MHz, without any adjustable parameters in the interpretation of the simulations, and without any relaxation models. This level of agreement validates the simulation techniques and analysis of Gd3+1H dipole-dipole relaxation for unlike spins. The simulations show potential for predicting r1r_{1} at high frequencies in chelated Gd3+ contrast-agents for clinical MRI, or at the very least the simulations could reduce the number of free parameters in the Solomon-Bloembergen-Morgan (SBM) inner-sphere and Hwang-Freed (HF) outer-sphere relaxation models.

Simulations suggest residence times between τm1\tau_{m}\simeq 1\leftrightarrow 2 ns for a complete rejuvenation of the inner sphere waters of Gd3+. Further, the average coordination number is q8.5q\simeq 8.5. These observations are consistent with previously reported interpretation of experiments using the SBM model.

The autocorrelation function G(t)G(t) shows a multi-exponential decay, with an average correlation time of τ\left<\tau\right>\simeq 30 ps. The multi-exponential nature of G(t)G(t) is expected given that the simulation implicitly includes both inner-sphere and outer-sphere contributions. The results are analyzed assuming that the inner-sphere relaxation dominates, yielding approximations for the average Gd3+1H separation rin2.97r_{in}\simeq 2.97 Å and rotational correlation time τR=τ\tau_{R}=\left<\tau\right>\simeq 30 ps, both of which are consistent with previously published values which use the SBM model.

A distance of closest approach (i.e. hard-core diameter) for Gd3+–H2O of d4.64d\simeq 4.64 Å is determined from the local maximum in g(r)g(r) (attribute to outer-sphere water), which together with the simulated diffusion coefficients of Gd3+ and H2O are used to estimate the translational-diffusion correlation time τD=9/4τT99\tau_{D}=9/4\,\tau_{T}\simeq 99 ps in the HF outer-sphere model. Comparing τT\tau_{T} to the distribution in molecular modes P(τ)P(\tau) (determined from the modes expansion of G(t)G(t)) indicates that the outer-sphere may contribute to relaxivity. Further investigations are underway to separate inner-sphere from outer-sphere contributions in the simulations, without assuming any models.

Below f0f_{0}\lesssim 5 MHz the simulation overestimates r1r_{1} compared to measurements, which is used to estimate the zero-field electron-spin relaxation time Te0T_{e0}. The resulting fitted value Te0180T_{e0}\simeq 180 ps is consistent with the published range of values. Further investigations are underway to incorporate dispersion in the electron-spin relaxation time.

Acknowledgments

We thank Vinegar Technologies LLC, Chevron Energy Technology Company, and the Rice University Consortium on Processes in Porous Media for financial support. We gratefully acknowledge the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy (No. DE-AC02-05CH11231) and the Texas Advanced Computing Center (TACC) at The University of Texas at Austin for high-performance computer time and support.

References

  • Solomon [1955] I. Solomon, Relaxation processes in a system of two spins, Phys. Rev. 99, 559 (1955).
  • Bloembergen [1957] N. Bloembergen, Proton relaxation times in paramagnetic solutions, J. Chem. Phys. 27, 572 (1957).
  • Bloembergen and Morgan [1961] N. Bloembergen and L. O. Morgan, Proton relaxation times in paramagnetic solutions. Effects of electron spin relaxation, J. Chem. Phys. 34, 842 (1961).
  • Kowalewski et al. [1985] J. Kowalewski, L. Nordenskiöld, N. Benetis, and P.-O. Westlund, Theory of nuclear spin relaxation on paramagnetic systems in solution, Prog. Nucl. Magn. Reson. Spect. 17, 141 (1985).
  • Torrey [1953] H. C. Torrey, Nuclear spin relaxation by translational diffusion, Phys. Rev. 92, 962 (1953).
  • Hwang and Freed [1975] L.-P. Hwang and J. H. Freed, Dynamic effects of pair correlation functions on spin relaxation by translational diffusion in liquids, J. Chem. Phys. 63, 4017 (1975).
  • Morgan and Nolle [1959] L. O. Morgan and A. W. Nolle, Proton spin relaxation in aqueous solutions of paramagnetic ions, II. Cr+++, Mn++, Ni++, Cu++, and Gd+++, J. Chem. Phys. 31, 365 (1959).
  • Bernheim et al. [1959] R. A. Bernheim, T. H. Brown, H. S. Gutowsky, and D. E. Woessner, Temperature dependence of proton relaxation times in aqueous solutions of paramagnetic ions, J. Chem. Phys. 30, 905 (1959).
  • Fries et al. [2003] P. H. Fries, G. Ferrante, E. Belorizky, and S. Rast, The rotational motion and electronic relaxation of the Gd(III) aqua complex in water revisited through a full proton relaxivity study of a probe solute, J. Chem. Phys. 119, 8636 (2003).
  • Helm [2006] L. Helm, Relaxivity in paramagnetic systems: Theory and mechanisms, Prog. Nucl. Magn. Reson. Spect. 49, 45 (2006).
  • Korringa et al. [1962] J. Korringa, D. O. Seevers, and H. C. Torrey, Theory of spin pumping and relaxation in systems with a low concentration of electron spin resonance centers, Phys. Rev. 127, 1143 (1962).
  • Westlund [1995] P.-O. Westlund, Dynamics of Solutions and Fluid Mixtures by NMR (John Wiley &\& Son, Chichester, 1995) Chap. : Nuclear paramagnetic spin relaxation theory, p. 174–229.
  • Kowalewski and Mäler [2006] J. Kowalewski and L. Mäler, Nuclear Spin Relaxation in Liquids: Theory, Experiments, and Applications (Taylor & Francis Group, 2006).
  • Belorizky et al. [2008] E. Belorizky, P. H. Fries, L. Helm, J. Kowalewski, D. Kruk, R. R. Sharp, and P.-O. Westlund, Comparison of different methods for calculating the paramagnetic relaxation enhancement of nuclear spins as a function of the magnetic field, J. Chem. Phys. 128, 052315 (2008).
  • Koenig and Epstein [1975] S. H. Koenig and M. Epstein, Ambiguities in the interpretation of proton magnetic relaxation data in water solutions of Gd3+ ions, J. Chem. Phys. 63, 2279 (1975).
  • Southwood-Jones et al. [1980] R. V. Southwood-Jones, W. L. Earl, K. E. Newman, and A. E. Merbach, Oxygen-17 NMR and EPR studies of water exchange from the first coordination sphere of gadolinium(III) aquoion and gadolinium(III) propylenediaminetetraacetate, J. Chem. Phys. 73, 5909 (1980).
  • Banci et al. [1985] L. Banci, I. Bertini, and C. Luchinat, 1H NMR studies of solutions of paramagentic metal ions in etheleneglycol, Inorgan. Chimica Acta 100, 173 (1985).
  • Lauffer [1987] R. B. Lauffer, Paramagnetic metal complexes as water proton relaxation agents for NMR imaging: Theory and design, Chem. Rev. 87, 901 (1987).
  • Koenig and III [1990] S. H. Koenig and R. D. B. III, Field-cycling relaxometry of protein solutions and tissue: implications for MRI, Prog. Nucl. Magn. Reson. Spect. 22, 487 (1990).
  • Micskei et al. [1993] K. Micskei, D. H. Powell, L. Helm, E. Brücher, and A. E. Merbach, Water exchange on [Gd(H20)8]3+ and [Gd(PDTA)(H2O)2]- in aqueous solution: a variable-pressure, -temperature and -magnetic field 17O NMR Study, Magn. Res. Chem. 31, 1011 (1993).
  • Strandberg and Westlund [1996] E. Strandberg and P.-O. Westlund, 1H NMRD profile and ESR lineshape calculation for an isotropic electron spin system with S = 7/2 a generalized modified Solomon–Bloembergen–Morgan theory for nonextreme-narrowing conditions, J. Magn. Reson. 25, 261 (1996).
  • Powell et al. [1996] D. H. Powell, O. M. N. Dhubhghaill, D. Pubanz, L. Helm, Y. S. Lebedev, W. Schlaepfer, and A. E. Merbach, Structural and dynamic parameters obtained from 17O NMR, EPR, and NMRD studies of monomeric and dimeric Gd3+ complexes of interest in magnetic resonance imaging: An integrated and theoretically self-consistent approach, J. Amer. Chem. Soc. 118, 9333 (1996).
  • Caravan et al. [1999] P. Caravan, J. J. Ellison, T. J. McMurry, and R. B. Lauffer, Gadolinium (III) chelates as MRI contrast agents: Structure, dynamics, and applications, Chem. Rev. 99, 2293 (1999).
  • Rast et al. [2001] S. Rast, P. H. Fries, E. Belorizky, A. Borel, L. Helm, and A. E. Merbach, A general approach to the electronic spin relaxation of Gd(III) complexes in solutions. Monte Carlo simulations beyond the Redfield limit, J. Chem. Phys. 115, 7554 (2001).
  • Bertini et al. [2001] I. Bertini, C. Luchinat, and G. Parigi, Solution NMR of Paramagntic Molecules; Applications to Metallobiomolecules and Models (Elsevier, 2001).
  • Borel et al. [2002] A. Borel, F. Yerly, L. Helm, and A. E. Merbach, Multiexponential electronic spin relaxation and redfield’s limit in Gd(III) complexes in solution: Consequences for 17O/1H NMR and EPR simultaneous analysis, J. Amer. Chem. Soc. 124, 2042 (2002).
  • Zhou et al. [2004] X. Zhou, P. Caravan, R. B. Clarkson, and P.-O. Westlund, On the philosophy of optimizing contrast agents. An analysis of 1H NMRD profiles and ESR lineshapes of the Gd(III) complex MS-325 +HSA, J. Magn. Reson. 167, 147 (2004).
  • Zhou and Westlund [2005] X. Zhou and P.-O. Westlund, The viscosity and temperature dependence of X-band ESR lineshapes of Gd(III) aqueous complex, Spectrochim. Acta A 62, 76 (2005).
  • Yazyev and Helm [2007] O. V. Yazyev and L. Helm, Gadolinium (III) ion in liquid water: Structure, dynamics, and magnetic interactions from first principles, J. Chem. Phys. 127, 084506 (2007).
  • Yazyev and Helm [2008] O. V. Yazyev and L. Helm, Nuclear spin relaxation parameters of MRI contrast agents - insight from quantum mechanical calculations, Eur. J. Inorg. Chem. , 201 (2008).
  • Lindgren et al. [2009] M. Lindgren, A. Laaksonenb, and P.-O. Westlund, A theoretical spin relaxation and molecular dynamics simulation study of the Gd(H2O)+39{}_{9}^{+3} complex, Phys. Chem. Chem. Phys. 11, 10368 (2009).
  • Luchinat et al. [2014] C. Luchinat, G. Parigi, and E. Ravera, Can metal ion complexes be used as polarizing agents for solution DNP? A theoretical discussion, J. Biomol. NMR 58, 239–249 (2014).
  • Aime et al. [2019] S. Aime, M. Botta, D. Esteban-Gómez, and C. Platas-Iglesias, Characterisation of magnetic resonance imaging (MRI) contrast agents using NMR relaxometry, Mol. Phys. 117, 898 (2019).
  • Fragai et al. [2019] M. Fragai, E. Ravera, F. Tedoldi, C. Luchinat, and G. Parigi, Relaxivity of Gd-based MRI contrast agents in crosslinked hyaluronic acid as a model for tissues, ChemPhysChem 20, 2204 (2019).
  • Li and Meade [2019] H. Li and T. J. Meade, Molecular magnetic resonance imaging with Gd(III)-based contrast agents: Challenges and key advances, J. Amer. Chem. Soc. 141, 17025 (2019).
  • Wahsner et al. [2019] J. Wahsner, E. M. Gale, A. Rodríguez-Rodríguez, and P. Caravan, Chemistry of MRI contrast agents: current challenges and new frontiers, Chem. Rev. 119, 957 (2019).
  • Kleinberg et al. [1994] R. L. Kleinberg, W. E. Kenyon, and P. P. Mitra, Mechanism of NMR relaxation of fluids in rocks, J. Magn. Reson. 108, 206 (1994).
  • Foley et al. [1996] I. Foley, S. A. Farooqui, and R. L. Kleinberg, Effect of paramagnetic ions on NMR relaxation of fluids at solid surfaces, J. Magn. Reson. 123, 95 (1996).
  • Straley [2002] C. Straley, A mechanism for the temperature dependence of the surface relaxation rate in carbonates, Soc. Core Analysts SCA2002-27 (2002).
  • Zhang et al. [2003] G. Q. Zhang, G. J. Hirasaki, and W. V. House, Internal field gradients in porous media, Petrophysics 44, 422 (2003).
  • Korb et al. [2009] J.-P. Korb, G. Freiman, B. Nicot, and P. Ligneul, Dynamical surface affinity of diphasic liquids as a probe of wettability of multimodal porous media, Phys. Rev. E 80, 061601 (2009).
  • Mitchell et al. [2014] J. Mitchell, L. F. Gladden, T. C. Chandrasekera, and E. J. Fordham, Low-field permanent magnets for industrial process and quality control, Prog. Nucl. Magn. Reson. Spect. 76, 1 (2014).
  • Faux et al. [2015] D. A. Faux, S. H. P. Cachia, P. J. McDonald, J. S. Bhatt, N. C. Howlett, and S. V. Churakov, Model for the interpretation of nuclear magnetic resonance relaxometry of hydrated porous silicate materials, Phys. Rev. E 91, 032311 (2015).
  • Saidian and Prasad [2015] M. Saidian and M. Prasad, Effect of mineralogy on nuclear magnetic resonance surface relaxivity: A case study of Middle Bakken and Three Forks formations, Fuel 161, 197 (2015).
  • Bloembergen et al. [1948] N. Bloembergen, E. M. Purcell, and R. V. Pound, Relaxation effects in nuclear magnetic resonance absorption, Phys. Rev. 73, 679 (1948).
  • Lipari and Szabo [1982a] G. Lipari and A. Szabo, Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. 1. Theory and range of validity, J. Amer. Chem. Soc. 104, 4546 (1982a).
  • Lipari and Szabo [1982b] G. Lipari and A. Szabo, Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. 2. Analysis of experimental results, J. Amer. Chem. Soc. 104, 4559 (1982b).
  • Singer et al. [2017] P. M. Singer, D. Asthagiri, W. G. Chapman, and G. J. Hirasaki, Molecular dynamics simulations of NMR relaxation and diffusion of bulk hydrocarbons and water, J. Magn. Reson. 277, 15 (2017).
  • Asthagiri et al. [2018] D. Asthagiri, P. M. Singer, A. Valiya Parambathu, Z. Chen, G. J. Hirasaki, and W. G. Chapman, Molecular dynamics simulations of NMR relaxation and diffusion of bulk hydrocarbons, SEG/AAPG/EAGE/SPE Research and Development Petroleum Conference and Exhibition , 101 (2018).
  • Singer et al. [2018a] P. M. Singer, D. Asthagiri, Z. Chen, A. Valiya Parambathu, G. J. Hirasaki, and W. G. Chapman, Role of internal motions and molecular geometry on the NMR relaxation of hydrocarbons, J. Chem. Phys. 148, 164507 (2018a).
  • Asthagiri et al. [2020] D. Asthagiri, W. G. Chapman, G. J. Hirasaki, and P. M. Singer, NMR 1H–1H dipole relaxation in fluids: Relaxation of individual 1H–1H pairs versus relaxation of molecular modes, J. Phys. Chem. B 124, 10802 (2020).
  • Singer et al. [2018b] P. M. Singer, D. Asthagiri, W. G. Chapman, and G. J. Hirasaki, NMR spin-rotation relaxation and diffusion of methane, J. Chem. Phys. 148, 204504 (2018b).
  • Singer et al. [2020a] P. M. Singer, A. Valiya Parambathu, X. Wang, D. Asthagiri, W. G. Chapman, G. J. Hirasaki, and M. Fleury, Elucidating the 1H NMR relaxation mechanism in polydisperse polymers and bitumen using measurements, MD simulations, and models, J. Phys. Chem. B 124, 4222–4233 (2020a).
  • Valiya Parambathu et al. [2020] A. Valiya Parambathu, P. M. Singer, G. J. Hirasaki, W. G. Chapman, and D. Asthagiri, Critical role of confinement in the NMR surface relaxation and diffusion of nn-heptane in a polymer matrix revealed by MD simulations, J. Phys. Chem. B 124, 3801 (2020).
  • Ren and Ponder [2003] P. Ren and J. W. Ponder, Polarizable Atomic Multipole Water Model for Molecular Mechanics Simulation, J. Phys. Chem. B 107, 5933 (2003).
  • Clavaguéra et al. [2006] C. Clavaguéra, F. Calvo, and J.-P. Dognon, Theoretical study of the hydrated Gd3+ ion: Structure, dynamics, and charge transfer, J. Chem. Phys. 124, 074505 (2006).
  • Asthagiri et al. [2004] D. Asthagiri, L. R. Pratt, M. E. Paulaitis, and S. B. Rempe, Hydration structure and free energy of biomolecularly specific aqueous dications, including Zn2+ and first transition row metals, J. Amer. Chem. Soc. 126, 1285 (2004).
  • Eastman et al. [2017] P. Eastman, J. Swails, J. D. Chodera, R. T. McGibbon, Y. Zhao, K. A. Beauchamp, L. P. Wang, A. C. Simmonett, M. P. Harrigan, C. D. Stern, R. P. Wiewiora, B. R. Brooks, and V. S. Pande, OpenMM 7: Rapid development of high performance algorithms for molecular dynamics, PLOS Comp. Bio. 13, e1005659 (2017).
  • Nosé [1984] S. Nosé, A molecular dynamics method for simulation in the canonical ensemble, Mol. Phys. 52, 255 (1984).
  • Hoover [1985] W. G. Hoover, Canonical dynamics: Equilibrium phase-space distribution, Phys. Rev. A 31, 1695 (1985).
  • Zhang et al. [2019] Z. Zhang, X. Liu, K. Yan, M. E. Tuckerman, and J. Liu, Unified efficient thermostat scheme for the canonical ensemble with holonomic or isokinetic constraints via molecular dynamics, J. Phys. Chem. A 123, 6056 (2019).
  • Maginn et al. [2018] E. J. Maginn, R. A. Messerly, D. J. Carlson, D. R. Roe, and J. R. Elliott, Best practices for computing transport properties 1. self-diffusivity and viscosity from equilibrium molecular dynamics [article v1.0], Living J. Comp. Mol. Sc. 1, 1 (2018).
  • Powell et al. [1993] D. H. Powell, A. E. Merbach, G. González, E. Brücher, K. Micskei, M. F. Ottaviani, K. Köhler, A. V. Zelewsky, O. Y. Grinberg, and Y. S. Lebedev, Magnetic‐field‐dependent electronic relaxation of Gd3+ in aqueous solutions of the complexes [Gd(H2O)8]3+, [Gd(propane-1,3-diamine-N,N,N’,N’-tetraacetate)(H2O)2]-, and [Gd(N,N’-bis[(N-methylcarbamoyl)methyl]-3-azapentane-1,5-diamine-3,N,N’-triaetate)(H2O)] of interest in magnetic‐resonance imaging, Helv. Chim. Acta 76, 2129 (1993).
  • Helm and Merbach [1999] L. Helm and A. E. Merbach, Water exchange on metal ions: experiments and simulations, Coord. Chem. Rev. 187, 151 (1999).
  • Abragam [1961] A. Abragam, Principles of Nuclear Magnetism (Oxford University Press, International Series of Monographs on Physics, 1961).
  • McConnell [1987] J. McConnell, The Theory of Nuclear Magnetic Relaxation in Liquids (Cambridge University Press, 1987).
  • Cowan [1997] B. Cowan, Nuclear Magnetic Resonance and Relaxation (Cambridge University Press, 1997).
  • Kimmich [1997] R. Kimmich, NMR Tomography, Diffusometry and Relaxometry (Springer-Verlag, 1997).
  • Peter et al. [2001] C. Peter, X. Daura, and W. F. van Gunsteren, Calculation of NMR-relaxation parameters for flexible molecules from molecular dynamics simulations, J. Biomol. NMR 20, 297 (2001).
  • Case [2002] D. A. Case, Molecular dynamics and NMR spin relaxation in proteins, Acc. Chem. Res. 35, 325 (2002).
  • Singer et al. [2020b] P. M. Singer, A. Arsenault, T. Imai, and M. Fujita, 139La NMR investigation of the interplay between lattice, charge, and spin dynamics in the charge-ordered high-TcT_{c} cuprate La1.875Ba0.125CuO4, Phys. Rev. B 101, 174508 (2020b).
  • Wang et al. [2021] J. Wang, W. Yuan, T. Imai, P. M. Singer, F. Bahrami, and F. Tafti, NMR investigation of the honeycomb irridate Ag3LiIr2O6, Phys. Rev. B 103, 214405 (2021).
  • Imai et al. [2021] T. Imai, P. M. Singer, A. Arsenault, and M. Fujita, Revisiting the 63Cu NMR signature of charge order in La1.875Ba0.125CuO4, J. Phys. Soc. Jap. 90, 034705 (2021).
  • Fordham et al. [2017] E. J. Fordham, L. Venkataramanan, J. Mitchell, and A. Valori, What are, and what are not inverse Laplace transforms, Diffusion Fundam. 29, 1 (2017).
  • Yeh and Hummer [2004] I.-C. Yeh and G. Hummer, System-size dependence of diffusion coefficients and viscosities from molecular dynamics simulations with periodic boundary conditions, J. Phys. Chem. B 108, 15873 (2004).
  • Dünweg and Kremer [1993] B. Dünweg and K. Kremer, Molecular dynamics simulation of a polymer chain in solution, J. Chem. Phys. 99, 6983 (1993).
  • Shikhov and Arns [2016] I. Shikhov and C. Arns, Temperature-dependent oxygen effect on NMR D{D}-T2{T}_{2} relaxation-diffusion correlation of nn-alkanes, Appl. Magn. Reson. 47, 1391 (2016).
  • Madhavi et al. [2017] W. A. M. Madhavi, S. Weerasinghe, and K. I. Momot, Rotational-diffusion propagator of the intramolecular proton - proton vector in liquid water: A molecular dynamics study, J. Phys. Chem. B 121, 10893 (2017).