The hydration shells of phosphate ions and phosphate groups of nucleotides and phospholipid membranes display markedly different structures and hydrogen-bond strengths. Understanding phosphate hydration requires insight into the spatial arrangements of water molecules around phosphates and in thermally activated structure fluctuations on ultrafast time scales. Femtosecond two-dimensional infrared spectroscopy of phosphate vibrations, particularly asymmetric stretching vibrations between 1000 and 1200 cm−1, and ab initio molecular dynamics (AIMD) simulations are combined to map and characterize dynamic local hydration structures and phosphate–water interactions. Phosphoric acid H3PO4 and its anions H2PO4, HPO42−, and PO43− are studied in aqueous environments of different pH value. The hydration shells of phosphates providing OH donor groups in hydrogen bonds with the first water layer undergo ultrafast structural fluctuations, which induce a pronounced spectral diffusion of vibrational excitations on a sub-300 fs time scale. With a decreasing number of phosphate OH groups, the hydration shell becomes more ordered and rigid. The 2D-IR line shapes observed with hydrated PO43− ions display a pronounced inhomogeneous broadening, reflecting a distribution of hydration geometries without fast equilibration. The AIMD simulations allow for an in-depth characterization of the hydration geometries with different numbers of water molecules in the first hydration layer and different correlation functions of the fluctuating electric field that the water environment exerts on the vibrational phosphate oscillators.

Phosphate groups are key structural units and hydration sites of nucleotides, such as adenosine di-and tri-phosphate, the backbone of DNA and RNA, and/or the head groups in phospholipid cellular membranes.1–3 In biochemistry, numerous metabolic and signaling processes rely on phosphorylation reactions for storing and releasing energy.4,5 Beyond their fundamental relevance in chemistry and biology, phosphates have found widespread applications in pH buffering of solutions, materials treatment, fertilizers, or food processing. Monomeric phosphates exist as the charge neutral phosphoric acid H3PO4 or anions H2PO4, HPO42−, and PO43−, depending on the pH value of an aqueous environment. The different species display markedly different hydration structures, which are characterized by a pattern of hydrogen bonds with water molecules in the first solvation shell and electric interactions with solvent molecules beyond the first shell.

Phosphate hydration has been studied in detail by linear vibrational and dielectric spectroscopy.6–10 In parallel, extensive theoretical work based on density functional theory, quantum mechanics/molecular mechanics (QM/MM) methods, and molecular dynamics simulations has addressed the basic molecular interactions and spatial arrangements of solute and solvent molecules.7,11–17 Such work has mainly focused on time-averaged structures and interactions in thermal equilibrium. Hydration dynamics on the intrinsic molecular length and time scales has been addressed for a variety of ions by nonlinear femtosecond infrared spectroscopy.18–23 Recently, such methods have been applied to elucidate structural fluctuations around phosphates and the related fluctuating electric fields on their intrinsic ultrashort time scale.24–27 In a most direct approach, phosphate vibrations serve as specific probes for mapping local interactions and the structure and dynamics of the water shell.

The asymmetric phosphate stretching vibration νAS(PO2) mainly involves displacement of the PO2 unit of the phosphate group and has been used as a local probe in ultrafast infrared spectroscopy28 to study model systems with hydrated PO2 groups, e.g., (CH3)2PO4 or KH2PO4, and larger DNA and RNA backbone structures. Two-dimensional infrared (2D-IR) spectroscopy has given detailed insight into vibrational line shapes, which exhibit a moderate inhomogeneous broadening due to site variations in hydration and line broadening by vibrational dephasing on a 300 fs time scale. Moreover, the formation of solvent-separated and contact ion pairs of PO2 with Na+, Ca2+, and, particularly Mg2+ cations has been detected via characteristic frequency shifts and line shape changes.29–31 The theoretical analysis of such results by DFT cluster calculations and QM/MM simulations of ultrafast molecular dynamics has demonstrated a moderate slowing-down of water dynamics in the first few hydration shells compared to bulk water and fluctuating electric fields with amplitudes of several ten megavolts/cm from the aqueous environment. The first hydration shell around the PO2 unit consists of up to 6 water molecules forming hydrogen bonds with the two oxygen atoms. In contact ion pairs, repulsive short-range interactions of PO2 with the cation lead to a spectral blue shift of the νAS(PO2) absorption band.

There is limited and partly conflicting information on the hydration structure and dynamics of phosphoric acid H3PO4 and its anions H2PO4, HPO42−, and PO43−.7,8,10,12,13,16 So far, the ionic species have mainly been studied in aqueous solutions of sodium and potassium phosphate salts at high concentration, where, beyond separately solvated ions, cation/anion pairs are formed and enhance the structural heterogeneity. Such behavior becomes increasingly complex at high molar concentrations of phosphate ions, with a direct impact on vibrational8 and dielectric spectra.10 

H3PO4, H2PO4, and HPO42− can act as hydrogen bond acceptors and hydrogen bond donors, while PO43− acts as hydrogen bond acceptor only. The impact of such dual hydrogen bond functionality on hydration dynamics is largely unexplored and insight from femtosecond nonlinear vibrational spectroscopy has remained very limited. In this article, we present a systematic 2D-IR study of hydrated phosphoric acid and its anions in the frequency range of the symmetric and asymmetric phosphate stretching vibrations. The relative abundance of the different species is tuned by changing the pH value of diluted aqueous solutions, thus avoiding the issue of ion pair formation. In the samples, there are ∼280 water molecules per phosphate, in order to suppress interactions between solvated ions and provide a sufficient fraction of bulk-like water. The fluctuating structure of the water environment around the phosphates gives rise to spectral diffusion of vibrational excitations, which is mapped onto the 2D-IR line shapes of symmetric and asymmetric phosphate stretching vibrations. The 2D-IR spectra demonstrate a pronounced slowing-down of water shell dynamics when going from H3PO4 to PO43−. The experimental results are complemented by ab initio molecular dynamics simulations of hydration geometries and dynamics for H3PO4 and its anions H2PO4, HPO42−, and PO43− in water.

Solutions of phosphoric acid H3PO4 (Thermo Fisher) in ultrapure water (Rotipuran Ultra, Roth) were prepared with a H3PO4 concentration of c = 0.2M. At this concentration, there are some 280 water molecules per PO4 unit. The large water excess allows for safely suppressing interactions between PO4 moieties at different solvation sites.

In order to control the dissociation equilibria of H3PO4, the pH value of the solutions was tuned by adding NaOH and measured with a pH meter (Hanna Instruments). The respective concentrations of H3PO4, H2PO4, HPO42−, and PO43− presented in Table I were calculated from the dissociation equilibria,
(1)
(2)
(3)
TABLE I.

Concentration of phosphoric acid H3PO4 and the anions H2PO4, HPO42−, and PO43− at different pH values. Bold numbers denote the predominant species. The total concentration of the different HnPO4(3−n)− species (n = 0, …, 3) is c = 0.2M. The pH value was changed by adding NaOH to the aqueous solutions. The respective Na+ ion concentration is given in the last row of the table.

SpeciesConcentration (M)
pH = 1.0pH = 2.3pH = 9.7pH = 13.2
H3PO4 0.187 0.055 <10−10 <10−18 
H2PO4 0.013 0.145 0.0005 <10−7 
HPO42− <10−8 <10−6 0.199 0.025 
PO43− <10−19 <10−16 0.0005 0.175 
Na+ 10−13 2 × 10−12 5 × 10−5 0.158 
SpeciesConcentration (M)
pH = 1.0pH = 2.3pH = 9.7pH = 13.2
H3PO4 0.187 0.055 <10−10 <10−18 
H2PO4 0.013 0.145 0.0005 <10−7 
HPO42− <10−8 <10−6 0.199 0.025 
PO43− <10−19 <10−16 0.0005 0.175 
Na+ 10−13 2 × 10−12 5 × 10−5 0.158 

The concentration of Na+ ions at different pH values is presented at the bottom line of Table I. For pH = 1.0, 9.7, and 13.2, the species H3PO4, HPO42−, and PO43− are predominant, whereas appreciable amounts of both H3PO4 and H2PO4 exist at pH = 2.3. At pH = 13.2, there are similar concentrations of PO43− and Na+ ions, which could result in the formation of different ion pair species, contributing to the spectroscopic observables. The results on PO22−/Na+ contact ion pairs reported in Ref. 30 suggest that the fraction of such ion pairs and their contributions to the linear infrared absorption and 2D-IR spectra is minor for equimolar solutions with c ≈ 0.2M.

Linear infrared absorption spectra were recorded with a Fourier transform infrared spectrometer (Bruker Vertex 80, spectral resolution 2 cm−1). The liquid samples were held in a demountable cell (Harrick) with two 1 mm thick CaF2 windows separated by a Teflon spacer of 12 or 25 µm thickness. The infrared absorption of neat water was measured under the same conditions and subtracted from the spectra of the solutions (cf. Fig. 1).

FIG. 1.

Infrared absorption spectra of aqueous phosphoric acid at different pH values of the solution (solid lines, concentration c = 0.2M). The spectra were measured with a sample thickness of 25 µm and the water background absorption subtracted. The weak band marked with the symbol * in panel (a) is due to the presence of H2PO4 (cf. Table I). The dashed line in panel (b): absorption spectrum of an aqueous solution of KH2PO4 (c = 1.0M), normalized to the peak absorbance of the phosphoric acid sample at 1077 cm−1. The mode assignments are presented in Table II.

FIG. 1.

Infrared absorption spectra of aqueous phosphoric acid at different pH values of the solution (solid lines, concentration c = 0.2M). The spectra were measured with a sample thickness of 25 µm and the water background absorption subtracted. The weak band marked with the symbol * in panel (a) is due to the presence of H2PO4 (cf. Table I). The dashed line in panel (b): absorption spectrum of an aqueous solution of KH2PO4 (c = 1.0M), normalized to the peak absorbance of the phosphoric acid sample at 1077 cm−1. The mode assignments are presented in Table II.

Close modal

2D-IR spectra were measured in 3-pulse photon-echo experiments with heterodyne detection of the nonlinear signal. Infrared pulses with a center frequency of ∼1100 cm−1, a spectral width of ∼190 cm−1 (intensity FWHM), a duration of 120 fs, and a pulse energy of 5 µJ are generated by optical parametric amplification with a 1 kHz repetition rate.32 Two passively phase-stabilized pulse pairs with wavevectors (k1, k2) (mutual delay τ, coherence time) and (k3, kLO) are generated by reflection from a diffractive optic element.

In a box-CARS beam geometry, three pulses are focused into the sample by an off-axis parabolic mirror to generate a photon echo signal. The third-order signal emitted in the phase-matching direction −k1 + k2 + k3 = kSig is overlapped with a local oscillator pulse (kLO) for spectrally resolved detection by a 64 pixel mercury cadmium telluride detector array (frequency ν3, spectral resolution 2 cm−1). Signals measured for different coherence times τ are Fourier-transformed along τ to generate the excitation frequency coordinate ν1. In Fig. 3, the absorptive 2D-IR signal, which is given by the real part of the sum of the rephasing and nonrephasing signals, is plotted as a function of ν1 and ν3. Femtosecond pump–probe experiments were performed in the 2D-IR setup with pulse 1 as the pump and the attenuated pulse 3 as the probe pulse.

Ab initio (first principles) molecular dynamics (AIMD) simulations of H3PO4, H2PO4, HPO42−, and PO43− in water were performed with the CP2K program package (version 8.2).33 AIMD simulations start from a pre-equilibrated box of 64 water molecules (TIP5P force field, cell dimensions 12.4138 × 12.4138 × 12.4138 Å3; the box is taken from Ref. 34) by replacing one water molecule at the center of the simulation box by a H3PO4 molecule, followed by relaxation of the simulation box. Trajectories of the ions H2PO4, HPO42−, and PO43− in water were generated by successively removing a proton after every ∼5 ps of simulation time. After an initial 5 ps equilibration period of every trajectory that was discarded, AIMD simulations cover a simulation time of 80 ps for H3PO4 and HPO42− and 77 ps for H2PO4 and PO43− (total simulation time: 314 ps).

AIMD simulations were performed on the hybrid density functional level of theory using the PW6B95 meta-GGA functional35 that uses 28% of exact exchange in the exchange correlation functional. The PW6B95 functional has shown an excellent performance for the thermodynamics and structure of main group elements36 and yields the radial distribution function of liquid water with excellent accuracy [cf. inlay Fig. 5(e)].

AIMD simulations used a high energy cutoff of 280 a.u. and a relative cutoff of 30 a.u. The screening threshold EPS_SCHWARZ for exact exchange was set to 1.0 × 10−6, and the interaction potential was truncated at 4.5 Å. The TZV2P-GTH basis set was used for hydrogen, oxygen, and phosphorous atoms together with the auxiliary cFit3 basis and the PBE pseudopotential was used for core electrons. Simulations were performed in the NVT ensemble at 300 K using a simulation time step of 0.5 fs and a generalized Langevin equation thermostat (ndim = 5, relaxation time scale 1.0 ps−1). All the AIMD simulations were performed on the CooLMUC-2 Linux Cluster of the Leibniz Supercomputing Centre.

For analysis of the simulation results, radial distribution functions were calculated with the VMD program37 that was also used to produce graphical representations of the solvation shell. Electric fields imposed by the water solvent were evaluated for H3PO4, H2PO4, HPO42−, and PO43− after mapping the real space trajectory to the initial simulation box, centered around the phosphorus atom. The simulation of electric field employs fixed charges of the TIP3P-FB water model,38 thereby neglecting polarization effects in the solvation shell due to interactions with the solute. Simulation with a polarizable model15 has shown that induction contributions to the electric field are a minor contribution (10%–15%) in the solvation forces of phosphates. Moreover, it was demonstrated that the pronounced librational fluctuation dynamics are dominated by the electrostatic contribution to the correlation function. Electric fields have been evaluated at the midpoint of P=O (H3PO4) and P–OH bonds (H3PO4, H2PO4, HPO42−), followed by projection along the bond axis. For PO2 groups of H2PO4, HPO42−, and PO43−, electric fields have been evaluated at the bisector midpoint of the PO2 units, followed by projection along the PO2 bisector axis. This procedure has been shown to quantitatively explain the spectral properties of the asymmetric phosphate stretching vibration νAS(PO2) via the vibrational Stark effect.15 

Infrared absorption spectra of the different phosphate species are shown in Fig. 1. The absorbance A = −log(T) (T: sample transmission) measured with a sample thickness of 25 µm is plotted as a function of wavenumber between 950 and 1250 cm−1 (solid lines). The dashed line in panel (b) gives the absorption spectrum of KH2PO4 dissolved in water for a concentration of c = 1.0M, at which KH2PO4 fully dissociates in hydrated K+ and H2PO4 ions.26 This spectrum is normalized to the peak absorbance at 1077 cm−1 of the H3PO4 solution at pH = 2.3.

The predominant phosphate species at pH = 1.0, 9.7, and 13.2 is H3PO4, HPO42−, and PO43−, respectively (cf. Table I). Their absorption spectra are shown in panels (a), (c), and (d) of Fig. 1. At pH = 2.3 [Fig. 1(b)], there is a mixture of 72.5% H2PO4 and 27.5% H3PO4. The corresponding infrared spectrum displays strong bands of H2PO4 at 1077 and 1160 cm−1 and weak bands of H3PO4 at 1010 cm−1 and between 1130 and 1210 cm−1. The spectrum of the KH2PO4 sample [dashed line in panel (b)] exhibits the H2PO4 bands only. The asymmetric and symmetric P–OH vibrations of H2PO2 are centered at 940 and 880 cm−1, i.e., outside the frequency range shown in Fig. 1(b) and have been analyzed in detail in Ref. 27. The absorption bands of the symmetric and asymmetric phosphate stretching vibrations νS,AS are marked in Fig. 1 and the respective frequencies are presented in Table II. The symmetric stretching mode νS of H3PO4 shows a very small transition dipole but is clearly identified in the Raman spectrum.8 For the phosphate species H3PO4, H2PO4, HPO42−, and PO43−, we observe a successive redshift from 1170 cm−1 [ν(P = O)] to 1008 cm−1AS(PO4)], reflecting the weakening of the bonding potential upon the increase in negative charge density. In the present study, we focus on the asymmetric phosphate stretching vibrations νAS as probes of hydration dynamics.

TABLE II.

Symmetric (νS) and asymmetric (νAS) phosphate stretching frequencies of phosphoric acid and ions at different pH values.

SpeciesVibrational frequency (cm−1)
νSνAS
H3PO4 890 (Raman)a 1010 
H2PO4 1077 1158 
HPO42− 990 1080 
PO43− 936b 1008 
SpeciesVibrational frequency (cm−1)
νSνAS
H3PO4 890 (Raman)a 1010 
H2PO4 1077 1158 
HPO42− 990 1080 
PO43− 936b 1008 
a

Reference 9.

b

Reference 7.

Population relaxation of the phosphate stretching vibrations was studied in spectrally and temporally resolved pump–robe measurements. The data presented in Fig. 2 allow for estimating the decay times of the v = 1 states of the νAS vibrations of H3PO4 and PO43− from the fast recovery of the absorption decrease at νpr = 1030 cm−1, which is caused by the bleaching of the v = 0 → 1 absorption and stimulated emission on the v = 1 → 0 transition.
FIG. 2.

Femtosecond pump–probe spectra of the asymmetric phosphate stretching vibration of (a) H3PO4 and (b) PO43− (colored lines). The change of absorbance ΔA = −log(T/T0) is plotted as a function of probe frequency νpr (T, T0: sample transmission with and without excitation). The prominent absorption decrease is due to ground-state bleaching and stimulated emission on the v = 0 → 1 vibrational transitions. The dashed lines represent the infrared absorption spectrum (in arbitrary units). (c) Time-resolved absorbance changes ΔA at a fixed probe frequency of νpr = 1030 cm−1 (symbols). The solid lines are numerical fits to the data, giving an initial decay time of 390 fs for H3PO4 and 200 fs for PO43−. The fast decays reflect the population relaxation from the v = 1 state, while the residual long-lived signals are due to vibrational cooling.

FIG. 2.

Femtosecond pump–probe spectra of the asymmetric phosphate stretching vibration of (a) H3PO4 and (b) PO43− (colored lines). The change of absorbance ΔA = −log(T/T0) is plotted as a function of probe frequency νpr (T, T0: sample transmission with and without excitation). The prominent absorption decrease is due to ground-state bleaching and stimulated emission on the v = 0 → 1 vibrational transitions. The dashed lines represent the infrared absorption spectrum (in arbitrary units). (c) Time-resolved absorbance changes ΔA at a fixed probe frequency of νpr = 1030 cm−1 (symbols). The solid lines are numerical fits to the data, giving an initial decay time of 390 fs for H3PO4 and 200 fs for PO43−. The fast decays reflect the population relaxation from the v = 1 state, while the residual long-lived signals are due to vibrational cooling.

Close modal
A numerical fit gives a value of 390 fs for H3PO4 and of 200 fs for PO43−. The νAS vibration of H2PO4 displays a v = 1 lifetime of 330 fs (cf. Fig. 3 of Ref. 26). For HPO42−, the lifetime is on the order of 300 fs, as suggested by the amplitudes of the 2D-IR signal measured at different waiting times T(cf. Fig. 3). The long-lived absorption changes extending into the picosecond time range are due to the slower dissipation of vibrational excess energy in the solution, a well-characterized process typically extending over tens of picoseconds.24,39
FIG. 3.

Infrared absorption and 2D-IR spectra of (a), (e), and (i) H3PO4; (b), (f), and (j) H2PO4; (c), (g), and (k) HPO42−; and (d), (h), and (l) PO43− in water (cf. Table I for concentrations). In the 2D-IR spectra, the absorptive 2D signal recorded at a waiting time T is plotted as a function of excitation frequency ν1 (ordinate) and detection frequency ν3 (abscissa). The yellow-red contours close to the frequency diagonal represent positive 2D signals on the v = 0 → 1 transitions of the vibrations, while the negative blue contours originate from v = 1 → 2 transitions. The 2D signal changes by 7.5% between neighboring contour lines. The blue solid lines represent the center lines of the positive 2D signals due to the asymmetric phosphate stretching modes. The spectra in panels (b) and (f) measured at a sample pH = 2.3 display contributions from H3PO4 and H2PO4. The 2D-IR spectrum in panel (j) was measured with KH2PO4 in water (c = 1.0M, taken from Ref. 18). The corresponding infrared absorption spectrum is shown in panel (b) (dashed line).

FIG. 3.

Infrared absorption and 2D-IR spectra of (a), (e), and (i) H3PO4; (b), (f), and (j) H2PO4; (c), (g), and (k) HPO42−; and (d), (h), and (l) PO43− in water (cf. Table I for concentrations). In the 2D-IR spectra, the absorptive 2D signal recorded at a waiting time T is plotted as a function of excitation frequency ν1 (ordinate) and detection frequency ν3 (abscissa). The yellow-red contours close to the frequency diagonal represent positive 2D signals on the v = 0 → 1 transitions of the vibrations, while the negative blue contours originate from v = 1 → 2 transitions. The 2D signal changes by 7.5% between neighboring contour lines. The blue solid lines represent the center lines of the positive 2D signals due to the asymmetric phosphate stretching modes. The spectra in panels (b) and (f) measured at a sample pH = 2.3 display contributions from H3PO4 and H2PO4. The 2D-IR spectrum in panel (j) was measured with KH2PO4 in water (c = 1.0M, taken from Ref. 18). The corresponding infrared absorption spectrum is shown in panel (b) (dashed line).

Close modal

The symmetric and asymmetric stretching vibrations of PO2 groups in (CH3)2PO4, KH2PO4, DNA, and RNA embedded in an aqueous environment display lifetimes between 250 and 400 fs, as has been discussed in Refs. 26 and 28. The fast population decays of phosphate stretching vibrations and their independence from the particular molecular species and environment point to a relaxation via a population transfer to PO4 vibrations at lower frequencies, which are anharmonically coupled to the stretching vibrations.

The 2D-IR spectra of the different phosphate species are shown in Fig. 3. The spectra were measured with phosphoric acid samples at different pH values, except for panel (j), which displays the 2D-IR spectrum of a KH2PO4 sample.26 In panels (e)–(l), the absorptive 2D-IR signal is plotted as a function of excitation frequency ν1 and detection frequency ν3 for different waiting times T. At the early waiting time T = 300 fs [panels (e)–(h), (j)], there is a negligible temporal overlap of the three femtosecond pulses in the photon-echo experiment. The range toward longer waiting times is limited to some 600 fs [panels (i), (k), (l)] because of the femtosecond population decay times of the vibrational excitations. The 2D-IR spectra display diagonal peaks at detection frequencies around the frequency positions of all the vibrational absorption bands in the linear infrared absorption spectra [panels (a)–(d); please note the different ν3 ranges of the spectra of different phosphate species (columns)].

The yellow-red contours of the diagonal peaks are due to excitations on the v = 0 → 1 vibrational transitions (positive 2D-IR signal), while the blue contours originate from v = 1 → 2 excitations (negative 2D-IR signal). The shift of the blue contours to smaller detection frequencies ν3 reflects the diagonal anharmonicity of the vibrations. Slices of the 2D-IR peaks of the asymmetric phosphate stretching excitations νAS along the detection frequency ν3 are shown in Fig. 4. The maxima of the positive 2D-IR signals arise at a higher detection frequency ν3 than the maxima of the corresponding linear absorption bands. Such frequency shifts are due to the strong spectral overlap of the positive v = 0 → 1 and the negative v = 1 → 2 components of the 2D-IR signal, leading to a pronounced compensation of positive and negative signals around the zero crossings.
FIG. 4.

Slices of the 2D-IR spectra recorded at a waiting time T = 300 fs (solid lines) along detection frequency ν3 for the excitation frequencies ν1 given in the different panels. Panel (d) includes a slice for T = 500 fs (dashed–dotted line). The 2D-IR signal is normalized to the maximum positive value. The strong positive and negative peaks originate from the diagonal peaks due to the νAS vibrations, with the positive component representing a signal on the v = 0 → 1 transition and the negative component on the v = 1 → 2 transition. The arrows mark the frequency positions of the related maximum in the linear absorption spectrum.

FIG. 4.

Slices of the 2D-IR spectra recorded at a waiting time T = 300 fs (solid lines) along detection frequency ν3 for the excitation frequencies ν1 given in the different panels. Panel (d) includes a slice for T = 500 fs (dashed–dotted line). The 2D-IR signal is normalized to the maximum positive value. The strong positive and negative peaks originate from the diagonal peaks due to the νAS vibrations, with the positive component representing a signal on the v = 0 → 1 transition and the negative component on the v = 1 → 2 transition. The arrows mark the frequency positions of the related maximum in the linear absorption spectrum.

Close modal
The pump–probe spectra shown in Figs. 2(a) and 2(b) exhibit the same behavior. In other words, the spectral widths of the positive and negative signals are larger than the diagonal anharmonicity of the vibrations, the latter being in a range of 10–20 cm−1.24,26 We should note the increased anharmonicity of the P=O mode of H3PO4 [Fig. 3(e), ν3 ≅ 1170 cm−1] that reflects a change in the mode character, being an isolated local P=O oscillator for H3PO4 but delocalized asymmetric/symmetric modes with smaller diagonal anharmonicities in H2PO4, HPO42−, and PO43−.

The 2D-IR line shapes of the v = 0 → 1 signals (yellow-red peaks) display elliptical contours with a different tilting angle relative to the frequency diagonal ν1 = ν3. To quantify the orientation of the elliptic peaks, we derive center lines (CL), which connect the frequency positions of the maximum signals in (horizontal) cuts along the detection frequency axis ν3 at a fixed excitation frequency ν1.40 The thick blue lines shown in Fig. 3 give the CL for the asymmetric stretching modes νAS. The inverse slope of such lines, the centerline slope (CLS), is a measure for the correlation of excitation (ν1) and detection (ν3) frequencies. A CLS = 1 corresponds to CL parallel to the ν1 = ν3 frequency diagonal and, thus a perfect correlation of ν1 and ν3, such as in the case of static inhomogeneous broadening. In contrast, a CLS = 0 corresponds to vertical CL (parallel to the ν1 axis) and total loss of correlation, such as induced by spectral diffusion.

The CL of the H3PO4 spectra shown in Figs. 3(e) and 3(i) are vertical within the experimental accuracy, corresponding to CLS≈0. This observation points to a very fast spectral diffusion within the initial time window of up to T = 300 fs. The 2D-IR peak of H2PO4 at (ν1, ν3)=(1156, 1159) cm−1 [T = 300 fs, Fig. 3(f)] is superimposed by a broad component originating from H3PO4, which affects the CLS. Thus, we analyze the 2D-IR spectrum of KH2PO4 [Fig. 3(j)], displaying a small CLS ≈ 0.25 ± 0.05/0.1 of the νAS peak and again pointing to pronounced spectral diffusion within the first 300 fs. The results presented in Ref. 26 suggest a predominant 50 fs decay of the frequency-fluctuation correlation function, which governs the experimental 2D-IR line shape (Fig. 6(b) of Ref. 26).

The CLS for HPO42− [Figs. 3(g) and 3(k)] has a value of 0.36 ± 0.05, whereas the PO43− peaks at T = 300 and 500 fs display a CLS = 0.62 ± 0.05. The frequency cuts of the PO43− spectra shown in Fig. 4(d) exhibit a similar width of the positive 2D-IR signal at T = 300 and 500 fs, pointing to a limited spectral diffusion up to 500 fs. A similar behavior is borne out by cuts along the diagonal ν1 = ν3 for the two waiting times (not shown). The CLS for the different species suggest that the loss of frequency correlation becomes slower with a decreasing number of hydrogen atoms attached to the PO4 unit, which corresponds to a decreasing number of OH donor groups from the phosphate in hydrogen bonds with the first water solvation shell. Moreover, the Coulomb interaction between the phosphate and the water shell becomes stronger with increasing ion charge. Such two mechanisms appear to make the water environment more rigid and to slow down spectral diffusion via the fluctuating electric force from the solvent.

The time averaged hydration structure of phosphoric acid H3PO4 and its anions H2PO4, HPO42−, and PO43− obtained from high-level AIMD simulations is shown in Fig. 5. The solvation shell is quantified via the phosphorous–water–oxygen (P…Ow) radial distribution function [rdf, Fig. 5(e)]. Representative snapshots of first hydration shell geometries are shown in Figs. 5(a)5(d). For H3PO4, the average simulated phosphorous-oxygen bond lengths are 1.491 Å (P=O) and 1.572 Å (P–OH), in excellent agreement with bond lengths derived from EXAFS experiments41 (1.49 and 1.54 Å, respectively). Similar good agreement is found for all ions (e.g., PO43−: mean P–O distance 1.558 Å; exp: 1.54 Å), suggesting a high quality of AIMD simulations of phosphates in liquid water with the PW6B95 functional.

FIG. 5.

Hydration structure of (a) H3PO4 and the anions (b) H2PO4, (c) HPO42−, and (d) PO43−. Representative snapshots of the first hydration shells are shown (atoms within 4.5 Å of the phosphate compound); hydrogen bonds are indicated by the dashed lines. (e) Radial distribution functions g(r) of phosphorous P–water oxygen Ow distances of H3PO4, H2PO4, HPO42−, and PO43−. (f) Radial distribution functions g(r) of hydroxy oxygen atoms OH–water oxygen Ow distances of H3PO4, H2PO4, and HPO42−. (g) Radial distribution functions g(r) of ester oxygen atoms P=O–water oxygen Ow distances of H3PO4, H2PO4, HPO42−, and PO43−. The inlay of panel (e) shows the oxygen–oxygen radial distribution functions g(r) of neat water simulated with the PW6B95 functional, experimental reference from Ref. 42.

FIG. 5.

Hydration structure of (a) H3PO4 and the anions (b) H2PO4, (c) HPO42−, and (d) PO43−. Representative snapshots of the first hydration shells are shown (atoms within 4.5 Å of the phosphate compound); hydrogen bonds are indicated by the dashed lines. (e) Radial distribution functions g(r) of phosphorous P–water oxygen Ow distances of H3PO4, H2PO4, HPO42−, and PO43−. (f) Radial distribution functions g(r) of hydroxy oxygen atoms OH–water oxygen Ow distances of H3PO4, H2PO4, and HPO42−. (g) Radial distribution functions g(r) of ester oxygen atoms P=O–water oxygen Ow distances of H3PO4, H2PO4, HPO42−, and PO43−. The inlay of panel (e) shows the oxygen–oxygen radial distribution functions g(r) of neat water simulated with the PW6B95 functional, experimental reference from Ref. 42.

Close modal

For H3PO4, the P…Ow rdf shows a moderately structured first solvation shell (P…Ow distance <4.0 Å) and negligible long-range structure beyond the first hydration shell (P…Ow distances >4.0 Å). The first solvation shell is made up of 4.7 water molecules (integrated first maximum of P…Ow rdf). For the ions H2PO4 and HPO42−, the solvation shell is strikingly similar and substantially more structured than for H3PO4. The first hydration layer is composed of 8 and 9 water molecules, respectively (integrated first maximum of rdf P…Ow until 4.5 Å), and a pronounced second maximum appears in the rdf. For PO43−, the first hydration layer is composed of 13 water molecules (integrated first maximum of rdf P…Ow until 4.5 Å). As reflected in the different height of the first peak of the rdf of HPO42− and PO43−, the first solvation shell of PO43− contains a substantially larger number of water molecules. On average, each of the P=O units is surrounded by about 3 water molecules in a tetrahedral arrangement. Apparent is a long-range structure in the hydration shell around PO43− with a pronounced second minimum and recognizable third maximum in the P…Ow rdf (P…Ow distance ∼8.2 Å, not shown). Comparison of the first maxima of the P…Ow rdf yields that maxima are shifted to slightly larger P…Ow distances in the order H3PO4, H2PO4, HPO42−, and PO43− (3.68, 3.71, 3.76, and 3.78 Å, respectively), presumably due to increasing steric hindrance with increasing water content of the first hydration shell.

Comparing H3PO4 and its anions H2PO4, HPO42−, and PO43−, we find increasingly long-range structured hydration shells around the phosphates with increasing Coulomb interactions, i.e., phosphate ion charge (Fig. 5), which also increases the structuring in Ow…Ow rdf in comparison with neat water. Such increased water structuring correlates with the structure making properties of the highly charged ions HPO42− and PO43− in the Hofmeister series.

The rdf of hydroxy oxygen–water oxygen atoms [POH…Ow rdf, Fig. 5(f)] characterizes the water population at the hydroxy group and the hydrogen bond donor properties of H3PO4, H2PO4, and HPO42−. For H3PO4, we find particularly strong hydrogen bonds with the first hydration shell [first maximum of POH…Ow rdf = 2.56 Å]. For H2PO4 and HPO42−, the first maxima of the POH…Ow rdf are shifted to larger distances (2.68 and 2.76 Å) that become comparable in strength with the water–water hydrogen bond [2.78 Å, inlay Fig. 5(e)]. The population of the first hydration shell is lower for H2PO4 than for HPO42−.

The rdf of phosphate oxygen–water oxygen atoms (P=O…Ow rdf) were calculated to quantify the hydrogen bond strength of P=O groups and water molecules [Fig. 5(g)]. We find an increasing hydrogen bond strength in the order H3PO4, H2PO4, HPO42−, and PO43− with respective P=O⋯OH2 distances of 2.76, 2.74, 2.72, and 2.66 Å (first maximum of P=O…Ow rdf). On average, the P=O unit of H3PO4 is populated by about 2.1 water molecules (integrated first maximum of rdf P=O…Ow until 3.0 Å), while the P=O units of the ions H2PO4, HPO42−, and PO43− are populated by 2.5–2.75 water molecules.

Compared to previous ab initio MD simulations,16 a very similar hydration structure and comparable number of water molecules is found in the first solvation shell of the ions H2PO4, HPO42−, and PO43−, but a substantially reduced water population of the P=O unit of H3PO4 is observed in the current study. A possible origin of such discrepancies is given by the current higher level, hybrid density functional treatment, compared to a local GGA functional in Ref. 16. The hybrid density functional level of theory is found to provide a substantially improved water structure [inlay, Fig. 5(e)] and P=O bond length in H3PO4 [Figs. 6(a) and 6(b)]. Such balanced description of interactions is expected to be particularly relevant for describing the hydration shell of H3PO4, thus, the presented simulations are considered more reliable.

FIG. 6.

Time evolution and distribution of phosphorous–oxygen bond length of (a) and (b) H3PO4, (c) and (d) H2PO4, (e) and (f) HPO42−, and (g) and (h) PO43−. The P=O bond length of H3PO4 and P–O bond length of the POx groups of the ions H2PO4, HPO42−, and PO43− are shown.

FIG. 6.

Time evolution and distribution of phosphorous–oxygen bond length of (a) and (b) H3PO4, (c) and (d) H2PO4, (e) and (f) HPO42−, and (g) and (h) PO43−. The P=O bond length of H3PO4 and P–O bond length of the POx groups of the ions H2PO4, HPO42−, and PO43− are shown.

Close modal

While PO43− has tetrahedral symmetry in the gas phase, instantaneous geometries in a fluctuating liquid are subject to symmetry distortions due to stochastic interactions with the solvation shell. Figure 6 shows the fluctuations of bond length of the POx groups (x = 1, ..., 4) of H3PO4 and the ions H2PO4, HPO42−, and PO43−. Figures 6(b), 6(d), 6(f), and 6(h) show a shift of the bond length distributions to longer phosphorous oxygen distances with increasing negative charge density, reflecting the gradual weakening of bond order from a P=O to a P–O unit. For a particular solvent configuration, the oxygen atoms of the POx groups are not equivalent but differ due to the local environment. Nevertheless, the time evolution of the two PO2 bond length of H2PO4 [Fig. 6(c)] is strongly (anti-)correlated, reflecting the superimposed motion of symmetric and asymmetric PO2 stretching vibrations. For HPO42− and PO43− [Figs. 6(e) and 6(g)], the correlated time evolution of two oxygens is preserved on short time scales (∼200 fs), while the bond length of the additional oxygen atoms of the POx unit is decoupled from the motion of the PO2 unit, as reflected in the larger amplitude bond length excursions of a single alternating bond phosphorous oxygen bond length. In other words, the symmetry of degenerate PO2 units is broken in HPO42− and PO43−, leading to decoupling of P–O and PO2 units. The latter are randomized by the fluctuating forces arising from librational motions and hydrogen bond rearrangements in the solvation shell, eventually making the four oxygens indistinguishable on a time scale of tens of picoseconds.

The experimentally observed differences in the spectral diffusion properties of H3PO4 and the anions H2PO4, HPO42−, and PO43− are related to fluctuations and inhomogeneities of the hydration shells, sensed via electrostatic coupling of the phosphate vibrational oscillator with the solvent environment. Thus, the electric field acting on the phosphates represents a collective coordinate of solvation, allowing a characterization of the hydration dynamics. We have calculated the fluctuations of the electric field at the bond midpoint, projected on the bond axis of P=O and P–OH bonds (H3PO4, H2PO4, and HPO42−) and at the bisector midpoint of PO2 groups (H2PO4, HPO42−, and PO43−) [see Fig. 7(a)]. Electric field fluctuation amplitudes and electric field correlation functions are shown in Fig. 7. As demonstrated in Fig. 6, the gas phase tetrahedral symmetry of PO43− is broken in a fluctuating liquid, leading to a localization of modes in PO2 units, a behavior similar to other tetrahedral ions.43 We have thus chosen to evaluate the electric field at the bisector midpoint of successive PO2 groups of H2PO4, HPO42−, and PO43−.

FIG. 7.

(a) Schematic of the bond midpoint (yellow) along the bond axis of P=O in H3PO4, and the bisector midpoints (yellow) of PO2 groups (H2PO4, HPO42−, and PO43−) used for evaluation of the electric field. Electric fields were projected along the bisector axis (dashed lines). Electric field fluctuations of H3PO4 (b) and (c) distribution of electric fields imposed on the midpoint of the P=O bond. (d) Electric field fluctuation correlation function for P=O and P–OH bonds of H3PO4. (e)–(g) Electric field fluctuation correlation functions for H2PO4, HPO42−, and PO43−. Panels (d) and (e) show the correlation functions for the P=O and PO2 groups (solid blue lines), the average of the P–OH bonds (red solid lines) and for individual P–OH bonds of H3PO4 and H2PO4 (dashed lines). Panels (f) and (g) show the average correlation functions of PO2 groups (solid blue lines) and of individual PO2 groups (dashed lines) of HPO42− and PO43−.

FIG. 7.

(a) Schematic of the bond midpoint (yellow) along the bond axis of P=O in H3PO4, and the bisector midpoints (yellow) of PO2 groups (H2PO4, HPO42−, and PO43−) used for evaluation of the electric field. Electric fields were projected along the bisector axis (dashed lines). Electric field fluctuations of H3PO4 (b) and (c) distribution of electric fields imposed on the midpoint of the P=O bond. (d) Electric field fluctuation correlation function for P=O and P–OH bonds of H3PO4. (e)–(g) Electric field fluctuation correlation functions for H2PO4, HPO42−, and PO43−. Panels (d) and (e) show the correlation functions for the P=O and PO2 groups (solid blue lines), the average of the P–OH bonds (red solid lines) and for individual P–OH bonds of H3PO4 and H2PO4 (dashed lines). Panels (f) and (g) show the average correlation functions of PO2 groups (solid blue lines) and of individual PO2 groups (dashed lines) of HPO42− and PO43−.

Close modal

The trajectory of the electric field imposed on the P=O bond of H3PO4 [Fig. 7(b)] reveals fast and large amplitude, sub-100 fs fluctuations due to librational motions of water molecules together with slower modulations on the ∼1–10 ps time scale. On average, the imposed electric field is substantial [84 MV/cm, Fig. 7(c)] with large amplitude modulations reaching up to 200 MV/cm, providing a rationale for the modification of the gas phase molecular structure of H3PO4 in the aqueous solution, i.e., the contraction of P=O and P–OH bond distances from 1.54 and 1.69 Å16 to 1.491 and 1.572 Å in the liquid. The temporal behavior of field fluctuations is characterized via the electric field fluctuation correlation function [Fig. 7(d)]. For the P=O bond of H3PO4, the fast (∼100 fs) decay is followed by a slower decay extending into the picosecond time scale. In comparison, the amplitude of the initial decay of the field correlation function for the P–OH groups of H3PO4 is more pronounced, and slower dynamics on the 400–1000 fs time scale almost negligible. The weak residual inhomogeneity among the three P–OH sites stems from the strong hydrogen bonds of hydroxy groups, with the first hydration shell that lock the conformation of the hydroxy groups on the low picosecond time scale.

The electric field fluctuation correlation functions of H2PO4 and HPO42− [Figs. 7(e) and 7(f)] are very similar; for both ions, the correlation function recorded at PO2 groups shows a longer lived plateau extending into the picosecond time scale due to the inhomogeneity of hydrogen bond geometries, while the correlation functions recorded at P–OH sites display a fast (∼100 fs) decay.

The averaged electric field fluctuation correlation function recorded at the bisector midpoints of the PO2 units of PO43− [solid line in Fig. 7(g)] shows an initial fast (50–100 fs) decay, followed by slower dynamics on the 400 fs to 1.2 ps time scale. Notably, the different PO2 oscillators exhibit a similar temporal behavior, but the amplitudes of the fast and slower time scale in the individual correlation function of different PO2 sites vary [dashed lines in Fig. 7(g)]. The varying amplitudes of the slow temporal component in the correlation functions reflect inhomogeneities of the water population and water hydrogen bond geometries at the different sites of the PO43− tetrahedron leading to an inhomogeneity persisting up to picoseconds in the PO43− case.

Comparing the correlation decays of H3PO4 and the anions H2PO4, HPO42−, and PO43−, pronounced initial fast (50–100 fs) decay is followed by varying degrees of slower dynamics. The fast dynamics arise from librational motion predominantly of first solvation shell water molecules15 for a fixed hydrogen bond geometry. The slower (∼1 ps) time scale leads to a decay of residual correlation of correlation functions due to the randomization of hydrogen bond geometries that interchange on the 1–20 ps time scale and impose particularly pronounced inhomogeneities in the PO43− case. The breaking and reformation of hydrogen bonds involves large angular jumps of water molecules stochastically occurring in the hydration shell.44 The dual hydrogen bond functionality of H3PO4, H2PO4, and HPO42 is an additional source of fluctuations of the vibrational Hamiltonian due to the strong coupling of P–OH and PO2 mode displacements. The detailed mechanism of vibrational couplings among these modes has been investigated in Ref. 26. In essence, the fast (quasi-homogeneous) fluctuations at a hydrogen bond donor site [cf. red lines shown in Figs. 7(e) and 7(f)] fully propagate into fluctuations of νAS(PO2) vibrations, in addition to the field fluctuations experienced at PO2 units. This behavior explains the absence of major inhomogeneities in the 2D-IR spectra of H2PO4 and HPO42−. In contrast, due to the absence of H-bond donor functionality, the PO2 units of PO43− most clearly reveal inhomogeneities in the hydrogen bond arrangement around the ions with picosecond lifetime.

The experiments and simulations presented here are in the limit of independent phosphate hydration sites, i.e., interactions between neighboring sites play a negligible role. When increasing phosphate concentrations to the molar range, hydration shells of neighboring sites begin to interfere in space and influence each other. This scenario is particularly relevant for solvated PO43− ions, where the first two hydration layers contain up to 40 water molecules in a structured arrangement.10 At high PO43− concentrations, one expects a replacement of bulk-like water by an arrangement of solvent molecules structured by the ions.

It is interesting to compare the hydration dynamics of the tetrahedral ions PO43− and SO42− in the femtosecond time domain. While the current study reveals inherent inhomogeneities of hydrogen bond arrangement and hydration structure, inhomogeneities of the hydration shell of SO42− on the ∼500 fs time scale are only observed in the presence of Mg2+ ions (but not for singly charged Na+ and NH4+ ions).43 The tightly bound solvation shell or PO43− thus contrasts with a more labile and fluxional solvation shell of SO42−, where inhomogeneity is only induced via the formation of solvent shared ions pairs with Mg2+ ions.

In conclusion, the combination of femtosecond 2D-IR spectroscopy and ab initio molecular dynamics simulations reveals distinctly different hydration structures and dynamics of phosphoric acid H3PO4 and its anions H2PO4, HPO42−, and PO43−. The first solvent shells around the solutes contain between 6 and 13 water molecules and a molecular arrangement with phosphate OH groups acting as hydrogen bond donors to water oxygens. The next two water layers show a different degree of structural ordering and rigidity. The latter is most pronounced around PO43− ions. As a result of such structural differences, the fluctuating electric forces the water environment exerts on the vibrational phosphate oscillators display different time correlation functions. The predominant fast (50 fs) correlation decays around H3PO4 and H2PO4- result in a quasi-homogeneous 2D-IR line shape of the asymmetric phosphate stretching vibration. The slower picosecond correlation decay components found for HPO42− and PO43− point to an incomplete randomization of electric forces. For PO43− ions, such components manifest in an incomplete spectral diffusion on a 600 fs time scale and partly inhomogeneous 2D-IR line shapes.

This research received funding from the European Research Council (ERC) under European Union’s Horizon 2020 research and innovation program (Grant Agreements Nos. 833365 and 802817). The authors acknowledge the computational and data resources provided by the Leibniz Supercomputing Centre (www.lrz.de).

The authors have no conflicts to disclose.

T.E., A.K., and B.P.F. conceived the study. A.K. performed the experiments. A.K. and TE analyzed the experimental results. B.P.F. performed the theoretical calculations and simulations. The manuscript was written by all authors.

Achintya Kundu: Conceptualization (lead); Formal analysis (equal); Investigation (equal); Project administration (lead); Supervision (lead); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Benjamin P. Fingerhut: Conceptualization (equal); Data curation (equal); Funding acquisition (equal); Investigation (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Thomas Elsaesser: Conceptualization (lead); Formal analysis (equal); Investigation (equal); Project administration (lead); Supervision (lead); Validation (equal); Writing – original draft (lead); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
D.
Laage
,
T.
Elsaesser
, and
J. T.
Hynes
, “
Water dynamics in the hydration shells of biomolecules
,”
Chem. Rev.
117
,
10694
10725
(
2017
).
2.
B.
Schneider
,
K.
Patel
, and
H. M.
Berman
, “
Hydration of the phosphate group in double-helical DNA
,”
Biophys. J.
75
,
2422
2434
(
1998
).
3.
P.
Ball
, “
Water as an active constituent in cell biology
,”
Chem. Rev.
108
,
74
108
(
2008
).
4.
M. K.
Tarrant
and
P. A.
Cole
, “
The chemical biology of protein phosphorylation
,”
Annu. Rev. Biochem.
78
,
797
825
(
2009
).
5.
T.
Hunter
, “
Signaling—2000 and beyond
,”
Cell
100
,
113
127
(
2000
).
6.
A. C.
Chapman
and
L. E.
Thirlwell
, “
Spectra of phosphorus compounds—I the infra-red spectra of orthophosphates
,”
Spectrochim. Acta
20
,
937
947
(
1964
).
7.
C. C.
Pye
and
W. W.
Rudolph
, “
An ab initio, infrared, and Raman investigation of phosphate ion hydration
,”
J. Phys. Chem. A
107
,
8746
8755
(
2003
).
8.
W. W.
Rudolph
and
G.
Irmer
, “
Raman and infrared spectroscopic investigations on aqueous alkali metal phosphate solutions and density functional theory calculations of phosphate—Water clusters
,”
Appl. Spectrosc.
61
,
1312
1327
(
2007
).
9.
W. W.
Rudolph
, “
Raman- and infrared-spectroscopic investigations of dilute aqueous phosphoric acid solutions
,”
Dalton Trans.
39
,
9642
9653
(
2010
).
10.
A.
Eiberweiser
,
A.
Nazet
,
G.
Hefter
, and
R.
Buchner
, “
Ion hydration and association in aqueous potassium phosphate solutions
,”
J. Phys. Chem. B
119
,
5270
5281
(
2015
).
11.
M.
Klähn
,
G.
Mathias
,
C.
Kötting
,
M.
Nonella
,
J.
Schlitter
,
K.
Gerwert
, and
P.
Tavan
, “
IR spectra of phosphate ions in aqueous solution: Predictions of a DFT/MM approach compared with observations
,”
J. Phys. Chem. A
108
,
6186
6194
(
2004
).
12.
C.
Ebner
,
U.
Onthong
, and
M.
Probst
, “
Computational study of hydrated phosphate anions
,”
J. Mol. Liq.
118
,
15
25
(
2005
).
13.
A. B.
Pribil
,
T. S.
Hofer
,
B. R.
Randolf
, and
B. M.
Rode
, “
Structure and dynamics of phosphate ion in aqueous solution: An ab initio QMCF MD study
,”
J. Comput. Chem.
29
,
2330
2334
(
2008
).
14.
J.
VandeVondele
,
P.
Tröster
,
P.
Tavan
, and
G.
Mathias
, “
Vibrational spectra of phosphate ions in aqueous solution probed by first-principles molecular dynamics
,”
J. Phys. Chem. A
116
,
2466
2474
(
2012
).
15.
B. P.
Fingerhut
,
R.
Costard
, and
T.
Elsaesser
, “
Predominance of short range Coulomb forces in phosphate-water interactions—A theoretical analysis
,”
J. Chem. Phys.
145
,
115101
(
2016
).
16.
S.
Borah
, “
Hydration properties of HnPO4n−3 (n = 0–3) from ab initio molecular dynamics simulations
,”
J. Phys. Chem. B
124
,
5454
5464
(
2020
).
17.
B.
Sharma
and
A.
Chandra
, “
Ab initio molecular dynamics simulation of the phosphate ion in water: Insights into solvation shell structure, dynamics, and kosmotropic activity
,”
J. Phys. Chem. B
121
,
10519
10529
(
2017
).
18.
P.
Hamm
,
M.
Lim
, and
R. M.
Hochstrasser
, “
Non-Markovian dynamics of the vibrations of ions in water from femtosecond infrared three-pulse photon echoes
,”
Phys. Rev. Lett.
81
,
5326
5329
(
1998
).
19.
D. E.
Moilanen
,
D.
Wong
,
D. E.
Rosenfeld
,
E. E.
Fenn
, and
M. D.
Fayer
, “
Ion–water hydrogen-bond switching observed with 2D IR vibrational echo chemical exchange spectroscopy
,”
Proc. Natl. Acad. Sci. U. S. A.
106
,
375
380
(
2009
).
20.
K. J.
Tielrooij
,
N.
Garcia-Araez
,
M.
Bonn
, and
H. J.
Bakker
, “
Cooperativity in ion hydration
,”
Science
328
,
1006
1009
(
2010
).
21.
K.
Ohta
,
J.
Tayama
,
S.
Saito
, and
K.
Tominaga
, “
Vibrational frequency fluctuation of ions in aqueous solutions studied by three-pulse infrared photon echo method
,”
Acc. Chem. Res.
45
,
1982
1991
(
2012
).
22.
D.
Czurlok
,
M.
von Domaros
,
M.
Thomas
,
J.
Gleim
,
J.
Lindner
,
B.
Kirchner
, and
P.
Vöhringer
, “
Femtosecond 2DIR spectroscopy of the nitrile stretching vibration of thiocyanate anions in liquid-to-supercritical heavy water. Spectral diffusion and libration-induced hydrogen-bond dynamics
,”
Phys. Chem. Chem. Phys.
17
,
29776
29785
(
2015
).
23.
J. A.
Fournier
,
W.
Carpenter
,
L.
De Marco
, and
A.
Tokmakoff
, “
Interplay of ion–water and water–water interactions within the hydration shells of nitrate and carbonate directly probed with 2D IR spectroscopy
,”
J. Am. Chem. Soc.
138
,
9634
9645
(
2016
).
24.
Ł.
Szyc
,
M.
Yang
, and
T.
Elsaesser
, “
Ultrafast energy exchange via water–phosphate interactions in hydrated DNA
,”
J. Phys. Chem. B
114
,
7951
7957
(
2010
).
25.
R.
Costard
,
I. A.
Heisler
, and
T.
Elsaesser
, “
Structural dynamics of hydrated phospholipid surfaces probed by ultrafast 2D spectroscopy of phosphate vibrations
,”
J. Phys. Chem. Lett.
5
,
506
511
(
2014
).
26.
R.
Costard
,
T.
Tyborski
,
B. P.
Fingerhut
, and
T.
Elsaesser
, “
Ultrafast phosphate hydration dynamics in bulk H2O
,”
J. Chem. Phys.
142
,
212406
(
2015
).
27.
R.
Costard
,
T.
Tyborski
, and
B. P.
Fingerhut
, “
Anharmonicities and coherent vibrational dynamics of phosphate ions in bulk H2O
,”
Phys. Chem. Chem. Phys.
17
,
29906
29917
(
2015
).
28.
T.
Elsaesser
,
J.
Schauss
,
A.
Kundu
, and
B. P.
Fingerhut
, “
Phosphate vibrations probe electric fields in hydrated biomolecules: Spectroscopy, dynamics, and interactions
,”
J. Phys. Chem. B
125
,
3899
3908
(
2021
).
29.
J.
Schauss
,
F.
Dahms
,
B. P.
Fingerhut
, and
T.
Elsaesser
, “
Phosphate-magnesium ion interactions in water probed by ultrafast two-dimensional infrared spectroscopy
,”
J. Phys. Chem. Lett.
10
,
238
243
(
2019
).
30.
J.
Schauss
,
A.
Kundu
,
B. P.
Fingerhut
, and
T.
Elsaesser
, “
Contact ion pairs of phosphate groups in water: Two-dimensional infrared spectroscopy of dimethyl phosphate and ab initio simulations
,”
J. Phys. Chem. Lett.
10
,
6281
6286
(
2019
).
31.
J.
Schauss
,
A.
Kundu
,
B. P.
Fingerhut
, and
T.
Elsaesser
, “
Magnesium contact ions stabilize the tertiary structure of transfer RNA: Electrostatics mapped by two-dimensional infrared spectra and theoretical simulations
,”
J. Phys. Chem. B
125
,
740
747
(
2021
).
32.
T.
Siebert
,
B.
Guchhait
,
Y.
Liu
,
R.
Costard
, and
T.
Elsaesser
, “
Anharmonic backbone vibrations in ultrafast processes at the DNA–water interface
,”
J. Phys. Chem. B
119
,
9670
9677
(
2015
).
33.
T. D.
Kühne
et al, “
CP2K: An electronic structure and molecular dynamics software package - quickstep: Efficient and accurate electronic structure calculations
,”
J. Chem. Phys.
152
,
194103
(
2020
).
35.
Y.
Zhao
and
D. G.
Truhlar
, “
Design of density functionals that are broadly accurate for thermochemistry, thermochemical kinetics, and nonbonded interactions
,”
J. Phys. Chem. A
109
,
5656
5667
(
2005
).
36.
L.
Goerigk
,
A.
Hansen
,
C.
Bauer
,
S.
Ehrlich
,
A.
Najibi
, and
S.
Grimme
, “
A look at the density functional theory zoo with the advanced GMTKN55 database for general main group thermochemistry, kinetics and noncovalent interactions
,”
Phys. Chem. Chem. Phys.
19
,
32184
32215
(
2017
).
37.
W.
Humphrey
,
A.
Dalke
, and
K.
Schulten
, “
VMD: Visual molecular dynamics
,”
J. Mol. Graphics
14
,
33
38
(
1996
).
38.
L.-P.
Wang
,
T. J.
Martinez
, and
V. S.
Pande
, “
Building force fields: An automatic, systematic, and reproducible approach
,”
J. Phys. Chem. Lett.
5
,
1885
1891
(
2014
).
39.
E. T. J.
Nibbering
and
T.
Elsaesser
, “
Ultrafast vibrational dynamics of hydrogen bonds in the condensed phase
,”
Chem. Rev.
104
,
1887
1914
(
2004
).
40.
K.
Kwak
,
D. E.
Rosenfeld
, and
M. D.
Fayer
, “
Taking apart the two-dimensional infrared vibrational echo spectra: More information and elimination of distortions
,”
J. Chem. Phys.
128
,
204505
(
2008
).
41.
I.
Persson
,
M.
Trublet
, and
W.
Klysubun
, “
Structure determination of phosphoric acid and phosphate ions in aqueous solution using EXAFS spectroscopy and large angle X-ray scattering
,”
J. Phys. Chem. A
122
,
7413
7420
(
2018
).
42.
L. B.
Skinner
,
C.
Huang
,
D.
Schlesinger
,
L. G. M.
Pettersson
,
A.
Nilsson
, and
C. J.
Benmore
, “
Benchmark oxygen-oxygen pair-distribution function of ambient water from x-ray diffraction measurements with a wide Q-range
,”
J. Chem. Phys.
138
,
074506
(
2013
).
43.
A.
Kundu
,
S. I.
Mamatkulov
,
F. N.
Brünig
,
D. J.
Bonthuis
,
R. R.
Netz
,
T.
Elsaesser
, and
B. P.
Fingerhut
, “
Short-range cooperative slow-down of water solvation dynamics around SO42−–Mg2+ ion pairs
,”
ACS Phys. Chem. Au
2
,
506
514
(
2022
).
44.
D.
Laage
and
J. T.
Hynes
, “
A molecular jump mechanism of water reorientation
,”
Science
311
,
832
835
(
2006
).