3.2. GIAO-DFT Calculations of NMR Chemical Shifts Based on MP2 Geometries
Table 2 presents the computed average chemical shifts for modeled water clusters (assuming fast exchange) at the GIAO/SMD/MPW1PW91/6-311+G-(2d,p)//MP2/CBS-e level of theory. The MP2/CBS-e geometries were obtained from references [
15,
16].
Figure 2 shows the relationship between the number of water molecules in clusters (H
2O)
n (
n = 2–5) and their corresponding chemical shift (δ). To model water clusters, the geometries were optimized using various levels of theory, including MP2/aug-cc-pVTZ and MP2/6-31G*. As the number of water molecules increases, the chemical shifts (in ppm) also increase due to enhanced hydrogen bonding in small clusters such as (H
2O)
2, (H
2O)
3, (H
2O)
4, and (H
2O)
5. Both the MP2/aug-cc-pVTZ and MP2/6-31G* methods predict similar chemical shifts, although MP2/aug-cc-pVTZ yields slightly higher values, suggesting a more accurate description of hydrogen bonding in these systems.
The chemical shift continues to rise exponentially with the inclusion of pentamers (H
2O)
5. This supports the analysis that
n = 5 marks a critical point in water clusters’ structural and energetic cooperativity [
4,
7]. Both computational methods exhibit a consistent exponential increase up to this point, rather than a plateau, indicating that cooperative hydrogen bonding continues to strengthen through the fifth molecule.
The relationship between chemical shifts (ppm) and the number of hydrogen bonds in (H2O)n clusters (n = 2–4) calculated for distilled water using the MP2/aug-cc-pVTZ method demonstrates an exponential trend. The exponential fit yields a residual sum of squares (RSS) value of 0.0617, compared to 0.1207 for the linear fit and 0.2655 for the logarithmic fit. A similar exponential dependence is confirmed using the MP2/6-31G* method for (H2O)n (n = 2–4), where the exponential model achieves a lower RSS value of 0.1656, compared to 0.2579 for the linear model and 0.4687 for the logarithmic model.
These results consistently support the exponential model as the best fit for chemical shift data from the dimer to pentamer in water. The coordinates obtained from the MP2/aug-cc-pVTZ and MP2/6-31G* calculations show a strong positive correlation for (H2O)n, n = 2–4, with a Pearson correlation coefficient of r = 0.997, respectively, demonstrating consistency between the high-level computational approaches.
For (H2O)n, n = 2–5, with RSS, the exponential model is most prominent. For MP2/aug-cc-pVTZ, the value for exponential dependence is 0.0839; for linear, it is 0.1247; and for logarithmic, it is 0.4277. For MP2/6-31G*, the value for exponential dependence is 0.1739; for linear, it is 0.2829; and for logarithmic, it is 0.7626. The coordinates obtained from MP2/aug-cc-pVTZ and MP2/6-31G* calculations show a strong positive correlation for (H2O)n (n = 2–4), with a Pearson correlation coefficient of r = 0.997, highlighting the consistency between these high-level computational approaches.
Experimentally observed chemical shifts (δ, ppm) in hydrogen-rich water (HRW) show a clear dependence on the number of hydrogen bonds, particularly in small symmetrical clusters such as trimers and tetramers. This combined interpretation of DFT calculations and NMR measurements confirms that the experimentally observed chemical shifts in small water clusters arise from the cooperative effect of hydrogen bonding, wherein water structures exhibit an exponential stabilization trend to 10 water molecules.
The formation of primary hydrogen bonds in dimers leads to significant changes in electronic structure, while trimers begin to form cooperative hydrogen bond networks. Tetramers emerge as stable configurations in which the first signs of exponential hydrogen bond stabilization become evident. Overall, the exponential trend highlights that the stability of the hydrogen bonding network increases nonlinearly, reaching a saturation point at n = 4 water molecules.
In this study, we focused on the most energetically favorable and frequently reported structures. These low-energy configurations are also the most likely to occur under experimental conditions relevant to our NMR and FT-IR measurements. In
Table 3, we present the global minima and dominant low-energy structures for clusters up to
n = 5, optimized at the MP2 level. Although alternative isomers—such as star-shaped or figure-eight structures—are theoretically possible, they are significantly higher in energy and less relevant to the cooperative hydrogen-bonding effects under investigation.
The computed average (fast exchange) chemical shifts in modeled water clusters were analyzed using the RSS method at the GIAO/SMD/MPW1PW91/6-311+G(2d,p)//MP2/CBS-e level of theory, based on MP2/CBS-e geometries (
Table 3).
Table 4 presents the results achieved by fitting various mathematical models—linear, logarithmic, and exponential—for the relationship between the number of water molecules in clusters, from (H
2O)
2 to (H
2O)
n (
n = 4–7), and the corresponding NMR chemical shifts (δ, ppm), based on the data in
Table 3. These fits allow us to identify the most appropriate trend describing how cooperative hydrogen bonding within small clusters influences the observed chemical shifts. Theoretical analysis based on the data from
Table 4 shows that symmetric cyclic trimers and tetramers exhibit an exponential dependence of chemical shifts (δ, ppm) on the number of hydrogen bonds. This trend results from the cooperative effect in symmetric structures, where nearly equivalent hydrogen bonds lead to the uniform strengthening of the hydrogen bond network. The theoretical model for the (H
2O)
2–(H
2O)
4 range in
Table 4 demonstrates the best fit with an exponential trend, as indicated by the lowest residual sum of squares (RSS) value of 0.0005. For the (H
2O)
2–(H
2O)
5 range, the linear model yields an RSS of 0.0521. These results align closely with MP2/aug-cc-pVTZ and MP2/6-31G* calculations, which produce RSS values of 0.00153 and 0.0000347, respectively, for the exponential model. The notably lower RSS, compared to MP2/aug-cc-pVTZ, underscores the strong exponential dependence observed in the experimental data.
Although MP2/6-31G* achieves an even smaller RSS, the consistency across all methods and data sources confirms that an exponential stabilization model best describes cooperative hydrogen bonding in small symmetric clusters. However, this exponential trend diminishes when extending the range to (H2O)n, n = 2–28. In such larger clusters, logarithmic and linear models provide RSS values of 1.692 and 5.510, respectively, indicating that exponential behavior is primarily characteristic of smaller, symmetric systems.
A comparative analysis of MP2/aug-cc-pVTZ and MP2/6-31G* for modeling chemical shifts in small water clusters reveals that MP2/aug-cc-pVTZ consistently provides higher accuracy and reliability, particularly in systems with three or more hydrogen bonds. At
n = 3, where stable, symmetric configurations begin to form, the difference in δ (ppm) between the two methods is minimal. However, for
n = 4–5, MP2/6-31G* systematically overestimates the chemical shift, leading to a growing deviation and a higher mean absolute error (MAE) of 0.460 ppm, compared to 0.296 ppm with MP2/aug-cc-pVTZ. This trend is visualized in
Figure 3, which plots the difference in chemical shifts as a function of the number of hydrogen bonds.
Figure 2 presents the GIAO-DFT-calculated chemical shift values (δ, ppm) a function of the cluster size (
n = 2–5), with n plotted along the X-axis and δ along the Y-axis.
Table 2 lists the corresponding δ values and the corresponding number of hydrogen bonds for each cluster. An exponential increase in δ is observed for
n = 2–4, with a trend toward saturation at
n = 5, indicating nonlinear and cooperative interactions within the hydrogen bond network.
Table 4 compares different fitting models. The exponential model provides an exponential trend for
n = 2–4. Linear and logarithmic approximations applied to larger clusters and hydrogen bond counts showed lower accuracy.
This behavior is consistent with previous studies, which showed that larger and more flexible basis sets, such as MP2/aug-cc-pVTZ, provided improved accuracy in molecular electronic properties compared to smaller sets, such as MP2/6-31G* [
30].
We calculated the difference in calculated chemical shifts (δ, ppm) between MP2/aug-cc-pVTZ and MP2/6-31G* as a function of the number of hydrogen bonds in small water clusters. A minimum difference occurs at n = 3 hydrogen bonds, corresponding to the formation of symmetric cluster structures. Beyond this point (n = 4–5), the MP2/6-31G* level increasingly overestimates the chemical shift compared to MP2/aug-cc-pVTZ, resulting in larger deviations.
In 1996, Luzar and Chandler demonstrated that the dynamics of hydrogen bonds in liquid water are governed by cooperative effects, leading to logarithmic or power-law dependencies in the formation and breakage of the hydrogen bond network. Our computed chemical shifts for water clusters exhibit a similar logarithmic trend, with clusters ranging from 8 to 24 molecules, indicating the onset of dynamic equilibrium within the hydrogen bond network. These results confirm that, after the initial rapid stabilization phase observed in smaller clusters, the further growth of the system results in progressively minor changes, characteristic of a well-established cooperative hydrogen-bonded network [
31]. The cooperative effects lead to the water clusters having a longer lifetime [
32].
In previous studies, the cooperative hydrogen bonding behavior of small water clusters was analyzed using quantum chemical methods based on density functional theory (DFT) [
5,
7]. These analyses revealed a clear nonlinear trend in the energetic and structural properties of water clusters, particularly from the dimer to the pentamer, with a distinct saturation effect emerging beyond the fifth molecule. For instance, it has been demonstrated that both the free energy and internal energy per water molecule decrease rapidly up to
n = 5, after which they stabilize, indicating a threshold in hydrogen bond cooperativity [
5]. Similarly, a pronounced nonlinear increase in hydrogen bond interactions and NMR chemical shifts was observed from the trimer to the pentamer, transitioning toward liquid-like behavior at larger cluster sizes [
7]. In the present study, we introduced an empirical exponential model to qualitatively describe the variation in chemical shifts resulting from cooperative hydrogen bonding:
where δ
o is the baseline chemical shift, A is the change amplitude due to hydrogen bonding, and k represents the exponential growth rate.
The formula was derived using nonlinear regression on δ values computed at the GIAO/MPW1PW91/6-311+G(2d,p)//MP2/CGS-e level of theory. The resulting model successfully captures the saturation trend of cooperative hydrogen bonding in (H2O)n clusters for n = 2–5 and provides a superior statistical fit, reflected by a lower residual sum of squares (RSS) compared to linear and logarithmic alternatives. This model fits both experimental and theoretical data with high accuracy, reflecting the rapid increase in, and saturation of, δ values up to n = 4. As such, this formulation offers a novel yet physically grounded framework for describing cooperative behavior in hydrogen-bonded systems, building upon the conceptual foundation laid by earlier computational studies.
Figure 4 shows the exponential model of chemical shifts and hydrogen bonds from the formula, confirming cooperative hydrogen bond enhancement up to the water pentamer.
Table 5 summarizes the correlation between the number of hydrogen bonds in small water clusters and their corresponding NMR chemical shifts and compares the model predictions with the calculated data.
The exponential formulas are as follows:
Theoretical model: δ (n) = 1.94 + 3.20 (1 − e−0.605n);
Model-MP2/6-31G*: δ (n) = 1.89 + 3.10 (1 − e−0.678n);
Model-MP2/aug-cc-pVTZ: δ (n) = 1.62 + 3.50 (1 − e−0.760n).
A comparative analysis between the theoretical and MP2/6-31G* experimental models yields a coefficient of determination R2 = 0.9987, indicating excellent agreement. In contrast, the MP2/aug-cc-pVTZ model yields a lower R2 = 0.7685, indicating moderate alignment with theoretical predictions. The difference in R2 values highlights the superior accuracy of the CBS-extrapolated method, which better captures long-range cooperative effects in hydrogen bonding. Despite variations in the absolute δ values, both experimental models follow the exponential trends observed for hydrogen bonds in cyclic water clusters with the formula (H2O)n, n = 2–4, affirming the physical consistency of the hydrogen bond saturation effect. These findings demonstrate that the proposed exponential model is mathematically robust, physically meaningful, and broadly applicable for the characterization of chemical shifts in hydrogen-bonded water clusters.
The corrected MP2/aug-cc-pVTZ model is more accurate, as its predicted chemical shifts are close to the experimentally measured values for small water clusters, with the lowest mean absolute error (below 0.5 ppm). Unlike the other two models, it uses a high-quality aug-cc-pVTZ basis set that better captures the electronic structure and cooperative hydrogen bonding interactions. Moreover, comparison with experimental NMR data shows that the MP2/aug-cc-pVTZ model provides quantitative and qualitative agreement, outperforming the purely theoretical model and the MP2/6-31G* approximation.
In this section, we present a quantum chemical analysis of small water clusters (H2O)n (n = 1–5), combining second-order Møller–Plesset perturbation theory (MP2) with density functional theory (DFT) to evaluate structural stability and NMR chemical shifts (δ). The geometries of the water clusters were optimized using the MP2/6-31G* and MP2/aug-cc-pVTZ levels of theory, followed by harmonic frequency analysis to ensure vibrational stability. The results show that all clusters are vibrationally stable, as imaginary frequencies were not detected in any of the structures. This confirms that the optimized geometries correspond to the true local minima on the potential surface.
The hydrogen bond network formed in each cluster is characterized by intermolecular O···H distances between 1.7 and 2.0 Å and O···O separations of approximately 2.7 Å. These distances indicate strong hydrogen bonding and are consistent with the expected chemical shift trends. These geometrical parameters align with the increasing chemical shift trends observed in the GIAO-DFT calculations. Subsequent DFT/GIAO calculations of NMR chemical shifts were performed at the B3LYP/6-31G* and MPW1PW91/6-311+G(2d,p) levels, applied to the MP2-optimized geometries. This combined approach is considered cost-effective and reliable in accurately predicting magnetic shielding in small molecular systems [
15].
3.3. Fourier-Transform Infrared (FT-IR) Spectroscopy
Fourier-transform infrared (FT-IR) was performed on distilled water (
Figure 5).
The FT-IR spectrum reveals characteristic absorption bands corresponding to the vibrational dynamics of O–H bonds and the underlying hydrogen-bonding network. In liquid water, the broad hydroxyl-stretching region observed between 3201 and 3504 cm
−1 reflects the dynamic and extensive hydrogen bonding among water molecules. DFT calculations for isolated small water clusters—including dimers, trimers, and tetramers—predict the OH stretching frequency as the number of hydrogen bonds increases. FT-IR studies have shown that free OH groups not involved in hydrogen bonding exhibit stretching vibrations near 3700 cm
−1 [
33]. This spectral feature is characteristic of isolated water molecules (monomers) or environments where hydrogen bonding is partially disrupted. The red shift in the OH stretching band observed in small hydrogen-bonded clusters, such as dimers through tetramers, has been well documented in both experimental and theoretical studies. High-level ab initio calculations (e.g., MP2 and CCSD(T)) confirm that the OH stretching frequency shifts from ~3700 cm
−1 in free water molecules to the 3200–3500 cm
−1 range in small clusters, where cooperative hydrogen bonding stabilizes the network [
28]. This trend is further supported by FT-IR measurements at hydrophobic interfaces, where free OH groups are detected around 3700 cm
−1, while hydrogen-bonded OH groups exhibit lower stretching frequencies due to the weakening of the OH bond due to hydrogen bonding [
30].
The experimentally observed broad, asymmetric band in the 3201–3504 cm
−1 region (
Figure 4) reflects the strengthening and cooperative nature of the hydrogen-bonding network.
This spectral profile aligns with DFT predictions for OH stretching frequencies in small clusters, where a distribution of hydrogen bond strengths results in a wide range of vibrational shifts [
33]. The combined interpretation of DFT calculations and FT-IR measurements confirms that this broad band arises from water clusters of varying sizes and hydrogen-bonding configurations [
34]. This approach, which links theoretical and experimental data, has been successfully applied in previous studies on hydrogen-bonding dynamics in bulk water and water at interfaces, with the role of free and hydrogen-bonded OH groups being analyzed [
35,
36,
37].
In addition to the OH stretching region, the absorption band around 1650 cm
−1, attributed to the H–O–H bending (scissoring) mode, is also indicative of hydrogen bonding. In our experimental data, this band appears at 1658 cm
−1 [
38]. DFT calculations predict that as water forms larger clusters, the scissor mode undergoes subtle frequency shifts and increases in intensity. This is consistent with the increased coupling between bending and stretching vibrations in hydrogen-bonded systems.
Finally, the combination band observed at 2130 cm
−1, commonly seen in liquid water, arises from coupled bending and librational motions within the hydrogen bond network [
39].
Furthermore, the broad OH stretching band observed in the FT-IR spectrum spanning approximately 3200 to 3500 cm
−1 is indicative of threshold behavior at the size level of the water pentamer (H
2O)
5. The water pentamer is considered a critical threshold in the development of hydrogen-bonded aggregates, as it is the smallest cluster capable of forming a closed, cyclic network of five hydrogen bonds. This configuration enables the emergence of cooperative effects, vibrational redshifts, and geometries that closely resemble those found in bulk water [
40,
41].
Significant experimental work by Saykally and co-workers has shown, through high-resolution FT-IR and microwave spectroscopy, that, starting from the pentamer, water clusters exhibit broadened and redshifted O–H stretching bands akin to those in the liquid phase [
42]. These features are not observed in smaller clusters and point to the pentamer as the smallest structure with liquid-like spectral characteristics.
Recent findings further reinforce this view. Xie et al. [
4] reported that five water molecules are sufficient to completely dissociate a single HCl molecule, providing direct evidence that the pentamer exhibits incipient solvent behavior, including dielectric screening and ion stabilization. These results build on decades of computational studies that have identified the pentamer as the onset of three-dimensional hydrogen bond networks with enhanced cooperativity [
4,
40].
Thus, the distinct broadening and asymmetry of the OH stretching band observed in our FTIR measurements may be seen as spectroscopic confirmation of this structural threshold. The appearance of vibrational modes in the 3500–3600 cm
−1 range, which are absent from smaller clusters, further supports this assignment [
43].
3.4. Electrochemical Impedance Spectroscopy (EIS) Results
Complex electrical impedance spectroscopy is a powerful method of investigating the ionic conductivity of condensed matter layers.
Figure 6a shows the real (Z′) and imaginary (Z″) parts of the complex electrical impedance in the Nyquist plot of a thin layer of distilled water. It can be seen that the parallel arrangement of bulk resistance and bulk capacitance can explain the typical contour of a semicircle in the frequency range studied. In contrast with the Nyquist diagram, the Bode plot (
Figure 6b) presents the impedance modulus and phase dependencies on the frequency. This diagram allows the accurate localization of the relaxation processes and their relationship with the cooperative dynamics of the hydrogen network areas.
Figure 7 illustrates the frequency spectra of the real and imaginary parts of dielectric permittivity for a thin distilled water layer.
In this study, the frequency dependence of the complex dielectric permittivity of distilled water was analyzed, revealing a decrease in both the real (ε1) and imaginary (ε2) components of the dielectric permittivity with increasing frequency. High dielectric permittivity values are observed in the low-frequency range (0.1–1 Hz), indicating the occurrence of ionic conductivity and hydrogen bond interactions. At a frequency of 10 kHz, both components reach low values, suggesting a weakening in dipole relaxation. The results confirm that, although distilled water is chemically pure, it exhibits structural dynamics associated with hydrogen bonding and ionic effects.
The results with RSS in the 0.1–1 Hz range show an exponential trend. There is cooperative hydrogen bonding and cluster formation, where the exponential trend is connected with relaxation processes associated with the hydrogen bond network. The results with RSS in the 1 Hz–10 kHz range show a logarithmic trend.
Complex electrical impedance spectroscopy (EIS) provides valuable information on distilled water’s ionic conductivity and polarization dynamics. The Nyquist plot (
Figure 6a) reveals a characteristic semicircle, indicating that the water layer can be further modeled as a parallel combination of bulk resistance and capacitance.
The frequency dependence of the real (ε
1) and imaginary (ε
2) parts of the dielectric permittivity (
Figure 7) exhibits a dispersion effect at the frequency range (0.1–1 Hz), where both real (ε
1) and imaginary (ε
2) reach exponential trends. This behavior reflects the contributions of cooperative dipole orientation processes and ionic conductivity effects. The two methods are closely related to the dynamic formation and reorganization of hydrogen bond networks.
For the frequency range (1 Hz–10 kHz), real (ε1) and imaginary (ε2), the sharps decrease and indicate transition regimes dominated by local dipole fluctuations and a reduced ability of larger cooperative structures to follow the applied field. The high dielectric permittivity at low frequencies is particularly notable, as it correlates with larger hydrogen-bonded clusters and extended cooperative networks. DFT calculations show that small cyclic clusters (especially trimers and tetramers) exhibit enhanced cooperative hydrogen bonding, which can be expected to contribute to the observed dielectric response. The exponential increase in hydrogen bonding strength and dipole alignment observed in DFT-calculated water clusters correlates with the enhanced dielectric response at low frequencies in the EIS measurements. The high values for real (ε1) and imaginary (ε2) permittivity below 1 Hz suggest dynamic reorientation and cooperative dipole behavior, which is structurally supported by DFT models of (H2O)2 and (H2O)5 clusters. Thus, the molecular-level cooperative effects are reflected in the macroscopic dielectric permittivity of bulk water.
The exponential trend of cooperative enhancement of hydrogen bonding in (H2O)n, n = 2–4, established through NMR and DFT, is also reflected in electrochemical impedance spectroscopy (EIS) within the frequency range of 0.1–1 Hz, where an exponential decrease in dielectric permittivity is observed, associated with the dynamics of the hydrogen bond network. This analysis is further supported by FT-IR data, where the broad OH stretching band (3200–3550 cm−1) and the bending mode at 1658 cm−1 provide evidence of a heterogeneous hydrogen bond network containing both weak and strong hydrogen bonds. The low-frequency dielectric response (high ε1 and ε1) is thus a macroscopic reflection of this structural heterogeneity, capturing the combined polarization response of various small clusters in the dynamic hydrogen bond network.
This study demonstrates that the dielectric properties of distilled water arise from the collective dynamics of small water clusters stabilized by a cooperative hydrogen bond network. By applying Bode plots, which present both the impedance magnitude and phase angle as a function of the frequency, it was possible to localize relaxation processes with frequency resolution. Bode analysis revealed an exponential decrease in impedance |Z| and a characteristic phase transition within the 0.1–10 Hz range, typical of the coordinated dipolar relaxation associated with the hydrogen bond network.
The experimentally observed exponential trend corresponds to DFT models of chemical shifts and the number of hydrogen bonds in water clusters. The exponential models are found in the range n = 2–4. This reflects phase-intensifying cooperativity, while at n = 5, a structural threshold is reached, which is the water pentamer (H2O)5. At this point, the hydrogen bond network acquires three-dimensional stability and begins to exhibit properties characteristic of bulk liquid water.
To conclude, the combination of EIS, FT-IR, and DFT data reveals that the dielectric properties of distilled water cannot be attributed to isolated water molecules. Instead, they emerge from the collective polarization and reorientation dynamics of small, cooperative water clusters, whose structural and vibrational properties have been independently confirmed using DFT and FT-IR. These findings are further supported by recent EIS measurements of distilled water at temperatures near freezing point (−0.1 °C and +0.1 °C), which revealed the formation of ordered ice I
h-type hydrogen-bonded clusters. These results show that cooperative dipolar polarization persists even in quasi-solid water states, highlighting the continuity of dielectric response [
13].