Next Article in Journal
Comprehensive Analysis of Lutein and Loroxanthin in Scenedesmus obliquus: From Quantification to Isolation
Next Article in Special Issue
Towards the “Eldorado” of pKa Determination: A Reliable and Rapid DFT Model
Previous Article in Journal
Structure, Absolute Configuration, Antiproliferative and Phytotoxic Activities of Icetexane and Abietane Diterpenoids from Salvia carranzae and Chemotaxonomic Implications
Previous Article in Special Issue
Density Functional Theory Studies on the Chemical Reactivity of Allyl Mercaptan and Its Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantum-Chemistry Study of the Hydrolysis Reaction Profile in Borate Networks: A Benchmark

by
Francesco Muniz-Miranda
,
Leonardo Occhi
,
Francesco Fontanive
,
Maria Cristina Menziani
and
Alfonso Pedone
*
Dipartimento di Scienze Chimiche e Geologiche (DSCG), Università degli Studi di Modena e Reggio-Emilia (UNIMORE), Via G. Campi 103, 41125 Modena, Italy
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(6), 1227; https://doi.org/10.3390/molecules29061227
Submission received: 19 January 2024 / Revised: 5 March 2024 / Accepted: 8 March 2024 / Published: 9 March 2024
(This article belongs to the Special Issue Multiconfigurational and DFT Methods Applied to Chemical Systems)

Abstract

:
This investigation involved an ab initio and Density Functional Theory (DFT) analysis of the hydrolysis mechanism and energetics in a borate network. The focus was on understanding how water molecules interact with and disrupt the borate network, an area where the experimental data are scarce and unreliable. The modeled system consisted of two boron atoms, bridging oxygen atoms, and varying numbers of water molecules. This setup allows for an exploration of hydrolysis under different environmental conditions, including the presence of OH or H+ ions to simulate basic or acidic environments, respectively. Our investigation utilized both ab initio calculations at the MP2 and CCSD(T) levels and DFT with a range of exchange–correlation functionals. The findings indicate that the borate network is significantly more susceptible to hydrolysis in a basic environment, with respect to an acidic or to a neutral pH setting. The inclusion of explicit water molecules in the calculations can significantly affect the results, depending on the nature of the transition state. In fact, some transition states exhibited closed-ring configurations involving water and the boron–oxygen–boron network; in these cases, there were indeed more water molecules corresponding to lower energy barriers for the reaction, suggesting a crucial role of water in stabilizing the transition states. This study provides valuable insights into the hydrolysis process of borate networks, offering a detailed comparison between different computational approaches. The results demonstrate that the functionals B3LYP, PBE0, and wB97Xd closely approximated the reference MP2 and CCSD(T) calculated reaction pathways, both qualitatively in terms of the mechanism, and quantitatively in terms of the differences in the reaction barriers within the 0.1–0.2 eV interval for the most plausible reaction pathways. In addition, CAM-B3LYP also yielded acceptable results in all cases except for the most complicated pathway. These findings are useful for guiding further computational studies, including those employing machine learning approaches, and experimental investigations requiring accurate reference data for hydrolysis reactions in borate networks.

1. Introduction

Borate networks, which are particularly prominent in materials such as boron–silica glasses, are crucial in various applications, ranging from consumer products like Pyrex to advanced industrial materials. These networks’ significance stems not only from their widespread usage but also from their unique chemical and physical properties, making them subjects of extensive research. In particular, borate and borosilicate glasses have been shown to be able to promote the regeneration of enamel, bone, and nervous tissues when in contact with physiological fluids [1,2,3]. To enhance these functions, the dissolution rate of the glass in a water solution must match that of growth of the replaced human tissues. Unfortunately, the development and large-scale use of these glasses is hindered by a lack of knowledge of their chemical durability in physiologic conditions. This knowledge gap stems from the fact that a vast majority of investigative efforts in the field of glass corrosion has focused on silicate glasses and their derivatives, but the principles of silicate glass corrosion (e.g., composition–reactivity correlations) generally do not apply to borate glasses [4,5].
The hydrolysis of borate networks is a fundamental aspect of their chemistry, and knowledge of this process is essential for applications and a theoretical understanding. However, experimental studies in this area face significant challenges due to the complexity of the reactions and the difficulty in accurately measuring the reaction pathways and energetics. G. Werding and W. Schreyer [6] provided valuable insights into the behavior of borosilicates and selected borates under various conditions. The interaction between boric acid molecules/metaborate ions and water molecules was studied by Zhu et al. [7].
In addition to the challenges in studying borate hydrolysis, there was significant research in related areas of boron chemistry, which offers insights into the broader context of borate interactions. For example, studies on boronic acids and their derivatives provided an essential understanding of boron-based molecular recognition and sensor applications. The versatility of boronic acids in sensing and imaging applications was demonstrated by Nishiyabu et al. [8], Dai et al. [9], and Sun et al. [10]. Chan et al. [11] and Lippert et al. [12] have also contributed to the understanding of boronic acid chemistry, particularly in the context of bio-orthogonal reactions and fluorescence imaging.
Furthermore, the saccharide recognition capabilities of boronic acids, as explored by James et al. [13,14], provide a framework for understanding the interactions between borate networks and other molecular systems. The work of Jin et al. [15] on carbohydrate recognition by boronolectins and lectins illustrated the potential of borate and boronic acid systems in biological contexts.
Thus, while the experimental study of borate hydrolysis remains a challenging field, the broader research on boron chemistry, including boronic acids and their derivatives, has provided valuable insights. These studies not only enhanced our theoretical understanding but also had significant implications for various practical applications, from sensor development to biological recognition. As the field continues to evolve, the integration of experimental and computational approaches will undoubtedly yield a more comprehensive understanding of borate chemistry.
Indeed, the hydrolysis mechanism of borate networks is crucial for understanding the chemistry of materials like borosilicate glasses. Zhou et al. [16] delved into the specifics of pentaborate anion hydrolysis, shedding light on the intricate dynamics of borate networks in hydrolytic conditions, and suggested that the formation of ring structures plays a role, with hydrolization occurring at the bridge atom.
An in-depth look into the molecular events at the borosilicate glass–water interface, which is fundamental for understanding the interactions of borate networks with water, was achieved by Jabraoui et al. [17] using non-hybrid exchange–correlation functionals; they showed that hydrogen bonds between water and the oxygen atoms of a glassy borosilicate network play a role in the binding of the water to the glass.
Kagan et al. [18] contributed to the understanding of the dissolution processes in silicate networks, which is particularly relevant for the study of materials similar to borate networks, and found activation energies within the 15–40 kcal/mol with classical potentials.
To expand the application of borate networks, we need to develop a rigorous atomic-level fundamental understanding of the mechanisms and kinetics of the dissolution of these systems in water. Particular emphasis and attention should be given to explore the (i) atomistic origins of chemical and structural drivers that control the dissolution kinetics of borate networks and (ii) underlying mechanisms that govern glass dissolution.
Although a few systematic and important benchmark studies for hydrolysis reactions exist, these have been carried for organic reactions and their extension to borate networks is not obvious [19,20].
Therefore, in this study, we focused on the process of hydrolysis of a boronic dimer in three different pH conditions, namely neutral, acidic, and basic, with a certain number of explicitly treated water molecules, to extract the reaction pathways and energies. This is a first benchmarking study of DFT functionals against the MP2 and CCSD(T) calculations needed to individuate a low-cost level of theory to investigate more complex and large systems as well as to produce quantum mechanical databases for the development of reactive force fields and/or Machine Learning Potentials and to feed kinetic Monte Carlo Models to study borate and borosilicate glass dissolution at larger space and time scales.

2. Results and Discussion

2.1. Models

To model the hydrolysis reaction, we have considered a small network formed by two B atoms, connected by a bridging oxygen atom as the reagent (see reagent in Figure 1). The overall starting coordination number of each boron atom is three, and the non-bridging oxygen atoms are saturated with hydrogens to complete their coordination and to produce a neutral compound. We carried out attempts to simulate it without saturating the compound with hydrogens but the excessive negative charge on oxygen atoms led to unrealistic proton transfer from the water molecules. The products of the hydrolysis reaction were, instead, two B(OH)n compounds (with n equal to either 3 or 4).
In addition to these compounds, we introduced explicit water molecules to model a pH 7 environment, and H3O+ and OH ions to model acidic and basic conditions. For the neutral, acidic, and basic hydrolysis pathways, we also considered systems with two additional water molecules, as they are enough to create a ring-like transition state that can alter the energy barriers.
All the possible reactions pathways involved at least one transition state with a tetra-coordinated boron atom.

2.2. Assessment of the Basis Set and Ab Initio Levels of Theory

We considered the stationary points of the reaction pathway connecting a B-O-B containing unit, namely the B2O5H4 compound, and two separated orthoboric acid molecules, in an environment without a single water molecule, as a reference to check the reliability of the basis set employed as well as the MP2 level of theory. In fact, this was the simplest model studied here. The geometry of the stationary point is illustrated in Figure 1 and the model itself is discussed in more detail in Section 2.3.
Figure 1. Scheme of reaction pathway for the hydrolysis mechanism at neutral pH with a single water molecule. ξ represents the reaction coordinate. The energies are not to scale; see Table 1 for a complete report on the energy values.
Figure 1. Scheme of reaction pathway for the hydrolysis mechanism at neutral pH with a single water molecule. ξ represents the reaction coordinate. The energies are not to scale; see Table 1 for a complete report on the energy values.
Molecules 29 01227 g001
The stationary point was first found at the MP2/6-311++G(d,p) level of theory combined with PCM.
We performed single-point CCSD(T) [21]/Apr-cc-pVTZ calculations on each of these geometries, combined with PCM, to check that the MP2-computed energy barriers were reliable in comparison with the CCSD(T) level of theory. This was performed as Gaussian 16 currently does not allow geometry optimization at the CCSD(T) level of theory.
The computed energy differences are reported in Table 1, and suggest that CCSD(T) single points reproduce the results of MP2 optimizations with errors < 0.01 eV for the transition state barrier, and just slightly larger for the energy of the products.
Table 1. Energies for the stationary points of the reaction pathway at neutral pH with 1 water molecule computed with MP2 and CCSD(T) methods and different basis sets. The energies have been rescaled to have the energy of the reagents put to zero. The energy of the TS coincides with Δ E 1 W p H = 7 .
Table 1. Energies for the stationary points of the reaction pathway at neutral pH with 1 water molecule computed with MP2 and CCSD(T) methods and different basis sets. The energies have been rescaled to have the energy of the reagents put to zero. The energy of the TS coincides with Δ E 1 W p H = 7 .
Level of TheoryStationary PointEnergy/eV
Optimized MP2/6-311++G(d,p) + PCMReagents0
TS2.030
Product −0.015
Single-point CCSD(T)/Apr-cc-pVTZ + PCMReagents0
TS2.038
Products−0.026
Optimized MP2/6-311++G(2d,2p) + PCMReagents0
TS2.026
Products−0.022
Optimized MP2/6-31+G(d) + PCMReagents0
TS2.042
Products−0.034
Optimized MP2/6-21G + PCMReagents0
TS2.373
Products−0.009
Optimized MP2/LanL2DZ + PCMReagents0
TS2.118
Products−0.011
Also, we checked the effect of the basis set, in particular, using the 6-21G, 6-31+G*, 6-311++G(d,p), and 6-311++G(2d,2p) basis sets and the combined basis-set-pseudopotential LanL2DZ [22], optimizing the stationary points at the MP2 + PCM level of theory, as also reported in Table 1.
The energy differences remained well below the 0.01 eV threshold for the 6-311++G(d,p) and 6-311++G(2d,2p) cases, and the B-Obridge distances of the transition state changed from 1.613 Å (tetrahedral B) and 1.414 to 1.616 Å (tetrahedral B) and 1.399 Å, respectively. Overall, expanding the basis set seemed to only marginally affect the energies and very negligibly affect the structure of the models we adopted here. On the contrary, using smaller basis sets led to increased differences in the energy barrier, which were particularly evident in the case of the 6-21G basis set (>0.1 eV).
Overall, we considered the 6-311++G** as a good compromise between accuracy, computational burden, and availability in the many possible quantum chemistry codes.

2.3. Neutral pH System

To study the hydrolysis mechanism in a neutral aqueous environment, we found and considered two possible different pathways, one including a single water molecule (1 W) and another one (3 W) exploiting the presence of three water molecules.
The structures associated with the pathway 1 W are reported in Figure 1 of Section 2.2, as we used this model (the simplest one) to also assess the validity of the ab initio and basis set levels of theory.
Following pathway 1 W, we found a transition state (TS) resulting from the nucleophilic attack of the O atom of water on a boron atom, and at the same time, the formation of a bond between a hydrogen atom of water and the bridging oxygen atom of the B-O-B network. This mechanism thus resulted in a very unstable TS (2.03 eV at the MP2 level of theory) constituted by a 4-term ring (B-Owater-Hwater-Obridge). Indeed, the Δ E 1 W p H = 7 energies, i.e., the difference between the computed energy of the transition state and the energy of the reagents (see Figure 1), were the highest that we found for any mechanism, confirming the high unlikely and unstable nature of this TS and, consequently, of this reaction pathway. The actual energies calculated with different functionals are reported in Table 2.
Instead, models using three water molecules gave rise to a TS with a 7-term ring (see Figure 2), resulting in much lower transition energies (0.478 eV at the MP2 level), as shown in Table 3.
It must be pointed out that while the reaction model with just one water molecule, the DFT methods yielded reaction barriers about 0.5 eV lower than the MP2 result, in the model using three water molecules, this difference was significantly quenched. Indeed, ωB97XD approximated the MP2 result up to 0.01 eV, and B3LYP and PBE0 showed differences of about 0.1 and 0.2 eV, respectively. Also, some functional might yield an energy for products higher than that of the reactants, about hundredths of an eV.

2.4. Acidic pH System

In presence of an H3O+ cation, to mimic acidic conditions, the energy landscape of the reaction changes significantly. In particular, in the neutral environment, we were able to locate a single transition state, involving both the protonation of the bridging oxygen and the hydroxylation of a B atom; in acidic conditions, we observed these two processes occurring in sequence, in two different stationary points.
Namely, we consistently observed the protonation of the Obridge atom in the transition state TS1, which was followed by an intermediate state INT where a water molecule bound to a boron atom, thus making the coordination of the latter change from trigonal to tetrahedral. This was followed by a second transition state that we indicated as TS2, where the bond between the bridging oxygen and the tetrahedrally coordinated boron atom was weakened and lengthened, until it eventually broke to give rise to the products.
It is important to point out that we found this mechanism occurring both in the model with just one H3O+ cation (named 1 W, see Figure 3), and in that with one H3O+ cation and two water molecules (named 3 W, see Figure 4), at all levels of theory.
The presence of three stationary points between the reagents and products (two transition states and one intermediate) in the case of the acidic mechanism, instead of just one (a single, concerted transition state) as in the case of the mechanism at neutral conditions, dramatically lowers the energy barrier for the 1 W model, as reported in Table 4.
Also, it must be highlighted that for the 1 W model, the first energy barrier, Δ E 1 W p H < 7 , defined as the energy difference between the reagents and the TS1, was significantly higher than the second one, Δ E 1 W p H < 7 , defined as the energy difference between the intermediate (INT) and the TS2. A graphical representation of both Δ E 1 W p H < 7 and Δ E 1 W p H < 7 is shown in Figure 3. This means that the rate-determining step in this case is the protonation of the bridging oxygen. The actual energies found at the various levels of theory for this model are reported in Table 4.
The situation is more complicated in the 3 W model due to the increased complexity of the system. Indeed, at the DFT level of theory, the specific choice of the functional can even yield a negative energy for either Δ E 3 W p H < 7 (M06HF functional) or Δ E 3 W p H < 7 (CAM-B3LYP, BLYP, HSE06, PBE, and M06 functionals); this occurs since the TS2 has a negative vibrational frequency, but it still had a lower energy with respect to the intermediate INT. A graphical representation of both Δ E 3 W p H < 7 and Δ E 3 W p H < 7 is shown in Figure 4, while the actual energies are reported in Table 5.
We investigated the reasons behind this behavior, and found them to be dependent on these functional giving a higher hydrogen bond stabilization energy, which significantly lowered the energy of the TS2.
This also explains why the 1 W model, which employs just a single H3O+ cation, did not exhibit this behavior, since in that case, the number of hydrogen bonds potentially skewing the energy budget was smaller.
In fact, using these functionals to perform single-point calculations on the stationary points found and optimized at the MP2 level produced a behavior that is largely correct but, as expected, they cannot recognize transition states or intermediates based on their vibrational frequencies. We want to underline here that B3LYP, ωB97XD, and PBE0 all gave energies and energy differences close to those found at the MP2 level, with differences of less than 0.1 eV.
Overall, the energy barriers at the MP2 level and using B3LYP, ωB97XD, and PBE0 functionals yielded similar results in the 1 W and 3 W models; this can be explained by the lack of a concerted transition state in acidic conditions that (energetically) benefits from having a larger amount of water molecules explicitly included into the model. Therefore, different from the neutral simulations, in this case, the 1 W model seemed to be reliable for obtaining accurate energy barriers. However, for the functionals that poorly reproduce H-bonds, the 3 W model could indeed lead to erratic results.

2.5. Basic pH System

We simulated alkaline conditions by using a single hydroxyl anion in our system, combined with 0 (model 1 W) or 2 (model 3 W) neutral water molecules. The hydroxyl anion attacks a boron atom, turning its coordination from trigonal to tetrahedral, giving rise to a single TS and, thus, to a rather simple energy pathway (see Figure 5 and Figure 6).
In particular, it must be noted that the reaction barrier became extremely small even for the 1 W model, on the order of about 0.1 eV at the MP2, B3LYP, PBE0, and ωB97XD levels of theory, as reported in Table 6. In fact, the reaction barrier was so small that we had issues optimizing the reagents, as they very easily evolved into the transition state, which was very close in energy. Also, the reaction evolved without the need for the protonation of the bridging oxygen, which is different from the mechanisms in the neutral and acidic conditions.
In the more complex 3 W model, we observed that the reaction barriers decreased (see Table 7) while maintaining a rather similar transition geometry for the B2O5H4 and OH fragments but with a ring-like disposition for the other two additional water molecules. This decrease in the reaction barriers occurred at all levels of theory.

3. Computational Details

All QM calculations of the molecule–metal complexes presented here were performed with the Gaussian 16 suite of programs [23]. To produce the reference reaction energy profiles, we resorted to the 2nd order Moeller–Plesset perturbation theory approach (MP2) [24] and CCSD(T) calculations using 6-311++G(d,p) and Apr-cc-pVTZ basis sets. The final choice of MP2 as the reference for the ab initio method is discussed in more detail in Section 2.2.
Regarding density functional theory calculations, a plethora of exchange–correlation functionals were used, namely hybrid B3LYP [25,26], PBE0 [27], M06 [28], M06HF [29], HSE06 [30], CAM-B3LYP [31], ωB97Xd [32], and GGA and meta-GGA like BLYP [26,33], PBE [34], TPSS [35]. B3LYP, PBE0, M06, and M06HF, which are general hybrid functionals with different percentages of Hartree–Fock exchange (for instance, 100% for M06HF, 25% for PBE0). HSE06, CAM-B3LYP, and ωB97Xd are instead range-separated hybrids, the first is based on PBE, the second on B3LYP, and the third one also includes dispersion effects. BLYP, PBE, and TPSS are “pure” functionals, and do not use any Hartree–Fock exchange.
All the DFT calculations adopted Pople’s all-electrons 6-311++G(d,p) basis set, in combination with the Grimme’s GD3BJ correction [36] or the simpler GD3 correction [37] of dispersion interactions (when the former is not available for a specific DFT functional). For the ωB97Xd functional, no further correction was adopted.
In all calculations, the linear-response polarizable continuum model (PCM) [38,39] was adopted to include bulk solvent effects.

Search and Assessment of Transition States

The discovery of the transition states is, notoriously, a rather subtle research approach, and for this reason, we carried it out with 3 possible approaches: (1) performing optimizations with the Berny algorithm (corresponding to the TS keyword in Gaussian 16); (2) using the QST2/QST3 approach, thus giving a guess of the starting reagents and final product (and, in case of QST3, of the transition state as well) and letting the Transit-Guided Quasi-Newton algorithm optimize the transition state; and (3) stretching what we supposed was the reaction coordinates, performing a relaxed scan on it, and looking for a maximum of the potential energy.
Notwithstanding the procedure we adopted to find the transition states, these latter have been validated in two ways: (a) by the presence of a single imaginary frequency in the Hessian and (b) by using the intrinsic reaction coordinate approach (IRC keyword in Gaussian) to check that following the forward and reverse reaction pathways lead to the products and the reactants, respectively.

4. Concluding Remarks

In this study, we performed a comprehensive ab initio and Density Functional Theory (DFT) investigation of the hydrolysis mechanism and energetics of a borate network. Our focus was on modeling a system comprising two boron (B) atoms and oxygen (O) bridging atoms, along with various water molecules, to examine the interaction dynamics where water attacks and breaks the network. Additionally, we incorporated either hydroxide (OH) or proton (H+) ions to simulate basic or acidic environments, respectively.
The motivation for this research stems from the challenges in obtaining reliable experimental data regarding the hydrolysis pathway and energies. These energetic insights are crucial for subsequent studies, both in computational realms, such as machine learning models that require accurate reaction energies for calibration, and in experimental setups seeking a reliable reference point.
Our methodology included both ab initio calculations at the MP2 level of theory and DFT employing many different functionals.
The results obtained for neutral conditions exhibited reaction barriers greater than 1 eV, which is qualitatively of the same order of magnitude as the existing ab initio results for silica networks [13].
The principal findings of our research indicate that the borate network is significantly more susceptible to hydrolysis in a basic environment, followed by neutral and acidic environment. Indeed, we observed a drop in the activation energy from about 0.5 eV (neutral or acidic environment) to less than 0.1 eV (alkaline environment). These data are significantly lower than those found for silicate structures where the hydrolysis reaction barriers exceeded 1 eV [13].
Furthermore, our study reveals that the inclusion of a greater number of water molecules explicitly into the calculations significantly influences the results when the transition state has a concerted, ring-like topology. Specifically, it reduces the energy of the reaction barriers, with the transition states predominantly forming closed rings composed of water and the B-O-B network. This structural preference in transition geometries demonstrates why including additional explicit water molecules in the calculations results in lower reaction energies.
In terms of accuracy, we observed that for the hydrolysis reactions studied here, the B3LYP, PBE0, and ωB97Xd functionals exhibited an overall better performance, i.e., these functionals qualitatively reproduced the MP2-calculated reaction paths and quantitatively matched the reaction barriers within an approximate range of 0.1–0.2 eV for the most reasonable pathways. CAM-B3LYP also gave a similar accuracy with respect to the MP2 data [19] except for the most complicated reaction pathway involving two transition states and an intermediate.
Still, we must point out that, currently, no approximated functional [40] is known to be able to correctly and dependably reproduce the reaction barriers in chemical reactions. In fact, the description of reaction barriers is problematic for (1) GGA-type functionals because they might underestimate reaction barriers up to the eV scale (due to the very local nature of their functional form), while (2) non-local functionals involving Hartree–Fock exchange like general and range-separated hybrids might also lead to wrong results since they exhibit a spurious long-range repulsive behavior. We have also to point out Hait’s and Head-Gordon’s words: “it is difficult for a single functional to be effective at addressing delocalization error over many species” [41]. This is particularly an issue as boron oxides are not usually included in most libraries on which density functional approximations are tested.
However, in ref. [41], it was observed that functionals belonging to the CAM-B3LYP, ωB97, and PBE0 families showed surprising good behaviors in dealing with delocalization errors in the studied systems with fractional charges. This was also observed in ref. [42].
Also, part of the delocalization issue can be mitigated using dispersion corrections (as we did here) that, at least, should lead to more reliable relaxed structures as they partially account for some of the long-range forces.
While the subject of assessing DFT methods for hydrolysis reactions is a continuously researched subject (see for instance ref. [20]), here, we wanted to point out that we adopted an approach consisting of optimizing the structures at all levels of theory, as some of us already did previously [43,44]; this has the drawback that small changes in the geometry of the optimized structure can significantly change the energies of the models, contributing to the scattered nature of the results and somewhat obscuring the many contributions to the energies [41].
However, the great advantage of this approach is that it allows future investigators to directly apply the findings of this study that also include the structural relaxation, increasing the usefulness of this paper.
Indeed, this study provides valuable insights into the hydrolysis mechanism of borate networks, contributing to a deeper understanding of the process and offering a reliable computational reference for future experimental and computational investigations. The findings underscore the importance of environmental factors, like pH conditions, the composition of the reacting system, particularly the role of water molecules, in determining the hydrolysis behavior of borate networks, and the importance of the topology of the transition state(s). The agreement of our DFT results with the MP2 calculations reinforces the potential of DFT methods in accurately modeling complex chemical reactions, particularly in cases where the experimental data are scarce or otherwise challenging to obtain.

Author Contributions

Conceptualization, A.P. and M.C.M.; methodology, F.M.-M.; investigation, L.O., F.F. and F.M.-M.; resources, A.P.; data curation, L.O., F.F. and F.M.-M.; writing—original draft preparation, F.M.-M.; writing—review and editing, A.P., M.C.M. and F.M.-M.; visualization, F.M.-M.; supervision, A.P., M.C.M. and F.M.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant received by F.M.-M. named “Fondo di dipartimento per la ricerca anno 2022, Linea: progetti di ricerca” (FAR2022) from the “Dipartimento di Scienze Chimiche e Geologiche” (DSCG) of the Università degli Studi di Modena e Reggio-Emilia (UNIMORE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Balasubramanian, P.; Büttner, T.; Miguez Pacheco, V.; Boccaccini, A.R. Boron-Containing Bioactive Glasses in Bone and Soft Tissue Engineering. J. Eur. Ceram. Soc. 2018, 38, 855–869. [Google Scholar] [CrossRef]
  2. Rahaman, M.N.; Day, D.E.; Sonny Bal, B.; Fu, Q.; Jung, S.B.; Bonewald, L.F.; Tomsia, A.P. Bioactive Glass in Tissue Engineering. Acta Biomater. 2011, 7, 2355–2373. [Google Scholar] [CrossRef] [PubMed]
  3. Kargozar, S.; Mozafari, M.; Ghenaatgar-Kasbi, M.; Baino, F. Bioactive Glasses and Glass/Polymer Composites for Neuroregeneration: Should We Be Hopeful? Appl. Sci. 2020, 10, 3421. [Google Scholar] [CrossRef]
  4. Kapoor, S.; Youngman, R.E.; Zakharchuk, K.; Yaremchenko, A.; Smith, N.J.; Goel, A. Structural and Chemical Approach toward Understanding the Aqueous Corrosion of Sodium Aluminoborate Glasses. J. Phys. Chem. B 2018, 122, 10913–10927. [Google Scholar] [CrossRef] [PubMed]
  5. Stone-Weiss, N.; Pierce, E.M.; Youngman, R.E.; Gulbiten, O.; Smith, N.J.; Du, J.; Goel, A. Understanding the Structural Drivers Governing Glass–Water Interactions in Borosilicate Based Model Bioactive Glasses. Acta Biomater. 2018, 65, 436–449. [Google Scholar] [CrossRef] [PubMed]
  6. Werding, G.; Schreyer, W. Experimental studies on borosilicates and selected borates. Rev. Miner. Geochem. 1996, 33, 117–163. [Google Scholar]
  7. Zhu, F.; Bowron, D.T.; Gärtner, S.; Fang, C.; Zhou, Y.; Liu, H.; Hannon, A.C. Structural analysis of potassium borate solutions. Phys. Chem. Chem. Phys. 2023, 25, 12207–12219. [Google Scholar] [CrossRef]
  8. Nishiyabu, R.; Kubo, Y.; James, T.D.; Fossey, J.S. Boronic acid building blocks: Tools for self-assembly. Chem. Commun. 2011, 47, 1124–1150. [Google Scholar] [CrossRef]
  9. Dai, C.; Cazares, L.H.; Wang, L.; Chu, Y.; Wang, S.L.; Troyer, D.A.; Semmes, O.J.; Drake, R.R.; Wang, B. Using boronolectin in MALDI-MS imaging for the histological analysis of cancer tissue expressing the sialyl Lewis X antigen. Chem. Commun. 2011, 47, 10338–10340. [Google Scholar] [CrossRef]
  10. Sun, X.; Zhai, W.; Fossey, J.S.; James, T.D. Boronic acids for fluorescence imaging of carbohydrates. Chem. Commun. 2016, 52, 3456–3469. [Google Scholar] [CrossRef]
  11. Chan, J.; Dodani, S.C.; Chang, C.J. Reaction-based small-molecule fluorescent probes for chemoselective bioimaging. Nat. Chem. 2012, 4, 973–984. [Google Scholar] [CrossRef]
  12. Lippert, A.R.; Van de Bittner, G.C.; Chang, C.J. Boronate oxidation as a bioorthogonal reaction approach for studying the chemistry of hydrogen peroxide in living systems. Acc. Chem. Res. 2011, 44, 793–804. [Google Scholar] [CrossRef] [PubMed]
  13. James, T.D.; Phillips, M.D.; Shinkai, S. Boronic Acids in Saccharide Recognition; Rowan, A.E., Rowan, S.J., Aida, T., Fraser Stoddart, J., Eds.; RSC Publishing: Cambridge, UK, 2006; ISBN 978-0-85404-537-2. [Google Scholar]
  14. James, T.D.; Shinkai, S. Saccharide sensing with molecular receptors based on boronic acid. Angew. Chem. Int. Ed. 1996, 35, 1910–1922. [Google Scholar] [CrossRef]
  15. Jin, S.; Cheng, Y.; Reid, S.; Li, M.; Wang, B. Carbohydrate recognition by boronolectins, small molecules, and lectins. Med. Res. Rev. 2010, 30, 171–257. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, S.-X.; Zhu, F.-Y.; Liu, H.-Y.; Zhang, W.-Q.; Zhou, Y.-Q.; Fang, C.-H.; Li, H.-B. Hydrolysis mechanism of double six–membered ring pentaborate anion. Chem. Phys. Lett. 2020, 739, 136930. [Google Scholar] [CrossRef]
  17. Jabraoui, H.; Charpentier, T.; Gin, S.; Delaye, J.-M.; Pollet, R. Atomic Insights into the Events Governing the Borosilicate Glass-Water Interface. J. Phys. Chem. C 2021, 125, 7919–7931. [Google Scholar] [CrossRef]
  18. Kagan, M.; Lockwood, G.K.; Garofalini, S.H. Reactive simulations of the activation barrier to dissolution of amorphous silica in water. Phys. Chem. Chem. Phys. 2014, 16, 9294–9301. [Google Scholar] [CrossRef]
  19. Peverati, R.; Truhlar, D.G. Quest for a universal density functional: The accuracy of density functionals across a broad spectrum of databases in chemistry and physics. Phil. Trans. R. Soc. A 2014, 372, 20120476. [Google Scholar] [CrossRef]
  20. Epstein, A.R.; Spotte-Smith, E.W.C.; Venetos, M.C.; Andriuc, O.; Persson, K.A. Assessing the Accuracy of Density Functional Approximations for Predicting Hydrolysis Reaction Kinetics. J. Chem. Theory Comput. 2023, 19, 3159–3171. [Google Scholar] [CrossRef]
  21. Pople, J.A.; Head-Gordon, M.; Raghavachari, K. Quadratic configuration interaction—A general technique for determining electron correlation energies. J. Chem. Phys. 1987, 87, 5968–5975. [Google Scholar] [CrossRef]
  22. Dunning, T.H., Jr.; Hay, P.J. Modern Theoretical Chemistry; Schaefer, H.F., III, Ed.; Plenum: Brooklyn, NY, USA, 1977; Volume 3, pp. 1–28. [Google Scholar]
  23. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  24. Frisch, M.J.; Head-Gordon, M.; Pople, J.A. Direct MP2 gradient method. Chem. Phys. Lett. 1990, 166, 275–280. [Google Scholar] [CrossRef]
  25. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  26. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  27. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2009, 120, 215–241. [Google Scholar]
  29. Zhao, Y.; Truhlar, D.G. Density Functional for Spectroscopy: No Long-Range Self-Interaction Error, Good Performance for Rydberg and Charge-Transfer States, and Better Performance on Average than B3LYP for Ground States. J. Phys. Chem. A 2006, 110, 13126–13130. [Google Scholar] [CrossRef] [PubMed]
  30. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef]
  31. Yanai, T.; Tew, D.; Handy, N. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  32. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef]
  33. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic-behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  34. Perdew, J.P.; Burke, K.; Ernzerhof, M. Errata: Generalized gradient approximation made simple. Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef]
  35. Tao, J.M.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the density functional ladder: Nonempirical meta-generalized gradient approximation designed for molecules and solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [PubMed]
  36. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comp. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  37. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  38. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef]
  39. Caricato, M. Absorption and Emission Spectra of Solvated Molecules with the EOM-CCSD-PCM Method. J. Chem. Theory Comput. 2012, 8, 4494–4502. [Google Scholar] [CrossRef]
  40. Johnson, E.R.; Mori-Sánchez, P.; Cohen, A.J.; Yang, W. Delocalization errors in density functionals and implications for main-group thermochemistry. J. Chem. Phys. 2008, 129, 204112. [Google Scholar] [CrossRef]
  41. Hait, D.; Head-Gordon, M. Delocalization Errors in Density Functional Theory Are Essentially Quadratic in Fractional Occupation Number. J. Phys. Chem. Lett. 2018, 9, 6280–6288. [Google Scholar] [CrossRef]
  42. Cohen, A.J.; Mori-Sánchez, P.; Yang, W. Challenges for Density Functional Theory. Chem. Rev. 2012, 112, 289–320. [Google Scholar] [CrossRef] [PubMed]
  43. Muniz-Miranda, F.; Menziani, M.C.; Pedone, A. DFT and TD-DFT Assessment of the Structural and Optoelectronic Properties of an Organic–Ag14 Nanocluster. J. Phys. Chem. A 2015, 119, 5088–5098. [Google Scholar] [CrossRef] [PubMed]
  44. Muniz-Miranda, F.; Menziani, M.C.; Pedone, A. Assessment of Exchange-Correlation Functionals in Reproducing the Structure and Optical Gap of Organic-Protected Gold Nanoclusters. J. Phys. Chem. C 2014, 118, 7532–7544. [Google Scholar] [CrossRef]
Figure 2. Scheme of reaction pathway for the hydrolysis mechanism at neutral pH with three water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 2 for a complete report on the energy values.
Figure 2. Scheme of reaction pathway for the hydrolysis mechanism at neutral pH with three water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 2 for a complete report on the energy values.
Molecules 29 01227 g002
Figure 3. Scheme of reaction pathway for the hydrolysis mechanism at acidic pH with a single hydroxonium cation. ξ represents the reaction coordinate. The energies are not to scale; see Table 4 for a complete report on the energy values.
Figure 3. Scheme of reaction pathway for the hydrolysis mechanism at acidic pH with a single hydroxonium cation. ξ represents the reaction coordinate. The energies are not to scale; see Table 4 for a complete report on the energy values.
Molecules 29 01227 g003
Figure 4. Scheme of reaction pathway for the hydrolysis mechanism at acidic pH with a single hydroxonium cation and two water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 4 for a complete report on the energy values.
Figure 4. Scheme of reaction pathway for the hydrolysis mechanism at acidic pH with a single hydroxonium cation and two water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 4 for a complete report on the energy values.
Molecules 29 01227 g004
Figure 5. Scheme of reaction pathway for the hydrolysis mechanism at alkaline pH with a single OH anion. ξ represents the reaction coordinate. The energies are not to scale; see Table 5 for a complete report on the energy values.
Figure 5. Scheme of reaction pathway for the hydrolysis mechanism at alkaline pH with a single OH anion. ξ represents the reaction coordinate. The energies are not to scale; see Table 5 for a complete report on the energy values.
Molecules 29 01227 g005
Figure 6. Scheme of reaction pathway for the hydrolysis mechanism at alkaline pH with a single OH anion and 2 water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 5 for a complete report on the energy values.
Figure 6. Scheme of reaction pathway for the hydrolysis mechanism at alkaline pH with a single OH anion and 2 water molecules. ξ represents the reaction coordinate. The energies are not to scale; see Table 5 for a complete report on the energy values.
Molecules 29 01227 g006
Table 2. DFT and MP2 Energies for the stationary points of the reaction pathway at neutral pH with 1 water molecule. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 1 W p H = 7 .
Table 2. DFT and MP2 Energies for the stationary points of the reaction pathway at neutral pH with 1 water molecule. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 1 W p H = 7 .
Level of TheoryStationary PointEnergy/eV
MP2Reagents0
TS2.030
Product −0.015
B3LYPReagents0
TS1.492
Products−0.046
CAM-B3LYPReagents0
TS1.403
Products−0.052
BLYPReagents0
TS1.457
Products0.050
PBE0Reagents0
TS1.338
Products−0.046
HSE06Reagents0
TS1.327
Products−0.077
PBEReagents0
TS1.200
Products−0.054
TPSSReagents0
TS1.233
Products−0.040
M06HFReagents0
TS1.078
Products−0.005
M06Reagents0
TS1.508
Products−0.022
ωB97XDReagents0
TS1.418
Products−0.042
Table 3. DFT and MP2 Energies for the stationary points of the reaction pathway at neutral pH with 3 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 3 W p H = 7 .
Table 3. DFT and MP2 Energies for the stationary points of the reaction pathway at neutral pH with 3 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 3 W p H = 7 .
Level of TheoryStationary PointEnergy/eV
MP2Reagents0
TS0.478
Product −0.189
B3LYPReagents0
TS0.572
Products−0.146
CAM-B3LYPReagents0
TS0.395
Products−0.211
BLYPReagents0
TS0.393
Products−0.347
PBE0Reagents0
TS0.363
Products−0.148
HSE06Reagents0
TS0.112
Products−0.210
PBEReagents0
TS0.252
Products−0.196
TPSSReagents0
TS0.303
Products−0.132
M06HFReagents0
TS0.178
Products0.004
M06Reagents0
TS0.038
Products−0.236
ωB97XDReagents0
TS0.482
Products−0.213
Table 4. DFT and MP2 Energies for the stationary points of the reaction pathway at acidic pH with 1 H3O+ cation. The energies have been rescaled to set the energy of the reagents to zero. The Δ E 1 W p H < 7 and Δ E 1 W p H < 7 energy differences are reported separately in a different column.
Table 4. DFT and MP2 Energies for the stationary points of the reaction pathway at acidic pH with 1 H3O+ cation. The energies have been rescaled to set the energy of the reagents to zero. The Δ E 1 W p H < 7 and Δ E 1 W p H < 7 energy differences are reported separately in a different column.
Level of TheoryStationary PointEnergy/eV Δ E 1 W p H < 7 /eV Δ E 1 W p H < 7 /eV
MP2Reagents0
TS10.4670.467
INT0.384
TS20.401 0.018
Products0.003
B3LYPReagents0
TS10.6230.623
INT0.445
TS20.496 0.050
Products0.080
CAM-B3LYPReagents0
TS10.6180.618
INT0.417
TS20.566 0.128
Products−0.051
BLYPReagents0
TS10.6430.618
INT0.541
TS20.590 0.048
Products−0.079
PBE0Reagents0
TS10.6080.608
INT0.378
TS20.554 0.177
Products−0.049
HSE06Reagents0
TS10.4890.489
INT0.421
TS20.564 0.143
Products−0.059
PBEReagents0
TS10.7190.719
INT0.462
TS20.516 0.055
Products−0.057
TPSSReagents0
TS10.5040.504
INT0.402
TS20.566 0.164
Products−0.043
M06HFReagents0
TS10.3070.307
INT0.225
TS20.587 0.363
Products0.045
M06Reagents0
TS10.5940.594
INT0.413
TS20.528 0.115
Products−0.046
ωB97XDReagents0
TS10.6370.637
INT0.396
TS20.496 0.099
Products−0.062
Table 5. DFT and MP2 Energies for the stationary points of the reaction pathway at acidic pH with 1 H3O+ cation and 2 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The Δ E 3 W p H < 7 and Δ E 3 W p H < 7 energy differences are reported separately in a different column.
Table 5. DFT and MP2 Energies for the stationary points of the reaction pathway at acidic pH with 1 H3O+ cation and 2 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The Δ E 3 W p H < 7 and Δ E 3 W p H < 7 energy differences are reported separately in a different column.
Level of TheoryStationary PointEnergy/eV Δ E 3 W p H < 7 /eV Δ E 3 W p H < 7 /eV
MP2Reagents0
TS10.5300.530
INT0.448
TS20.459 0.011
Products0.095
B3LYPReagents0
TS10.5680.568
INT0.556
TS20.610 0.054
Products0.364
CAM-B3LYPReagents0
TS10.2360.236
INT0.745
TS20.043 −0.695
Products0.591
BLYPReagents0
TS10.5820.582
INT0.527
TS2−0.252 −0.778
Products0.373
PBE0Reagents0
TS10.5490.549
INT0.721
TS20.825 0.104
Products0.314
HSE06Reagents0
TS10.8670.867
INT0.765
TS20.227 −0.538
Products0.465
PBEReagents0
TS10.8360.836
INT0.801
TS20.209 −0.591
Products0.624
TPSSReagents0
TS10.6680.668
INT0.740
TS20.211 −0.529
Products0.617
M06HFReagents0
TS1−0.100−0.100
INT0.266
TS20.310 0.043
Products0.323
M06Reagents0
TS10.1120.112
INT0.491
TS20.020 −0.462
Products0.325
ωB97XDReagents0
TS10.5490.549
INT0.484
TS20.498 0.014
Products0.322
Table 6. DFT and MP2 Energies for the stationary points of the reaction pathway at basic pH with 1 OH. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 1 W p H > 7 .
Table 6. DFT and MP2 Energies for the stationary points of the reaction pathway at basic pH with 1 OH. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 1 W p H > 7 .
Level of TheoryStationary PointEnergy/eV
MP2Reagents0
TS0.098
Product −0.002
B3LYPReagents0
TS0.126
Products−0.062
CAM-B3LYPReagents0
TS0.131
Products−0.073
BLYPReagents0
TS0.270
Products−0.069
PBE0Reagents0
TS0.122
Products−0.036
HSE06Reagents0
TS0.155
Products0.011
PBEReagents0
TS0.371
Products−0.041
TPSSReagents0
TS0.339
Products−0.140
M06HFReagents0
TS0.137
Products−0.672
M06Reagents0
TS0.288
Products−0.275
ωB97XDReagents0
TS0.087
Products−0.042
Table 7. DFT and MP2 Energies for the stationary points of the reaction pathway at basic pH with 1 OH and 2 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 3 W p H > 7 .
Table 7. DFT and MP2 Energies for the stationary points of the reaction pathway at basic pH with 1 OH and 2 water molecules. The energies have been rescaled to set the energy of the reagents to zero. The energy of the TS coincides with Δ E 3 W p H > 7 .
Level of TheoryStationary PointEnergy/eV
MP2Reagents0
TS0.062
Product −0.013
B3LYPReagents0
TS0.080
Products−0.068
CAM-B3LYPReagents0
TS0.089
Products−0.083
BLYPReagents0
TS0.251
Products−0.071
PBE0Reagents0
TS0.083
Products−0.044
HSE06Reagents0
TS0.094
Products0.009
PBEReagents0
TS0.239
Products−0.057
TPSSReagents0
TS0.247
Products−0.148
M06HFReagents0
TS0.099
Products−0.566
M06Reagents0
TS0.165
Products−0.234
ωB97XDReagents0
TS0.076
Products−0.038
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Muniz-Miranda, F.; Occhi, L.; Fontanive, F.; Menziani, M.C.; Pedone, A. Quantum-Chemistry Study of the Hydrolysis Reaction Profile in Borate Networks: A Benchmark. Molecules 2024, 29, 1227. https://doi.org/10.3390/molecules29061227

AMA Style

Muniz-Miranda F, Occhi L, Fontanive F, Menziani MC, Pedone A. Quantum-Chemistry Study of the Hydrolysis Reaction Profile in Borate Networks: A Benchmark. Molecules. 2024; 29(6):1227. https://doi.org/10.3390/molecules29061227

Chicago/Turabian Style

Muniz-Miranda, Francesco, Leonardo Occhi, Francesco Fontanive, Maria Cristina Menziani, and Alfonso Pedone. 2024. "Quantum-Chemistry Study of the Hydrolysis Reaction Profile in Borate Networks: A Benchmark" Molecules 29, no. 6: 1227. https://doi.org/10.3390/molecules29061227

Article Metrics

Back to TopTop