• Aucun résultat trouvé

6.2 Classical hexamer dynamics

6.2.1 Radial distribution functions

Firstly, we focus our attention on the equilibrium, or the static properties of the H+(H2O)6 ion. The latter are obtained by performing QMC-driven Langevin Dynamics using the Clas-sical Momentum-Position Correlator (CMPC) algorithm introduced in Chapter 3. However, since the protonated hexamer in much larger than the Zundel ion, its WF is more complex and one cannot straightforwardly apply the same strategy employed to carry out CPMC-LD simulations of the Zundel cation. Indeed, the number of electronic WF parameters increases linearly with the system size. In practice, if one uses exactly the same JAGP WF employed to describe the Zundel ion in Chapter 4, 15763 parameters should be optimized between each LD step, against only 2391. Consequently, the part dedicated to the WF optimization, which is the most time-consuming, would be multiplied at least by a factor of 5. We have the further constraint that the generated MD trajectories should be long enough to display several spontaneous PT proceses inside the cluster, implying that they should last at least 20-25 ps. Therefore, using the JAGP WF of Ref. 72 is not viable in a reasonable amount of computational time.

To solve that issue, we apply a Geminal embedding scheme[202] to the Jastrow factor (this technique has already been used for the AGP part of the WF), using the O[6]H[2] contraction described in Chapter 5. In this way, a significant fraction of the water dimer binding energy is recovered, with a reasonable number of electronic parameters. Thanks to this strategy, we have reduced the number of electronic parameters to 6418, representing a significant gain.

Furthermore, since the Hilbert space of the electronic parameters is larger than in the Zundel case, the WF optimization, using the SRH method, is less stable and one should reduce the acceleration of the electronic parameters δtpar = 0.3 (see Eq. (C.9)). We also increased the number of MC samples to ' 4.0×105 to provide an accurate evaluation of the ionic forces and energy derivatives at each MD step. These choices allow reaching an accuracy of 3mHa (1.9 kcal/mol) in the total energy per VMC energy minimization step. Thus, the statistical error on the Born-Oppenheimer (BO) surface sampling is slightly larger than the Zundel one, but is sufficient not to spoil the quality of the protonated hexamer dynamics.

To quantify the impact of thermal effects on the proton dynamics in the H+(H2O)6 clus-ter, we decided to generate 5 trajectories of 25 ps, propagated using the CPMC algorithm at various temperatures ranging from 200 to 400 K. This choice is motivated by the fact that very often PT reactions occur at room temperature, in chemical or biological systems. The Langevin dampingγBOand the time stepδtemployed to propagate the CMPC-LD dynamics are the same that those employed for the study of the Zundel ion (Chapter 4). We verified that the average simulation temperature is compatible with the target temperature of the considered calculation, within a 1% error. This result ensures that the sampling of the BO surface remains correct during the dynamics.

We represent in Figure 6.6 the normalized RDFs gOO for the two central oxygen atoms O1 and O2 for various temperatures ranging from 200 to 400 K. The vertical lines indicate the average distance hdO1O2i for each simulation. If we compare the protonated hexamer gOO with those extracted from MD simulations of the Zundel cation at room temperature (upper

right panel of Figures4.7and4.8), we notice that the shape of the function is almost identical, but the peak position is shifted by about∼0.05 Å. For instance, the averageO1O2 distance is equal tohdO1O2i= 2.447(5) Å at room temperature (T= 300 K), being ∼0.054Å larger than the zero temperature result. Consequently, going back to Figure 6.5, it seems that at room temperature, the H+(H2O)6 cluster mainly adopts an Eigen-like configuration, where the hydrated proton is covalently bound to one side water molecule. Therefore, the presence of solvating water molecule drastically reduces the mobility of the hydrated proton, which is much more localized with respect to the Zundel case, where it freely oscillates between the 2 neighboring water molecules.

The evolution of the shape of the gOOfunctions when one increases the temperature is some-what expected. Indeed, the low-temperature (T = 200 K) distribution is sharper than the high-temperature (T = 400 K) one, which displays longer tails in the large OO distances region. This systematic broadening can be interpreted as follows. We provide more thermal energy to the system, thus increasing its kinetic energy and the average velocity of the par-ticles. Consequently, the amplitude of the intra- and intermolecular vibrations is enhanced and larger values of the oxygen-oxygen distance are obtained. This is also consistent with the strongly anharmonic shape of the protonated hexamer PES (Figure 6.3) that becomes very flat for large O1O2. Finally, we emphasize that the temperature dependance of the oxygen-oxygen RDFs is monotonic and saturates at 350 K, confirming that thermal effects are straightforward to interpret.

From now on, we will analyze the temperature-dependence of the normalized oxygen-proton RDFs gOH that are plotted in Figure 6.7. At variance with the gOO, the protonated water hexamer gOH are completely different from the computed RDFs of the Zundel ion (bottom panels of Figures4.7and 4.8). Indeed, they are characterized by the presence of two distinct peaks around dOH = 1.1Å and dOH = 1.35Å , corresponding to a localized proton, covalently bound to its closest neighbor. We thus confirm that the more representative configuration of the protonated hexamer in the classical case is Eigen-type contributing to the long-range tails in the gOO. Looking closer to Figure 6.7, we notice that the height of the two peaks decreases with the temperature, and the distributions are broadened. Indeed, from T= 350 K, the second maximum around dOH = 1.35 Å , starts to disappear. This suggests that at this temperature, the ambient thermal energy is large enough to trigger spontaneous proton jumps inside the core of the protonated hexamer. Such proton hops involve the presence of a fully Zundel-like reaction intermediate, explaining the little increase of the gOH around dOH = 1.2 Å . Furthermore, we notice that the low-temperature (T = 200 K) gOH displays a first peak at slightly largerOH distance, compared with higher temperature RDFs where the peak position remains unchanged. This can be related to an enhanced proton mobility at higher temperatures.

To complete the analysis of the static properties of the classical protonated hexamer, we now analyze the bidimensional oxygen-oxygen and oxygen-proton probability distributions ρ2D, represented in Figure 6.8. This probability distribution provides a graphical representation of the protonated hexamer PES at finite temperature projected along the reaction coordinate O1O2. To further analyze the impact of thermal effects on the phase space configurations

2.3 2.4 2.5 2.6

0 0.05

0.1 200 K

g

OO

(r) (1/Å

3

)

2.3 2.4 2.5 2.6

0 0.05

0.1 250 K

2.3 2.4 2.5 2.6

0 0.05

0.1 300 K

2.3 2.4 2.5 2.6

0 0.05

0.1 350 K

2.3 2.4 2.5 2.6

r (Å) 0

0.05

0.1 400 K

Figure 6.6 – Oxygen-oxygen RDFs obtained by QMC-driven CMPC-LD simulations at various temperatures ranging from 200 K to 400 K. The vertical lines indicate thehdO1O2ifor each simulation.

0 0.1 0.2 0.3 0.4

0.80 1.00 1.20 1.40 1.60 1.80

gOH(r) (1/Å3 )

dOH (Å)

200 K 250 K 300 K 350 K 400 K

Figure 6.7 –Oxygen-proton RDFs obtained by QMC-driven CMPC-LD simulations at various tem-peratures ranging from 200 K to 400 K.

that may be visited by the hydrated proton, we also plot the equilibrium geometries obtained by VMC at T= 0 K (black circles).

First of all, we notice that the global shape of the distributionsρ2D is very similar to those obtained for the Zundel cation, forming two wings that stretch along the equilibrium geome-tries for largeOO distances. This expected result confirms that the presence of the solvation shell does not drastically impact the global core geometry with no proton hopping from the Zundel core to the outer shell. At low temperature (T = 200 K), the configurations in the range dO1O2 = 2.38−2.45Å are predominantly visited, while this region becomes larger and larger when one increases the cluster temperature: dO1O2 = 2.38−2.53 Å at T= 300 K and dO1O2 = 2.35−2.58 Å at T = 400 K), in agreement with the gOO. However, the striking difference with the Zundel case is the depletion of the densityρ2D in the short OO distances region at low and room temperature. This stems from the quasi absence of symmetric or Zundel-like configurations, confirming again that, in the presence of classical particles, the protonated hexamer is described by an Eigen-like motif. However, when looking at the high-temperature distribution (T = 400 K), the density depletion, although still present, is less noticeable. The thermal energy is thus large enough to enable PT processes inside the core of the hexamer, confirming that the hydrated proton can classically cross the static barrier. The proton jumps are certainly less easy than in the Zundel complex since they imply aconcerted rearrangement of the solvating molecules, which is more constraining.

The study of the structural properties of the classical protonated hexamer enabled us to stress the critical role played by the temperature to trigger spontaneous PT processes. Indeed, at low temperature, the hydrated proton is frozen around its Eigen-like minimum energy configuration. When the temperature becomes higher than the ambient one, the covalent and H-bonds fluctuations are large enough to enable proton jumps within the core, implying the hexamer goes through a symmetric Zundel-like transition state. These conclusions are however based on thermal equilibrium distribution functions, and must be confirmed by the analysis of the time-dependent properties of the system.