28.02.2013 Views

Handbook of Solvents - George Wypych - ChemTech - Ventech!

Handbook of Solvents - George Wypych - ChemTech - Ventech!

Handbook of Solvents - George Wypych - ChemTech - Ventech!

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

136 Abraham Nitzan<br />

where A is an integration constant and τ L is the longitudinal Debye relaxation time<br />

τ<br />

L<br />

ε<br />

ε τ<br />

e<br />

= D<br />

[4.3.17]<br />

s<br />

The integration constant A is determined from the initial conditions: Immediately following<br />

the switch-on <strong>of</strong> the charge distribution, i.e. <strong>of</strong> D, E is given by E(t=0) = D/εe,so -1<br />

A= ε −ε D<br />

−1<br />

. Thus, finally,<br />

( e s )<br />

1 ⎛ 1 1 ⎞<br />

Et () = D+ ⎜ − ⎟De<br />

ε<br />

⎟<br />

s ⎝εe<br />

εs<br />

⎠<br />

−t/<br />

τL<br />

[4.3.18]<br />

We see that in this case the relaxation is characterized by the time τL which can be very<br />

different from τD. For example, in water εe / εs<br />

≅1/ 40, and while τD ≅ 10 ps,<br />

τL is <strong>of</strong> the order<br />

<strong>of</strong> 0.25ps.<br />

4.3.3 LINEAR RESPONSE THEORY OF SOLVATION DYNAMICS<br />

The continuum dielectric theory <strong>of</strong> solvation dynamics is a linear response theory, as expressed<br />

by the linear relation between the perturbation D and the response <strong>of</strong> E, Eq. [4.3.2].<br />

Linear response theory <strong>of</strong> solvation dynamics may be cast in a general form that does not<br />

depend on the model used for the dielectric environment and can therefore be applied also in<br />

molecular theories. 13,14 Let<br />

H = H + H′<br />

0<br />

[4.3.19]<br />

where H0 describes the unperturbed system that is characterized by a given potential surface<br />

on which the nuclei move, and where<br />

H′=∑X jFj() t<br />

[4.3.20]<br />

j<br />

is some perturbation written as a sum <strong>of</strong> products <strong>of</strong> system variables Xj and external time<br />

dependent perturbations Fj(t). The nature <strong>of</strong> X and F depend on the particular experiment: If<br />

for example the perturbation is caused by a point charge q(t) at position rj, q(t)δ(r-rj), we may<br />

identify F(t) with this charge and the corresponding Xj is the electrostatic potential operator<br />

at the charge position. For a continuous distribution ρ(r,t) <strong>of</strong> such charge we may write<br />

3<br />

H′= ∫d<br />

rΦ() r ρ ( r,t)<br />

, and for ρ( r,t ) = ∑ qj() t δ(<br />

r−r j<br />

j ) this becomes ∑ Φ ( rj ) qj( t)<br />

.<br />

j<br />

Alternatively we may find it convenient to express the charge distribution in terms <strong>of</strong> point<br />

moments (dipoles, quadrupoles, etc.) coupled to the corresponding local potential gradient<br />

tensors, e.g. H′ will contain terms <strong>of</strong> the form μ∇Φ and Q:∇∇Φ where μ and Q are point dipoles<br />

and quadrupoles respectively.<br />

In linear response theory the corresponding solvation energies are proportional to the<br />

corresponding products q, μ and Q: where denotes the usual observable<br />

average. For example, the average potential is proportional in linear response to<br />

q<br />

the perturbation source q. The energy needed to create the charge q is therefore∫ dq′<br />

<br />

2<br />

0<br />

≈( 1/ 2) q ≈(<br />

1/ 2)<br />

q.<br />

Going back to the general expressions [4.3.19] and [4.3.20], linear response theory relates<br />

non-equilibrium relaxation close to equilibrium to the dynamics <strong>of</strong> equilibrium fluctu-

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!