the dynamics of polar solvation castner...364 e. w. castner et al.: the dynamics of polar solvation...

10
E. W. Castner et al.: The Dynamics of Polar Solvation 363 References [I] D. J. Clouthier and D. A. Ramsay, Ann. Rev. Phys. Chern. 34,31 (1983). [2] C. B. Moore and J. C. Weisshaar, Ann. Rev. Phys. Chern. 34, 525 (1983). [3] J. C. D. Brand and C. G. Stevens, J. Chern. Phys. 58, 3331 (1973). [4] P. Kusch and F. W. Loomis, Phys. Rev. 55, 850 (1939). [5] F. W. Birss, D. A. Ramsay, and S. M. Till, Chern. Phys. Lett. 53, 14 (1978). [6] D. A. Ramsay and S. M. Till, Can. J. Phys. 57, 1224 (1979). [7] C. M. L. Kerr, D. C. Moule, and D. A. Ramsay, Can. J. Phys. 61,6(1983). [8] D. J. Clouthier, A. M. Craig, and D. A. Ramsay, Can. J. Phys. 61,1073 (1983). [9] M. Barnett, D. A. Ramsay, and S. M. Till, Chern. Phys. Lett. 65, 440 (1979). [10] D. J. Clouthier, D. C. Moule, D. A. Ramsay, and F. W. Birss, Can. J. Phys. 60,1212 (1982). [11] D. J. Clouthier, D. A. Ramsay, and F. W. Birss, J. Chern. Phys. 79, 5851 (1983). [12] K. H. Fung, J. C. Petersen, and D. A. Ramsay, Can. J. Phys. 63, 933 (1985). [13] M. Lombardi, R.Jost, C. Michel, and A. Trarner, Chern. Phys. 46,273 (1980); 57, 341, 355 (1981). [14] K. H. Fung and D. A. Ramsay, J. Phys. Chern. 88, 395 (1984). [15] J. C. Petersen, D. A. Ramsay, and T. Amano, Chern. Phys. Lett. 103, 266 (1984). [16] J. C. Petersen, S. Saito, T. Amano, and D. A. Ramsay, Can. J. Phys. 62,1731 (1984). [17] J. C. Petersen, T. Amano, and D. A. Ramsay, J. Chern. Phys. 81,5449 (1984). [18] T. Suzuki, S. Saito, and E. Hirota, J. Chern. Phys. 79, 1641 (1983). [19] J. C. Petersen and D. A. Ramsay, Chern. Phys. Lett. 118,31 (1985). [20] J. C. Petersen and D. A. Ramsay, Chern. Phys. Lett. 124,406 (1986). [21] Q. S. Zhu and D. A. Ramsay, unpublished result. Presented at the Discussion Meeting of the E 6648 Deutsche Bunsen-Gesellschaft fur Physi- kalische Chemie "Intramolecular Proc- esses", in Grainau, August 17th to 20th, 1987 The Dynamics of Polar Solvation E. W. Castner, Jr., B. Bagchi, M. Maroncelli, S. P. Webb, A. J. Ruggiero, and G. R. Fleming*) Department of Chemistry and the James Franck Institute, The University of Chicago, Chicago, Illinois 60637 Computer Experiments / Dielectrics / Fluorescence / Non-equilibrium Phenomena Fluorescence upconversion experiments with subpicosecond resolution are described. The time dependent fluorescence spectra of several probe molecules In polar solvents are measured and used to construct the Stokes shift correlation function C(I). The correlation functions decay nonexponentially in contrast to the predictions of simple Debye/Onsager continuum theory. At high values of the static dielectric the times are much th.an predicted from standard continuum theory. Generalizations of continuum theory to mclude non-spherical shapes, response and saturation effects are discussed. The experimental data are qualitatively In accord With predictions based on a frequency and position dependent dielectric response. I. Introduction The study of the dynamics of the interaction between sol- utes and solvents has become an active area of research. Experimental and theoretical studies have been carried out with the goal of understanding the phenomena of how the solvent environment adapts to a rapid change in the electric properties, such as charge distribution and dipole moment, of a probe molecule or ion. It has become clear that the polar solvent can playa dominant role in the energetics and dynamics of electron transfer reactions. For example, in adi- abatic electron transfer reactions the rate of reaction is pro- posed to be inversely proportional to the solvent longitu- dinal relaxation time TL [1, 2]. TL is derived via a continuum model from the experimentally obtained solvent dielectric relaxation time To [3 - 5]. However, there have been rela- tively few determinations of solvent response in situations where TL should be observed directly. In this paper we dis- cussour recent work on the dynamics of dipolar solvation. Our experimental work has been devoted to the study of the time dependent fluorescence Stokes shift (TDFSS) of *) John Simon Guggenheim Fellow. highly polar fluorophores in polar solvents [6- 8]. Com- plementary theoretical work on the study of dynamical di- electric continuum models has been undertaken in an at- tempt to understand the experimental results. In Section II of this paper a brief discussion of the ex- perimental techniques used to measure the ultrafast fluores- cence emission is given. The methods used to convert the fluorescence decay data from the time domain to the (op- tical) frequency domain are described, along with the method for constructing the correlation function used to compare theory and experiment. Recent results on the TDFSS obtained in our laboratories are included in Section II. Section III is a description of theoretical work on the TDFSS. Three new extensions of the current dielectric con- tinuum dynamical models are introduced. First, predictions for the TDFSS obtained using non-Debye formulae for the frequency dependent dielectric susceptibility are shown to predict non-exponential TDFSS behavior that is in better accord with some of the experimental results. Next is a study of how the shape of a probe molecule represented by the spherical cavity in the continuum models affects the solva- tion dynamics. Finally, the effect of saturation of the dielec- Ber, Bunsenges. Phys. Chern. 92, 363-372 (1988) - (() VCH Verlagsgesellschaft rnbH, D-6940 Weinheirn, 1988. 0005 - 9021/88/0303 - 0363 $ 02.50/0

Upload: others

Post on 25-Feb-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

E. W. Castner et al.: The Dynamics of Polar Solvation 363

References

[I] D. J. Clouthier and D. A. Ramsay, Ann. Rev. Phys. Chern.34,31 (1983).

[2] C. B. Moore and J. C. Weisshaar, Ann. Rev. Phys. Chern. 34,525 (1983).

[3] J. C. D. Brand and C. G. Stevens, J. Chern. Phys. 58, 3331(1973).

[4] P. Kusch and F. W. Loomis, Phys. Rev. 55, 850 (1939).[5] F. W. Birss, D. A. Ramsay, and S. M. Till, Chern. Phys. Lett.

53, 14 (1978).[6] D. A. Ramsay and S. M. Till, Can. J. Phys. 57, 1224 (1979).[7] C. M. L. Kerr, D. C. Moule, and D. A. Ramsay, Can. J. Phys.

61,6(1983).[8] D. J. Clouthier, A. M. Craig, and D. A. Ramsay, Can. J. Phys.

61,1073 (1983).[9] M. Barnett, D. A. Ramsay, and S. M. Till, Chern. Phys. Lett.

65, 440 (1979).[10] D. J. Clouthier, D. C. Moule, D. A. Ramsay, and F. W. Birss,

Can. J. Phys. 60,1212 (1982).[11] D. J. Clouthier, D. A. Ramsay, and F. W. Birss, J. Chern.

Phys. 79, 5851 (1983).[12] K. H. Fung, J. C. Petersen, and D. A. Ramsay, Can. J. Phys.

63, 933 (1985).

[13] M. Lombardi, R. Jost, C. Michel, and A. Trarner, Chern. Phys.46,273 (1980); 57, 341, 355 (1981).

[14] K. H. Fung and D. A. Ramsay, J. Phys. Chern. 88, 395 (1984).[15] J. C. Petersen, D. A. Ramsay, and T. Amano, Chern. Phys.

Lett. 103, 266 (1984).[16] J. C. Petersen, S. Saito, T. Amano, and D. A. Ramsay, Can.

J. Phys. 62,1731 (1984).[17] J. C. Petersen, T. Amano, and D. A. Ramsay, J. Chern. Phys.

81,5449 (1984).[18] T. Suzuki, S. Saito, and E. Hirota, J. Chern. Phys. 79, 1641

(1983).[19] J. C. Petersen and D. A. Ramsay, Chern. Phys. Lett. 118,31

(1985).[20] J. C. Petersen and D. A. Ramsay, Chern. Phys. Lett. 124,406

(1986).[21] Q. S. Zhu and D. A. Ramsay, unpublished result.

Presented at the Discussion Meeting of the E 6648Deutsche Bunsen-Gesellschaft fur Physi-kalische Chemie "Intramolecular Proc-esses", in Grainau, August 17th to 20th,1987

The Dynamics of Polar SolvationE. W. Castner, Jr., B. Bagchi, M. Maroncelli, S. P. Webb, A. J. Ruggiero, and G. R. Fleming*)

Department of Chemistry and the James Franck Institute, The University of Chicago, Chicago, Illinois 60637

Computer Experiments / Dielectrics / Fluorescence / Non-equilibrium Phenomena

Fluorescence upconversion experiments with subpicosecond resolution are described. The time dependent fluorescence spectra of severalprobe molecules In polar solvents are measured and used to construct the Stokes shift correlation function C(I). The correlation functionsdecay nonexponentially in contrast to the predictions of simple Debye/Onsager continuum theory. At high values of the static dielectric~onstants the solv~tion times are much slo~er th.an predicted from standard continuum theory. Generalizations of continuum theory tomclude non-spherical shapes, non~Debye ?l~lectrIc response and saturation effects are discussed. The experimental data are qualitatively

In accord With predictions based on a frequency and position dependent dielectric response.

I. Introduction

The study of the dynamics of the interaction between sol­utes and solvents has become an active area of research.Experimental and theoretical studies have been carried outwith the goal of understanding the phenomena of how thesolvent environment adapts to a rapid change in the electricproperties, such as charge distribution and dipole moment,of a probe molecule or ion. It has become clear that thepolar solvent can playa dominant role in the energetics anddynamics of electron transfer reactions. For example, in adi­abatic electron transfer reactions the rate of reaction is pro­posed to be inversely proportional to the solvent longitu­dinal relaxation time TL [1, 2]. TL is derived via a continuummodel from the experimentally obtained solvent dielectricrelaxation time To [3 - 5]. However, there have been rela­tively few determinations of solvent response in situationswhere TL should be observed directly. In this paper we dis­cuss our recent work on the dynamics of dipolar solvation.Our experimental work has been devoted to the study ofthe time dependent fluorescence Stokes shift (TDFSS) of

*) John Simon Guggenheim Fellow.

highly polar fluorophores in polar solvents [6- 8]. Com­plementary theoretical work on the study of dynamical di­electric continuum models has been undertaken in an at­tempt to understand the experimental results.

In Section II of this paper a brief discussion of the ex­perimental techniques used to measure the ultrafast fluores­cence emission is given. The methods used to convert thefluorescence decay data from the time domain to the (op­tical) frequency domain are described, along with themethod for constructing the correlation function used tocompare theory and experiment. Recent results on theTDFSS obtained in our laboratories are included inSection II.

Section III is a description of theoretical work on theTDFSS. Three new extensions of the current dielectric con­tinuum dynamical models are introduced. First, predictionsfor the TDFSS obtained using non-Debye formulae for thefrequency dependent dielectric susceptibility are shown topredict non-exponential TDFSS behavior that is in betteraccord with some of the experimental results. Next is a studyof how the shape of a probe molecule represented by thespherical cavity in the continuum models affects the solva­tion dynamics. Finally, the effect of saturation of the dielec-

Ber, Bunsenges. Phys. Chern. 92, 363-372 (1988) - (() VCH Verlagsgesellschaft rnbH, D-6940 Weinheirn, 1988.0005 - 9021/88/0303 - 0363 $ 02.50/0

Page 2: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

364 E. W. Castner et al.: The Dynamics of Polar Solvation

L5

Nd YAG Laser

Ff

1.00mmurea

L3!L4

L1SL2

c

I.O-I.5W532om

Fig. 1Synchronously pumped hybrid mode-locked cavity-dumped tuna­ble femtosecond dye laser. 1.2 W of 532 nm radation is used topump the vertically flowing Rhodamine 6G gain jet. The quartzprisms allow for tunable net cavity group velocity dispersion, withthe variable slit at the end of the cavity (after the prism) providingthe wavelength selection. A second pair of fold mirrors surroundsthe DQOCI saturable absorber jet. A cavity dumper serves to in­crease the peak power of the pulses 20-fold while lowering therepetition rate. At 580 nm, the laser can produce 200 fs (FWHM)pulses at 3.8 MHz with 75 mW of average power. The five cavity

clements set at the Brewster angle provide linear polarization

SYNCHRONOUSLY PUMPEOHYBRID LASER

II. Experimental

The fluorescence from the first singlet state of a polar probemolecule in solution is measured by either of two techniques; timecorrelated single photon counting, or fluorescence upconversion.The photon counting technique has been described previously [10]and is the technique of choice for experiments on solvents whichhave relatively slow dielectric relaxation times, as the signal to noisecharacteristics are excellent, and it is a simple matter to refrigeratethe sample. However, the time resolution is currently limited toabout 20 ps, with a time window of about 5 ns on this level ofresolution. Many of the polar solvents of great interest, viz. thepolar aprotics such as DMSO, acetonitrile, etc. are predicted bycontinuum theories [3 - 5] to have solvation dynamics with ex­ponential lifetimes in the range of 0.2 - lOps at room temperature.Thus it is necessary to use a technique with superior time resolutionto accurately probe solvation dynamics. The fluorescence techniquewith the best time resolution is fluorescence upconversion [6,11- 13]. However, the upconversion technique is limited by therequirement of a mechanical optical delay, and thus the time win­dow of - 500 ps (in our lab) is not sufficient to record the entirefluorescence decay for most probe molecules. The optimal experi­mental situation is to use both photon counting and upconversionin a complementary way.

The dynamical Stokes shift can be related to solvation propertiesof a particular solute/solvent system by means of the time-corre­lation function C(l) [4].

tric susceptibility near a large dipole (analogous to the sat­uration predicted by Debye near an ion [9]) upon theTDFSS is considered.

Section IV is a discussion of the state of experimentalwork in the field of dipolar solvation in the context of cur­rent theory.

The Appendix is an extensive summary of fits to the tem­perature dependent dielectric dispersion data for several al­cohols.

C (I) = _i'_(t_)_-_v_(00_)v(O) - v(oo)

(1)

PBSl XI2

v(t) is defined to be the first moment of the emission spectrum atsome time 1. It would be preferable to be able to measure the entirefluorescence spectrum at any desired time delay. However, the needto obtain subpicosecond time resolution makes this very difficultbecause of the width of the emission spectrum of the probe mole­cules considered so far. The procedure adopted here is to measurethe emission spectrum over the time region of interest for severalwavelengths, e. g. fifteen. These spectra are then transformed intocomplete emission spectra at any desired time delay by a numericalcut-and-paste procedure [6, 7]. The resulting time resolved spectraare used to construct C(t).

The upconversion experiment has been described previously [6,8] but several improvements have since been made. The improve­ments to the upconversion experiment are in three areas. First, anew type of sub-picosecond dye laser was used (Fig. 1). This laseris synchronously pumped by the frequency doubled output of a cwmode-locked Quantronix 416 Nd J

; - YAG laser. The dye laser isa dispersion compensated hybrid mode-locked type of a design verysimilar to that of Dawson et al. [14- 16] with the addition of acavity dumper. The cavity dumper increases the peak power of theoutput pulse train by a factor of more than twenty-fold, while sub­picosecond pulses are still generated.

We have also changed from Type I (LiIOJ , KDP) sum frequencygeneration in LifO, to Type" mixing in urea. The major source ofnoice in frequency conversion optical gating experiments utilizingthe dye laser fundamental as both an excitation and gate pulse isthe second harmonic light generated along with the fluorescenceupconversion signal. The use of the Type" mixing scheme largelyeliminates this problem and greatly increases the SIN ratio of themeasurement. The perpendicular polarizations of the input beams

Fig. 2Type" phase matched upconversion experiment. The linear, hor­izontally polarized pulse train from the hybrid laser is directed uponthe first polarizing beam splitter cube (PBSt) after having the po­larization rotated by a half-wave plate (A/2). The transmitted por­tion is directed along the stepper motor controlled variable delaystage. The fluorescence is generated in a 1.0 mm quartz sample flowcell (S), and is collected by a microscope objective (L2). The or­thogonally polarized fluorescence and variably-delayed laser pulsesare focused into the Type" phase matched urea crystal. SPl/PMTland SP2/PMT2 are the spectrometer/photomultiplier systemswhich isolate and detect the fluorescence upconversion and laserpulse autocorrelation signals, respectively. The upconversion de­tection is by means of single photon counting, and the autocorre­lation is detected with a lock-in amplifier. Ls are lenses, PI a po­larizer, C a light chopper, and Fl is a Schott UGI 1 UV pass filter

also facilitate efficient separation and recombination of the laserand fluorescence with polarizing optics. This mixing scheme cou­pled with single photon counting detection has resulted in a dra­matic improvement in sensitivity and signal quality over the lock­in detection and Type I mixing used previously. The counting sta­tistics also allow a meaningful comparison of the reduced chi­squared values obtained in the fitting of the upconversion data tothe TCSPC data. In addition, the simultaneous acquisition of abackground free autocorrelation signal through the upconversionoptics is possible in this configuration. This has several advantages

Page 3: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

E. W. Castner et al.: The Dynamics of Polar Solvation 365

1500O. 30.1 0.2

T i ml;l (T)o

"--,

" -.p =1.0 {DebyeI

-,

"

o-4

c~ -3

-1

o

......'"8L<,

';} -2v

.s

Fig. 4Calculated TDFSS response vs. reduced time for the continuummodel using the Davidson-Cole form of the dielectric response. Co

= 25.0, CCf~ = 1.0, and the D-C exponent P(Eqn. (8)) values arebetween 0.4 and t.O

1000(JJ....c ....:J .~ ' . , .0u .... o·

"500

over the previous simultaneous autocorrelation measurementwhich was done with a separate nonlinear crystal. It provides botha more accurate response function for data analysis by standardnonlinear least squares fitting with iterative convolution, and anaccurate determination of time zero. It should be noted howeverthat the successful implementation of this technique is dependenton using focusing conditions that allow second harmonic genera­tion over the entire frequency bandwidth of the laser pulse. A sche­matic of our current upconversion spectrometer is shown in Fig. 2.Representative data from the upconversion experiment is shown inFig. 3 for the Nile Red molecule in methanol solution.

III. Continuum Theories

a) The Form of the Dielectric Response

In the continuum approximation the potential energyfrom polar interactions can be written in the form [4J

Fig. 3Representative data obtained from the Type II phase upconversionexperiment. The sample is t0 4 M Nile Red in spectro grade Meth­anol. The solid curve is the nonlinear least squares fit to the data(after deconvolution of the autocorrelation by means of iterativereconvolution); the weighted residuals of the nonlinear least squaresfit are displayed at the top of the graph. The time per channel is0.2 ps. The two fluorescence components are nearly equallyweighted with lifetimes of t8 and 600 ps. Laser excitation was at580 nm, and the fluorescence phase matched for detection by up-

conversion was 6t5 nm

r(t) = A exp( -tITd.

where To is the dielectric relaxation time, then r(t) takes theform

(5)

(4)

When the frequency dependent dielectric response is givenby the familiar Debye form:

Thus in this approximation the Stokes shift correlation func­tion, C (t) should relax with a single exponential time con­stant TL given by

1:0- !:c(w) = Eoo + 1 . ex.+ 1WTo

I

R(t)= J dt'r(t')tl(t-t')·

This equation can be Fourier inverted to give the time de­pendence of the reaction field

(2)

50 100T imQ Cps)

oo

E(t) = -tl(t) . R(t)

TL is the solvent longitudinal relaxation time. If the dielectricresponse corresponds to a sum of two Debye responses, thenC (t) will be a sum of exponentials [4].

Many solvents, particularly associated liquids, do nothave an E(W) adequately described by the Debye form [9,18, 19]. The Davidson-Cole form [17J,

where R(t) is the time dependent reaction field and tl(t) isthe dipole moment of the solute at time t. Since the shift inthe fluorescence spectrum, via Eq. (1), can be directly relatedto E(t) the problem reduces to the calculation of R(t). R(t)is obtained by Fourier transform of the frequency domainreaction field R (w). A quasi-static boundary-value calcula­tion for the reaction field of a non-polarizable point dipoleembedded in a vacuum cavity gives

(28 00 + 1)TL = (2eo + 1) To;

(6)

R(w) = r(w) tl(w)

where (3)

EO - E.x •

E(W) = Eoo + (1 . )fi+ 1WTo(7)

2 8(w)-1r(w) = a3 28(W) + I '

has a form that is skewed progressively more towardsshorter relaxation times as [3 is reduced from unity. Manypolar liquids have an E(W) well described by the Davidson-

Page 4: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

366 E. W. Castner et al.: The Dynamics of~P-=o::la r~Solv~a=-:t~ion~__~ _

We found empirically that the stretched exponential expo­nent IX was related to the Davidson-Cole exponent f3 via

Cole (D-C) equation [18, 19]. Examples are glycerol (f1 =

0.7 at 285 K), acetonitrile (f3 = 0.8), glycerol triacetate (f3 =0.2 -0.4), the haloalkanes and mixtures of various polar sol­vents. Molecules with internal rotations and associative liq­uids are classes of compounds for which the D-C formworks well. Because of this wide range of applicability it isof interest to calculate the form C (r) takes for this form ofdielectric response. The Davidson-Cole form has also re­cently been used in a theory of adiabatic electron transfer.reactions [20].

A more detailed description of the calculation of R (t) andhence Crt) for the Davidson-Cole form is given elsewhere[21]. The procedure is identical to that described by Bagchiet al. [4], except in this case r(t) must be obtained by anumerical inverse Fourier transform. The results for f3 valuesranging from 1 (Debye) to 0.5 are shown in Fig. 4. As ex­pected when f3 = 1 a single exponential decay with timeconstant TL is obtained. As f3 decreases and high frequencycomponents become more important the initial decay be­comes much more rapid than the Debye case. Even a smalldeviation from Debye behavior (e.g. f3 = 0.9) gives a highlynonexponential response with a substantially faster initialdecay, and a slower decay than 'L at long times. Solventswith f3 as low as 0.4 are known experimentally, and so it isclear that solvation may proceed in such solvents 1- 2 or­ders of magnitude faster than expected from a Debye model.This finding may be relevant to recent discussions of fem­tosecond hole burning experiments [22, 23].

The curves in Fig. 4 can be fit to the sum of two expo­nentials, however, they fit better to the stretched exponentialform [24, 25]

r(t) = r(O)exp[ - (t/,)"]. (8)

(9)

c) Saturation Effects

As will be seen in the summary of experimental data inthe following section, setting aside for the moment questionsof the detailed time dependence of the solvation, continuummodels work fairly well for low to medium dielectric con­stant solvents but appear to break down completely forhigher dielectric constant solvents.

As the dielectric constant of the solvent increases, theeffective polar interactions at short range become progres­sively more important because they are more effectivelyscreened by the solvent molecules. Around each dipolar sol­ute molecule, there exists a region where solvent moleculestend to align with the dipolar field of the solute molecule.In this region, the effective screening of the dipolar field islessened because of the alignment of the nearest neighbormolecules. In other words, if a measurement of the localdielectric constant of the liquid around a dipolar molecule(having a large dipole moment) in a strongly polar solventcould be made (by measuring, for example, the response toa weak external field), then a value of the dielectric constantthat is lower than the bulk value of the solvent should beobserved. Thus the dielectric constant around a dipolar mol­ecule should become position dependent. This is known asthe saturation effect [9], and is not included in the usualcontinuum model calculations. This saturation effect is ex­pected to be more important when the solute particle issmall, has a large dipole moment, and the solvent is highlypolar. Such a saturation, or inhomogeneity in the dielectricresponse, may have a drastic effect on the time dependenceof the reaction field. This is because the contribution of thenearest-neighbor molecules to the dynamics of the reactionfield may vary on a significantly slower time scale than thosein the bulk solvent far from the dipole. This will lead to anoverall slower variation of the polarization P. This pointhas also been addressed in a recent paper by Loring andMukamel [26].

The most general treatment of the saturation effect shouldbe based on the following relation [26 - 30]

where C, and C2 are constants. Work is in progress to see D(k,w) = ~ (k,w)E(k,w)if C, and C2 can be derived in closed form.

(10)

for the displacement vector D. We can now write a muchsimpler equation for the Laplace equation in D.

between the displacement vector D, electric field E, and thedielectric tensor s (k, w). The quantities depend both on fre-

'"quency wand wavevector k. Loring and Mukamel [26] havegiven a systematic method of incorporating position de­pendence of e by starting from Eqn. 11. Here we will discuss

'"a simpler approach involving a continuum calculation withposition dependent dielectric constant. A full description ofthis work will be published elsewhere [31].

To include saturation effects in the continuum treatmentof the TDFSS, we must obtain the time dependent reactionfield R (r), A number of assumptions are made about thesystem to make the problem tractable. First, we assume thatthe dielectric tensor in r-space is diagonal in r, and is a scalarfunction of r. This allows us to write

b) The Influence of Molecular Shape

Theoretical models of solvation dynamics have so far con­sidered only spherical molecules. In view of the well-knowninfluence of deviations from spherical shape in molecularreorientation we have investigated this aspect of the solva­tion dynamics. Details of the calculation are presented el­sewhere [21]. Calculations were carried out for a Debyeform of e(w) for both oblate and prolate ellipsoids. Evenwith a Debye form non-spherical molecules exhibit solvationdynamics which deviates from single exponential. Bothoblate and prolate ellipsoids give double exponential re­sponses. As an example for eo = 50 and eeY. = 2, 'Lsphcre =

0.0495 'D' For an oblate molecule with a ratio of major tominor semiaxes of 2.5 the two time constants are ,t = 0.0661'D and 'tt = 0.0450 'tD. For a prolate molecule with an axialratio of 2.5, 'tt = 0.0440 'D and ,t = 0.0553 'D. The differ­ence in time constants, particularly for the oblate case maybe large enough to permit experimental observation.

D(r,w) = ~ (r,w)E(r,w) (tt)

Page 5: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

E. W. Castner et al.: The Dynamics of Polar Solvation 367

and

eo (r) = eo (bulk) + (eo (a) - eo(bulkjj expt -(r - a)1A). (13b)

eo (r) = eo(a)+ (eo (bulk) - eo (a» tanh(r - all A) (13a)

0.6O. 2 O. 4Time (To)

o-4

Fig. 6Calculated TDFSS response vs. reduced time including saturationeffects. Bo(bulk) = 50, Bo(a) = 1.0. Cavity dipole moment is to D.The form of Bo(r) is given by Eq. (l3a). From left to right, the e(l)curves are for the Debye response, then for A/a = 2/3, t, 2, 5,

and 20.

-3

'"~u -2c

0 ......------------------,

-1

(12), and the electric field potential cpo Following Frohlich[32], the requirements of continuity of the electric field Eand displacement D lead to a pair of complex equations forthe reaction field r(w). To obtain r(t), we solve the coupledcomplex equations and numerically integrate the Laplaceequation by a standard Runge-Kutta quadrature. This so­lution provides values of the potential cp obtained at theboundary of the dipole cavity. Once r(w) is in hand, anumerical inverse Fourier transform provides the time de­pendent reaction field factor r(t) Eq. (4) then provides R(t).

The ratio R(t)1R (0) should be directly compared with C(t),Eq. (1), the experimental measure of the TDFSS.

About six hours of cpu time on a Celerity 1200 computer(~1 Megaflop) are required for the calculation of R (t). Allof the results reported here are for the values of eoo = 1,eo(bulk) = 50, with eo (a) between 1 and 25, and Ala between0.66 and 20. The results of our calculations of R (t)1R (0) are

(12)

A plot of Eq, (13a) as a function of position is shown in Fig.5 for several choices Ala. Eqs. (13a) and (13b) are similarfor the values we have used for A, but Eq. (13a) shows asharper response in the saturation region than does Eq.(Bb). The final results we have obtained for the TDFSSusing Eq. (13a) are quite similar to those using Eq. (13b) foreo (r).

The Laplace equation for the electric displacement vectorD may be rewritten in terms of the dielectric response, Eq.

Continuous [27, 28] and shell models for eo (r) [29] havebeen used to describe saturation effects on solvation energiesof ions in polar solvents.

Our approach is to take an empirical function for eo (r).This eo (r) is continuous and has a shape similar to thatobtained by Debye [9]. Our empirical functions contain twoparameters. One parameter is the value of the dielectric con­stant at the cavity radius a, eo (a). The possible values foreo (a) range between e"" and eo (bulk), the optical and staticdielectric constants for the pure solvent, respectively. Theother parameter is a scaling factor A describing the lengthof the region over which the position dependent dielectricconstant eo (r) changes between its limiting values eo (a) andeo (bulk). The two empirical functions we have used are

A form for the radially and frequency dependent dielectricresponse is now required. It is asumed that the reducedscreening of the cavity dipole field affects only the staticdielectric constant and that the dielectric frequency responseis described by Eqn. (5). We then write

eo (r) - eooe(r,w) = e,,,, + . .

1 + IWfD

1.50.5o-4

'"~u -2c

-3

o ......-----------------,

-I

305 10 15 20 25LQngth (cavity radius a)

5

.....g 100::

20

25

Fig. 5The real part of the distance dependent dielectric constant Bo (r) vs.reduced length, Eqns. (l3a) and (l3b). Bo(bulk) = 50, Bo(a) = 2.0,f oo = 1.5. From left to right the curves are for Eqns. (l3b), Ala =2; (13a), A/a = 2; (l3b), A/a = 5; (l3a), A/a = 5; (l3b), A/a = 10;

and (l3a), A/a = 10

Fig. 7Calculated TDFSS response vs. reduced time, using Eq. (t3b) forBo(r). Bo(bulk) = 50, f oo = t, A. = 2. From the left, the C(l) curvesare for the Debye response, then for Bo(a) = 25, to, 5, 2, and t,

respectively.

Page 6: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

368 E. W. Castner et al.: The Dynamics of Polar Solvation

shown in Figs. 6 and 7. Here we plot the natural log of R (t)/R (0) versus time t. When the calculation is made assuminga Debye type dielectric response, i.e. eo (a) = eo (bulk) andA/a large, we recover the results of the simpler continuumtheories for C (r) [3 - 5]. In this case the TDFSS is predictedto decay exponentially with time constant 'L.

Fig. 6 shows a study of the decay of the TDFSS as thesaturation length scale A/a is varied. As Ala is increased fromtwo-thirds of a cavity radius to greater than 20 cavity radii,the calculated TDFSS decay becomes progressively longerthan 'L' The decay is non-exponential for all A values butappears to become more nearly exponential as A becomeslarge especially for higher values of eo (a). However in thesecases the decay is still significantly longer than 'L'

Fig. 7 shows a series of calculated C (t) decays for a fixedvalue of A/a= 2 cavity radii, with eo (a) varied. For eo (a) =0.5 eo (bulk), for example, C (t) is nearly exponential but witha time constant about twice 'L' As eo (a) is decreased towardsunity, the value of eoo the decays become very nonexponen­tial, with average time constants more than an order ofmagnitude longer than 'L'

IV. Experimental Results

The extraction of an accurate Stokes shift correlationfunction, C (r) is a rather intricate procedure and is describedin detail in Refs. [6] and [8]. We have obtained data in aseries of protic and aprotic solvents with the time scale forC(t) varying from 0.4 ps (acetonitrile, 300 K) to 1.5 ns (pro­panol, 221 K) for four different probe molecules, LD-750[6], Coumarin 153 [8], Cl-naphthylamine and Nile Red. Ofthese four molecules the first three appear to be relativelystraightforward "single state" system where the spectrumsimply shifts and undergoes fairly small changes in shapeand width. The type of analysis described earlier should beapplicable to these systems. Nile Red however, despite everyappearance of normality in the steady state absorption andfluorescence spectra, is a "two state" system. A fairly sharpband at 610 nm evolves into a broader red shifted band ona time scale of 25 ps. Thus Nile Red is not suitable for thesimple analysis used for the other three molecules.

Fig. 8 shows a comparison of C (t) curves obtained usingCoumarin 153 (Cu 153) and o-naphthylamine in ethanol at253 K. The response is very similar for the first 3 1/2 life-

,......., -2+-

u -3'---'

c: -4

-5

-60 200 400 600 800 1000

Itme Ips)

Fig. 8Plots of In [C (t)] vs. t are given for the two probe moleculesl-amino-naphthalene (I-AN) and Coumarin 153 (CuI53) in ethanol

at T = 253 K

times. A series of such comparisons in different solvents in­dicates that within the limits of our accuracy the C (t) curvesare independent of the probe molecule. The curves are al­ways qualitatively similar (see below) and the maximum de­viation of the average decay time is ::::: 30%.

In the previous section we discussed three differentsources of non-exponentiality in solvation dynamics. At thepresent stage quantitative comparison with experiment ­with the exception noted below - have not been attemptedand our discussion will focus on qualititative aspects. Theexception is the work Nagarajan et al. [35] on solvation isthe highly associated liquid glycerol triacetate. The dielectricdata for this solvent fit well to both Davidson-Cole and asum of three Debye times [36], and both forms give a C(t)which fits the data fairly well. Although the time resolutionof this study was limited the authors concluded that theexperimental C(t)'s decayed more rapidly at short times thanthe calculated curves implying that the influence of higherfrequency components is underestimated in both the D-Cand triple exponential models.

o - Obs.

Cole.-1

....,...,

.j..l'-' -2u....C

-.J

-3

-40 1000 2000 3000 4000 .

Time (ps)

Fig. 9Comparison of observed ( x - x - x ) and calculated (- - - ) C(t)correlation functions in n-propanol. The calculated curves wereobtained using the continuum theory expressions Ref. [4] and ap­proximating the n-propanol dielectric response by a two-Debyeform. The dielectric parameters used in these calculations werefor253, 233, and 223 K respectively: Gol = 27.81, 31.43, and 33.80;Bxt = 3.94, 4.10, and 4.20; C~r2 = 3.10,3.14. and 3.16; EDt = 2140,

. 5190, and 9600 ps, TD2 = 12.3, 19.7, and 28.3 ps.

Fig. 9 compares three C(t) curves for Cu153 in n-pro­panol. The calculated curves used a two-Debye form follow­ing Ref. [4]. Comparing the calculated and experimentalcurves it is clear that the measured C(t) curves are non­exponential in a way quite different from the calculaledcurves. The calculated curves are double exponential andbecome exponential at long times. The presence of high fre­quency components in the propanol response - whichseems to be well fit as the sum of 3 Debye regions over thewhole temperature range of Fig. 7 - complicates the com­parison at short times. However, inclusion of the secondDebye region (monomer dynamics) reproduces the rapid in­itial decay rather well. The deviation at long times does notseem likely to be explainable in terms of non-sphericity of

Page 7: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

E. W. Castner et al.: The Dynamics of Polar Solvation 369

15

Fig. 10Summary of observed solvation times compared with the solventlongitudinal relaxation time. Plots are of robs/ rL vs. log (Eo) for thesolutes, LDS 750 (open rectangles), Coumarin 153 (open circles),I-aminoaphthalene (filled triangles), DMAPS (filled diamonds). The

DMAPS data was provided by John Simon

Qualitatively the results in Fig. 10 are quite consistentwith the calculations of Fig. 7. As the value of eo (a) decreasesfrom eo the response of the solvent slows down substantially.We thus believe that the data in Fig. 10 represent a directdemonstration of a saturation effect or radially dependentresponse in solution. Clearly much more work is requiredto investigate this fascinating effect quantitatively.

References

[I] D. F. Calef and P. G. Wolynes, J. Phys. Chern. 87. 3387 (1983).[2] M. Sparpaglione and S. Mukamel, J. Phys. Chern. 91. 3938

(1987).[3] Y. T. Mazurenko and N. G. Bakshiev, Opt. Spectrosc. 28. 490

(1970).[4] B. Bagchi, D. W. Oxtoby, and G. R. Fleming, Chern. Phys.

86. 257 (1984).[5] G. Van der Zwan and J. T. Hynes, J. Phys. Chern. 89. 4181

(1985).[6] E. W. Castner, Jr., M. Marconcelli, and G. R. Fleming, 1. Phys.

Chern. 86. 1090 (1987).[7] M. Maroncelli and G. R. Fleming, J. Phys. Chern. 86. 6221

(1987).[8] M. Maroncelli, E. W. Castner, Jr., S. P. Webb, and G. R.

Fleming, in Ultrafast Phenomena V, p. 303.[9] P. Debye, Polar Molecules Chemical Catalog Co., Chapter

VI, p. 109. New York 1929.[10] M. C. Chang, S. H. Courtney, A. J. Cross, R. J. Gulotty, J.

W. Petrich, and G. R. Fleming, Anal. Instrum. 14.433 (1985).[11] J. Shah, T. C. Damen, and B. Deveaud, Appl. Phys. Lett. 50,

1307 (1987).[12] D. Block, J. Shah, and A. C. Gossard, Sol. State Comm. 59,

527 (1986).[13] K. Kash and J. Shah, Appl. Phys. Lett. 45. 401 (1984).[14] M. D. Dawson, T. F. Boggess, and A. L. Smirl, Opt. Lett. 12.

254 (1987).[15] M. D. Dawson, T. F. Boggess, D. W. Garvey, and A. L. Smirl

in Ultrafast Phenomena V, eds. G. R. Fleming, and A. E.Siegman, p. 5, Springer-Verlag, Berlin 1986.

[16] M. D. Dawson, T. F. Boggess. D. W. Garvey, and A. L. Smirl,Opt. Comm. 60, 79 (1986).

[17] D. W. Davidson and R. H. Cole, J. Phys. Chern. 19. 1984(1951).

[18] N. E. Hill, W. E. Vaughan, A. H. Price, and M. Davies, Di­electric Properties and Molecular Behavior. Van NostrandReinhold Company 19B9.

[19] C. 1. F. Bottcher and P. Bordewijk, Theory of Electric Polar­ization Volume II, Elsevier, 1978.

[20] I. Rips and J. Jortner, Chern. Phys. Lett. 133.411 (1987).[21] E. W. Castner, Jr., G. R. Fleming, and B. Bagchi, submitted

to Chern. Phys. Lett.[22] C. H. Brito-Cruz, R. L. Fork, W. H. Knox, and C. V. Shank,

Chern. Phys. Lett. 132.341 (1986).[23] R. F. Loring and S. Mukamel, 1. Phys. Chern. 91.1302 (1987).[24] M. F. Shlesinger and E. W. Montroll, Proc. Natl. Acad. Sci.,

USA 81, 1280 (1984).[25] J. Klafter and M. F. Shlesinger, Proc. Natl. Acad. Sci, USA

83,848 (1986).[26] R. F. Loring and S. Mukamel, J. Phys. Chern. 87.1272 (1987).[27] A. A. Kornyshev, in The Chemical Physics of Solvation. Part

A. Theory of Solvation. eds. R. R. Dogonadze, E. Kalman, A.A., Kornyshev, and J. Ulstrny Chap 3 p. 77, Elsevier, Am­sterdam 1985.

[28] A. A. Kornyshev and A. G. Volkov, J. Electroanal. Chern. 180.363 (1984).

Concluding Remarks

In this paper we have described three extensions of con­ventional continuum theories of solvation dynamics, andtheir relevance to current experimental data. The nonex­ponential solvent responses, predicted and observed haveimportant implications for electron (and proton) transferreactions in solution. It will be of interest to try to observemanifestations of these phenomena in, for example, adi­abatic electron transfer reactions.

We thank Professors John Simon and Paul Barbara for providingus with preprints. This work was supported by a grant from theNational Science Foundation.

o

2.5

oo

oo

1.5

•cro.~~-~~-------------

5

o

.... .J10"- II

.Do

....

the probe molecule but could result, in principle, from eithera distribution of dielectric relaxation times or a positiondependent dielectric response. Judging from the data of Coleand Davidson [37] the response of n-propanol in the lowfrequency regime is well described by a Debye form at allof the experimental temperatures. It thus seems unlikely thatthe Davidson-Cole f3 could be much less than 0.9. Compar­ing Figs. 4 and 9 leads us to suggest that the long timebehavior in the experimental data does not result fromaD. C. distribution in this solvent. Rather we suggest thatthe non-exponentiality arising from a radial dependence ofdielectric constant (Figs. 6 and 7) is responsible for curvatureof C(t) in Fig. 9. Attempts to make a quantitative compar­ison with the "saturation effect" model are currently under­way. Fig. 10 shows the dependence of Tobs/'tL on eo for allthe systems we have studied plus several from the literature[33, 34]. For TL only the longest longitudinal time is usedTL = TLI = (e:<;l/eod TOI' The dashed line represents thecontinuum theory prediction of Refs. [4] and [5]. Clearlyfor eo > 50 this model is failing very badly. Several pointsabout Fig. 10 are worth noting. First, in the region whereTobs - TL' Tobs varies by a factor of 3 . 103

. Second, the valuesof Tobs for the high dielectric constants solvents, propylenecarbonate and n-methyl propionamide, lie in the middle ofthe range of Tobs values. Further, the spectral shifts found inthese solvents fit with the other solvents very well. Thus thestabilization in the high dielectric constant solvent conformsto expectations, but it occurs much more slowly than thesimple continuum theory predicts [3 - 5].

Page 8: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

370 E. W. Castner et al.: The Dynamics of Polar Solvation

[29] E. Ehrenson, J. Compo Chem. 2.41 (1981).[30] V. Friedrich and D. Kivelson, J. Phys. Chem. 86. 6425 (1987).[31] E. W. Castner, Jr., G. R. Fleming, and B. Bagchi, J. Chem.

Phys., in preparation.[32] H. Frohlich, Theory of Dielectrics. Oxford 1958.[33] S.-G. Su and J. D. Simon, J. Phys. Chem. 90.6475 (1986).[34] S.-G. Su and 1. D. Simon, J. Phys. Chem. 91. 1055 (1971).[35] V. Nagarajan, A. M. Brearley, T.-J. Kang, and P. F. Barbara,

J. Chem. Phys. 86.3183 (1987).[36] A. M. Ras and P. Bordewijk, Rec.Trav. Chim. 90.1055 (1971).[37] R. H. Cole and D. W. Davidson, J. Chem. Phys. 20. 1389

(1952).

summarizes our paramaterization of the temperature dependentdielectric properties of five alcohols: methanol, ethanol, n-propanol,n-butanol, and 2-propanol, which we hope will be of similar valueto other workers in the field.

Alcohols are widely used in time resolved studies of solvationbecause their response times are slow enough to beobservable with- 50 ps time resolution. Further, the shorter alcohols persist asliquids to temperatures well below room temperature where theirresponse times become much slower. For the same reasons a largebody of dielectric data exists for these solvents. The dielectric re­sponse of the alcohols considered here can be discribed by a sumof three Debye dispersion regimes as [1]:

The lowest frequency dispersion (j = I) accounts for the majorityof the dielectric response (for example, in n-propnaol (1:01 - 1:001)/(1:01 - 1:00 3) = 12 at 298 K) and is ascribed to reorientations ofalcohol clusters [1]. It is this dispersion for which the majority ofdielectric data exists, and, except for the case of n-propanol, wewillconsider regime 1 alone.

References to the original literature were obtained from severalsources. Most came from a search of Chemical Abstracts for theyears 1967-1985. Earlier references were found in the reviews byBottcher and Bordewijk [2], Buckley and Maryott [3], Jakusek

Appendix

Of special interest is the connection between solvent bulk dielec­tric properties, the dynamics of solvation of molecular probes, andthe effects polar solvation has in determining reaction rates. Tocompare the measured solution dynamics to behavior predicted bytheory, one often needs to known the bulk dielectric response (I:(w)of chosen solvents as a function of a convenient experimental var­iable such as temperature. While such data exists, it occurs scatteredamong many sources and it is a tedious task to find the particularvalues of interest and assess their uncertainties. During the courceof the studies presented in the body of this papar, we have collectedmany of the original results on the dielectric dispersion I:(w) of shortchain alcohols. In cases where sufficient temperature dependentdata exists, it is possible to fit the dielectric dispersion propertiesas a function of temperature, and we have found use of such fittedcurves an extremely convenient tool in our work. The present note

]

I:(w) = 1:00 3 Ij~1

with

I:Oj -I: L j

t - i UH j

(I)

Table 1Summary of fits to temperature-dependent dielectric data

Temperature" Parameter!"range (K)

Solvent

Methanol

Ethanol

n-Propanol

n-Butanol

2-Propanol

163-328

131- 323

117-333

135- 323

172-323

£0,

ex 1

f,

"0,t;>:.1

f,

"0,e~, = "0,f,

f,ey.2 = "ojf,

£x3

fJ

"0,

£<.;t'~ I

r,

"0,£-~_I

f,

Fitting coefficients"a" a, a,

-26.66 17.6870.51 1.491

-1.209 0.907 -0.0120

-23.09 14.795 -0.19432.01 0.720

-0.9213 0.94165

-21.19 13.302 -0.25252.07 0.4700.246 0.372 0.10032.64 0.1162.839 - 1.2158 0.22401.723 0.13951.065 -0.7853 0.1687

- 21.08 12.006 -0.17592.22 0.361

-0.153 0.4687 0.1151

-18.9 11.442.03 0.423

- 1.956 1.374

iTO

54 0.14 0.270.10 0.26

45 4.5' to-' Om5

24 0.039 0.1323 0.14 0.2917 5.7·10 3 5.7· 10- 3

28' 0.69 0.4614 0.046 0.1524 0.013 0.0868 0.019 0.10

to 0.010 0.0747 om 1 0.0737 1.4· 10- 3 0.020

30 0.18 0.3426 0.098 0.2633 8.4· to 3 0.071

18 3.0 1.515 0.088 0.2326 0.016 0.097

References"

[6,7.8][9,10,11]

[12, 13, 14][11,15.16,17]

[18, 19,20][1,21]

[I, 15,22][23,24,25]

[16. 11,22][26.27]

a, Temperature range of data used in the fits.b, Parameters of the multiple-Debye description of e(w) as defined in Eq. (1). The parameters were fit to a power series in t0 3

/ T as given by Eq. (2). Relaxationtimes were fit with p (Eq. 2) equal to 10gIO (f J) with f j expressed in picoseconds.

c See Eq. (2).d, Number of data points used in the fit.e Reduced chi-squared value for the fit defined by

x; = L [p, - p(Ti )]2/V

where p, is the observed parameter value at temperature T" and p(T,) is its calculated value, and v is the number of degrees of freedom of the fit.n Average absolute deviation defined by

iT = L IPi - p(T,)I.,., The symbols 1, 2, 3 etc. on the figures refer to the references in the Table in the order in which they are listed there. For example, 1 on Fig. 1 corresponds

to Ref. [8] in TabA, on Fig. 4 corresponds to Ref. [23].

Page 9: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

E. W. Castner et al.: The Dynamics of Polar Solvation 371

MethanolBO

7070

60 500 0

W 50 W

.0 30

3010

10 6

5B

B BW W •

63

• 2

3.5 10

"B

"0 0

l-' t 6...., 2. 5 enm 0 •0 ..J....J 2

1.5 03 • 5 6 3 5 7 9

103/T(K) 10

3/T(K)

Fig. 11Temperature dependent dielectric parameters (Eq, 1) for the lowestfrequency dielectric dispersion regime (j = 1)of methanol. The sym­bols 1, 2, 3 etc. denote the original source of the data as describedin the notes to the Table. The curves were generated from the fits

listed in the Table

Fig. 13Temperature dependent dielectric dispersion parameters ofn-propanol. Data for all three dispersion regimes are representedhere. In panel (a) eo is f~ll and in panels (b) and (c) the tree curvesto correspond to f..'XCj and !j withj = 1,2, and 3 from top to bottom.See caption to Fig. 11 and the Table for the key to the data sources

Ethanol n-ButanolBO ISO

50eo

0 o .0W W.0 30

2020

B5

7

B 6 8W W •

• 310

IS

" 5 B0 "t 0

en • t 60 en

..J 03 ..J •2

7 B2

3 • 5 IS 3 5 IS 710 31T

(K)•

1 031T (K)

Fig. 12Temperature dependent dielectric parameters for the lowest fre­quency dielectric dispersion regime of ethanol. See caption to

Fig. 11 and the Table for the key to the data sources

Fig. 14Temperature dependent dielectric parameters for the lowest fre­quency dielectric dispersion regime of n-butanol. See caption to

Fig. 11 and the Table for the key to the data sources

Page 10: The Dynamics of Polar Solvation Castner...364 E. W. Castner et al.: The Dynamics of Polar Solvation L5 Nd YAG Laser Ff 1.00mm urea L3!L4 L1 S L2 c I.O-I.5W 532om Fig. 1 Synchronously

Fig. 15Temperature dependent dielectric parameters for the lowest fre­quency dielectric dispersion regime of 2-propanol. The asterisks in(b) denote data points not included in the fit. See caption Fig. 11

and the Table for the key to the data sources

The results of such fitting are tabulated in the Table and plottedin Figs. 11 through 15. For the dielectric constant parameters inEq. (2) P is just f.o" c , I etc. The relaxation times 'j were fit as p =IOglO(',) with r expressed in picoseconds. In all cases a quadraticlit was sufficient to reproduce the data over the available temper­ature range (see figures). Base on the deviations of the data fromthe fitted curves the accuracy of the dielectric parameters computed

Appendix References

[1] S. K. Garg and C. P. Smyth, 1. Phys. Chern. 69. 1294 (1965).[2] C. J. F. Bottcher and P. Bordewijk, "Theory of Electric Po­

larization", Vol. II, (Elsevier, p. 106, New York 1978.[3] F. Buckley and A. A. Maryott, "Tables of Dielectric Disper­

sion Data for Pure Liquids and Dilute Solutions", NBS Cir­cular # 589 (1958).

[4] E. Jakusek and L. Sobyzyk in "Dielectric and Rolated Mo­lecular Processes" Vol. 3 p. 108, The Chemical Society, Lon­don 1977.

[5] C. Brot, M. Magat, and L. Reinsch, Kolloid Z. 134.101 (1953).[6] G. H. Barbenja, J. Chim. Phys. 65. 906 (1968).[7] D. J. Denney and R. H. Cole, J. Chern. Phys. 23. 1767 (1955).[8] D. W. Davidson, Can. J. Chern. 35. 458 (1957).[9] B. P. Jordan, R. J. Sheppard, and S. Szwarnowski, 1. Phys.

D: Appl. Phys. 11.695 (1978).[10] N. V. Chekalin, V. D. Paponov, and M. I. Shakharonov, Zh.

Strukt. Khim. 11.815 (1970).[11] D. Bertolini, M. Cassettari, and G. Salvetti, J. Chern. Phys.

78.365 (1983).[12] E. H. Grant, Proc. Phys. Soc. (London) B 70,937 (1957).[13] A. Suggett, P. A. Mackness, and M. J. Tait, Nature 228. 456

(1970).[14] H. Nakamura, S. Mashimo, and A. Wada, Jap. J. Appl. Phys.

21.1022 (1982).[15] M. W. Sagal, J. Chern. Phys. 36. 2437 (1962).[16] F. X. Hassion and R. H. Cole, J. Chern. Phys. 23.1756 (1955).[17] J. A. Lane and J. A. Saxon, Proc. Roy. Soc. (London) A213.

400 (1952) as cited in Ref. [5].[18] D. W. Davidson and R. H. Cole, J. Chern. Phys. 19. 1484

(1951).[19] R. H. Cole and D. W. Davidson, J. Chern. Phys. 20. 1389

(1952).[20] S. Mizushima and T. Aso, J. Chern. Soc. 49. 153 (1928) as

cited in Ref. [5].[21] I. Dutuit, S. L. Salefran, and A. M. Bottreau, Adv. Mol. Relax.

Interact. Processess 23. 75 (1982).[22] L. M. Imanov and Ya. M. Abbasor, Zh. Fiz. Khim. 37. 1510

(1963).[23] L. Dannhauser and R. H. Cole, J. Chern. Phys. 23. 1762(1955).[24] H. A. Rizk and N. Youssef, Z. Phys. Chern. 58. 100 (1968).[25] (a) L. Reinsch, J. Chim. Phys. 51, 113 (1954); (b) 1. F. van den

Berg, J. C. F. Michielsen, and J. A. A. Ketelaar, Proc. Kon.Ned. Wetensch. Ser B 76. 189 (1973); (c) A. Suggest, P. A.Mackness, and M. J. Tait, Nature 228. 456 (1970).

[26] H. A Rizk and I. M. Elanwar, Z. Phys. Chern. 62. 225 (1968).[27] S. Mizushima, Sci. Papers Inst. Phys. Chern. Research(Tokoy) 5. 201 as cited in Ref. [5].

Presented at the Discussion Meeting of the E 6692Deutsche Bunsen-Gesellschaft fur Physi-kalische Chemic "Intramolecular Proc-esses", in Grainau, August 17th to 20th,1987

as above is better then ± 1% for Go\ (except for 2-propanol whichis ±5%), roughly ±5% for 6\ (62, 6d)' and better than ±10% for'\'2. '3)'

(2)

6

E. W. Castner et al.: The Dynamics of Polar Solvation

3

3

6

~ S

~~ 4o

...J

and Sobczyk [4], and Brot et al. [5]. Actual references to data usedhere are provided in Table 1. In choosing which data to use in theparameterizations we did not attempt to include all possible results.For the most part only sets of data that covered some range oftemperatures and several od the dielectric quantities in Eq. (1) wereutilized. Thus the substantial number of reports of static dielectricconstant measurements that did not include frequency dependentdata were not included.

A convenient way to summarize the 6(W) data as a functionof temperature is to lit the parameters of Eq. (I) as a power in 1000/T (T in degrees Kelvin) as:

372

so

40

0W 30

20

7 s'(b)6

s'8 S

W4

3