1: \documentstyle[aps,preprint]{revtex}
2:
3: \tightenlines
4: %\documentstyle[aps]{revtex}
5:
6: \begin{document}
7:
8: \draft % To print out PACS
9:
10: %%%
11: %\twocolumn
12: %%%[{
13: %%%
14: \widetext
15:
16:
17: \title{Excited States of Linear Polyenes}
18:
19:
20: \author{William Barford$^{1\ast}$, Robert J. Bursill$^2$ and Mikhail Yu
21: Lavrentiev$^{1\ast\ast}$}
22:
23:
24: \address{
25: $^1$Department of Physics and Astronomy, The University of Sheffield,
26: \\
27: Sheffield, S3 7RH, United Kingdom.
28: \\
29: $^2$School of Physics, University of New South Wales, Sydney, NSW
30: 2052,
31: Australia.
32: }
33:
34: %%
35: \maketitle
36:
37: %%%
38: \mediumtext
39:
40: \begin{abstract}
41: We present density matrix renormalisation group calculations of the
42: Pariser-Parr-Pople-Peierls
43: model of linear polyenes within the adiabatic approximation.
44: We calculate the vertical and relaxed transition energies, and relaxed
45: geometries for various excitations on long
46: chains. The triplet ($1^3B_u^+$) and even-parity singlet ($2^1A_g^+$)
47: states have a
48: 2-soliton and 4-soliton form, respectively, both with large relaxation
49: energies. The dipole-allowed ($1^1B_u^-$) state forms an
50: exciton-polaron and
51: has a very small relaxation energy. The relaxed energy of the
52: $2^1A_g^+$ state lies below that of the $1^1B_u^-$ state.
53: We observe an attraction between the soliton-antisoliton
54: pairs in the $2^1A_g^+$
55: state.
56: The calculated excitation energies agree well with the
57: observed values for polyene oligomers; the agreement with polyacetylene
58: thin
59: films is less good, and we comment on the possible sources of
60: the discrepencies.
61: The photoinduced absorption is interpreted. The
62: spin-spin
63: correlation function shows that the unpaired spins coincide
64: with the geometrical soliton positions. We study the roles
65: of electron-electron
66: interactions and electron-lattice coupling in determining the excitation
67: energies and soliton structures. The electronic
68: interactions play the key role in determining the ground state dimerisation
69: and the excited state transition energies.
70: \end{abstract}
71:
72:
73: \pacs{PACS numbers: 71.10.F, 71.20.R, 71.35}
74:
75: %%% }]
76: \narrowtext
77:
78: \section{Introduction}
79:
80: The inter-play of electron-electron interactions and
81: electron-lattice
82: coupling in linear polyenes
83: results in
84: a wealth of low-lying excitations.
85: Electron-electron interactions
86: induce spin density wave correlations in the ground
87: state. The lowest lying excitations are
88: triplets, which combine to form dipole-forbidden
89: singlet
90: ($^1A_g^+$) excitations. Optical excitations are
91: gapped, lie above the $2^1A_g^+$ state, and are essentially
92: ionic
93: in
94: character, that is, there is charge transfer from one site to another.
95: The lowest optically allowed ($1^1B_u^-$) state lies below
96: the charge gap \cite{footnoteI}, and is thus excitonic in character. For
97: convenience, we show the
98: group
99: theoretic labelling of the states discussed in this paper in
100: Table I.
101:
102: Electron-phonon interactions result in a
103: dimerised semiconducting ground state.
104: Within the adiabatic approximation, the
105: non-linear excitations include charged-spinless and neutral-spin 1/2
106: solitons. Both
107: electronic
108: interactions and electron-lattice coupling lead to
109: a gap in the optical spectrum.
110: In contrast to the interacting limit, however, the
111: $2^1A_g^+$ state always lies above the $1^1B_u^-$ state in the
112: non-interacting electron-phonon model.
113:
114: The realisation that electronic interactions play a significant role in
115: polyenes
116: came via
117: the experimental observation,
118: by Hudson and Kohler\cite{hudson72} in 1972, that the $2^1A_g^+$ state lies
119: below the
120: $1^1B_u^-$
121: state. At the same time, by
122: perfoming
123: a double configuration interaction calculation on the
124: Pariser-Parr-Pople
125: model,
126: Schulten and Karplus\cite{schulten72} demonstrated that the $2^1A_g^+$
127: wavefunction has a strong triplet-triplet contribution, and has a lower
128: energy
129: than the
130: $1^1B_u^-$ state. The triplet-triplet and correlated
131: nature of the $2^1A_g^+$ state has
132: been further
133: investigated by Tavan and Schulten\cite{tavan87} and other workers\cite{refs}.
134: In 1986, Hayden and
135: Mele\cite{hayden86} performed a real space renormalisation group
136: calculation on the
137: Hubbard-Peierls model of up to sixteen sites and found that the $2^1A_g^+$
138: state was
139: composed of 4-solitons. This 4-soliton nature has
140: also been investigated by Su
141: \cite{su95},
142: and Wen and Su\cite{wen97}.
143: Ovchinnikov {\em et al.} also high-lighted the role of electronic
144: interactions,
145: by suggesting that they are largely responsible for the optical gap
146: \cite{ovchinnikov}.
147: In contrast to the strong deviations from the ground state geometry
148: predicted
149: for the
150: triplet and $2^1A_g^+$ state, Grabowski {\em et al.}\cite{grabowski85}
151: predicted
152: that the $1^1B_u^-$ state is an exciton-polaron.
153:
154: The existence of the $2^1A_g^+$ state below the $1^1B_u^-$ state in
155: polyacetylene
156: thin films has been suggested by a number of experiments. Third harmonic
157: generation
158: (THG) and two photon absorption by Halvorson and
159: co-workers\cite{halvorson93}
160: indicate that a $^1A_g^+$ state lies below 1.1 eV; while
161: the linear absorption,
162: locating
163: the
164: $1^1B_u^-$ state, typically rises at 1.8 eV and peaks at
165: 2.0
166: eV \cite{vardeny}. However, Fann
167: and co-workers\cite{fann89} performed THG, finding peaks at 0.6 eV and
168: 0.89 eV,
169: which they interpret as $^1A_g^+$ and $^1B_u^-$ states virtually
170: coincident at 1.8
171: eV.
172: The position of the $2^1A_g^+$ state is therefore
173: not definitively established.
174: We return to this point in section V when we discuss our own theoretical
175: predictions.
176: For a detailed review
177: of the experimental and theoretical studies of conjugated polymers up to 1992,
178: see \cite{kiess}.
179:
180: Electron-electron interactions in
181: $\pi$-conjugated systems, such as {\em trans}-polyacetylene,
182: are conveniently
183: modelled
184: by the
185: one-band
186: Pariser-Parr-Pople model, which includes long range Coulomb interactions.
187: This semi-empirical model has been extensively used to study the excited
188: states of small conjugated molecules with a remarkable degree of success
189: \cite{bursill98}.
190: The Peierls model describes the electron-lattice coupling in the adiabatic
191: limit.
192: Thus, the Pariser-Parr-Pople-Peierls model is a realistic and accurate model
193: of $\pi$-conjugated systems, which captures their essential physics.
194: In an earlier paper\cite{bursill99} we performed accurate calculations
195: on this model using
196: the infinite lattice
197: algorithm of
198: the
199: density matrix renormalisation group (DMRG)
200: method \cite{White}, \cite{book}.
201: The
202: Hellmann-Feynman
203: theorem
204: was used to calculate the low-lying excited states and the lattice
205: geometry associated with them. We showed that the $1^3B_u^+$ and
206: $2^1A_g^+$
207: states are modelled by 2 and 4 soliton fits, respectively, and that the
208: $1^1B_u^-$ state
209: is an exciton-polaron. In this paper we develop that work. In particular,
210: our objectives are:
211:
212: \begin{enumerate}
213:
214: \item
215: To further demonstrate that the DMRG calculations are reliable by, (i)
216: making
217: comparisons to the exact non-interacting limit, and (ii) comparing the
218: infinite
219: lattice method to the finite lattice method.
220:
221: \item
222: Use a realistic model of polyenes to understand the roles
223: of electron-electron
224: interactions and electron-lattice coupling in determining the
225: dimerisation of the ground state and the transition energies of the excited
226: states. In agreement with the earlier work of
227: Horsch\cite{horsch81}, and Konig and Stollhoff \cite{konig90}, we
228: find that the electronic interactions play the key role in driving the ground
229: state
230: dimerisation. Electronic interactions are also dominant
231: in determining the solitonic structures and transition energies of the
232: excited states.
233:
234: \item
235: To make more detailed comparisons to other experimental
236: probes, in particular
237: photo-induced absorption.
238: The agreement with a wide range of experiments
239: confirms the validity of the model, our calculational method,
240: and our predictions on the soliton structures and their interactions.
241:
242: \item
243: To further investigate both the geometry and electronic properties of solitons.
244:
245: \end{enumerate}
246:
247:
248: This paper also serves as a correction to \cite{bursill99}. In that paper
249: we used the dimerised ground state geometry in the Coulomb interactions
250: to calculate the energy of all
251: the states. Thus, the Coulomb interactions (unlike the one-electron transfer
252: integrals) were not updated in the Hellmann-Feynman minimisation procedure for
253: the relaxed states. We find that using the correct geometry in the Coulomb
254: interactions affects the excitation energies by ca.\ $0.1$ eV.
255: The geometry of the triplet excited state is modified, so that now there is no
256: soliton-antisoliton confinement in the triplet state. However, attractive
257: soliton interactions remain in the $2^1A_g^+$ state.
258:
259: The plan of this paper is as follows. In the next section we introduce
260: the Pariser-Parr-Pople-Peierls model. To establish the consequences
261: of the inter-play of electron-electron interactions and
262: electron-lattice
263: coupling, we consider these two limits separately in sections
264: III and IV. The
265: non-interacting
266: limit also allows us to compare the infinite and finite DMRG algorithms
267: to an exact
268: calculation.
269: In section V we solve the full model, and discuss
270: the vertical and relaxed energies
271: of the key excited states. As well as
272: linear
273: absorption and non-linear optical spectroscopies, photo-induced
274: absorption is a useful tool in determining the positions of excited
275: states.
276: We discuss the experimental situation and our theoretical interpretation
277: in section VI. In section VII we consider the solitonic structures. By
278: making comparisons between the
279: geometrical soliton structures and the spin-spin correlation functions,
280: we show how they are closely related.
281: We conclude and discuss in section VIII.
282:
283: As well as the work already mentioned, earlier work on the solitonic
284: structure of the
285: low-lying excitations include, a mean-field study of the
286: Heisenberg-Peierls
287: model
288: \cite{takimoto89} and an exact diagonalisation of a 12 site extended
289: Hubbard-Peierls
290: model \cite{gammel93}. The DMRG method has recently been used by
291: Yaron et al.\ \cite{yaron98} and Fano et al.\ \cite{fano98} to solve the
292: Pariser-Parr-Pople
293: model for linear and cyclic polyenes, respectively.
294:
295:
296: \section{The Pariser-Parr-Pople-Peierls Model}
297:
298: The Pariser-Parr-Pople-Peierls model is a realistic and accurate model
299: of $\pi$-conjugated systems, which includes the key features
300: of long range electron-electron interactions and electron-lattice
301: coupling. The Hamiltonian for an $N$ site chain with open
302: boundary conditions is defined as
303: %
304: \begin{eqnarray}
305: %
306: {\cal H}
307: %
308: & &
309: %
310: = - 2\sum_{\ell=1}^{N-1} t_{\ell} \hat{T}_{\ell}
311: + \frac{1}{4 \pi t_0 \lambda} \sum_{\ell=1}^{N-1} \Delta_{\ell}^2 +
312: \Gamma \sum_{\ell=1}^{N-1} \Delta_{\ell} \nonumber \\
313: %
314: & &
315: %
316: +\;
317: U \sum_{i=1}^N \left(n_{i\uparrow}- \frac{1} {2} \right)
318: \left(n_{i\downarrow} - \frac{1} {2} \right)
319: + \sum_{<ij>} V_{ij} (n_i - 1)(n_j - 1),
320: %
321: \end{eqnarray}
322: %
323: where, $<ij>$ indicates all pairs of sites,
324: %
325: $t_{\ell} = \left( t_0 + \frac{\Delta_{\ell}}{2} \right) $ and
326: \begin{eqnarray}
327: \hat{T}_{\ell}
328: = \frac{1} {2} \sum_{\sigma}
329: (c_{\ell+1 \sigma}^{\dagger} c_{\ell \sigma} + h.c.)
330: \end{eqnarray}
331: is the bond order operator of the $\ell$th bond. We use the Ohno function
332: for
333: the
334: Coulomb interaction:
335: \begin{eqnarray}
336: V_{ij} = U / \sqrt{ 1 + \beta r_{ij}^2 },
337: \end{eqnarray}
338: where
339: %
340: $\beta = (U/14.397)^2$
341: %
342: and bond lengths are in \AA.
343: The dimensionless
344: electron-phonon coupling constant,
345: $\lambda$, is defined by
346: \begin{eqnarray}
347: \lambda = \frac{2 \alpha^2}{\pi K t_0},
348: \end{eqnarray}
349: where $K$ is the elastic spring constant (estimated to be 46 eV
350: $\AA^{-2}$ from Raman analysis of C-C stretching modes in
351: {\em trans-}(CH)$_2$) \cite{ehrenfreund87}, and $\alpha$ relates
352: the actual distortion of the $\ell$th.\ bond from equilibrium,
353: $\delta r_{\ell}$,
354: to $\Delta_{\ell}$:
355: \begin{eqnarray}
356: \delta r_{\ell} = \Delta_{\ell} / 2 \alpha.
357: \end{eqnarray}
358: %
359: We take the undistorted chain to lie along the $x-$axis, with the
360: bonds oriented at $30^o$ to this axis. Then, for fixed bond angles,
361: the distorted chain coordinates are defined as:
362: \begin{eqnarray}
363: x_{ij} = x^0_{ij} - \frac{\sqrt{3}}{4\alpha}
364: \sum_{\ell=i}^{j-1} \Delta_{\ell},
365: \nonumber
366: \end{eqnarray}
367: \begin{eqnarray}
368: y_{ij} = y^0_{ij} - \frac{1}{4\alpha} \sum_{\ell=i}^{j-1} \Delta_{\ell}
369: (-1)^{\ell +1 },
370: \end{eqnarray}
371: where
372: \begin{eqnarray}
373: x_{ij}^0 = \frac{\sqrt{3}}{2} a_0 |j-i|
374: \nonumber
375: \end{eqnarray}
376: and
377: \begin{eqnarray}
378: y_{ij}^0
379: & &
380: = 0, {\rm if }\ |j-i|\ {\rm even} \\
381: \nonumber
382: & &
383: =\;
384: \frac{a_0}{2} (-1)^{(i+1)},\ {\rm otherwise}.
385: \nonumber
386: \end{eqnarray}
387: $a_0$ ($=1.40 \AA$) is the undistorted C-C bond length.
388:
389: The force per bond, $f_{\ell}$ is
390: \begin{eqnarray}
391: f_{\ell} =
392: -\frac{ \partial \langle {\cal H} \rangle}
393: {\partial \delta r_{\ell}}.
394: \end{eqnarray}
395: Using the Hellmann-Feynman theorem this can be re-written as,
396: \begin{eqnarray}
397: %
398: f_{\ell}
399: & &
400: = -2\alpha\left(
401: \frac {\Delta_{\ell}} {2\pi t_0\lambda} + \Gamma -
402: \langle T_{\ell} \rangle \right) \\
403: \nonumber
404: %
405: & &
406: -\;
407: \sum_{<ij>}' \frac{U \beta} {2\alpha(1+ \beta r_{ij}^2)^{3/2}}
408: \left( \frac{\sqrt{3}}{2} x_{ij}
409: + \frac{ (-1)^{(\ell+1)}}{2} y_{ij}
410: \langle (n_i - 1)(n_j - 1) \rangle \right).
411: \end{eqnarray}
412: The prime over the sum indicates that the sum runs over all pairs of sites
413: which span the $\ell$th. bond. The contribution to the bond force from the
414: Coulomb interaction is small compared to the kinetic energy term: the
415: value of the Coulomb force from the nearest neighbor
416: density-density correlator
417: is approximately one tenth of the kinetic term. Moreover, the
418: density-density correlator alternates in sign
419: and drops to less than one tenth of the nearest neighbor
420: density-density correlator,
421: so the sum over all bonds
422: is also small. Table II shows the correlator for
423: up to five nearest neighbors.
424: We therefore only include the nearest neighbor density-density
425: correlator
426: in the evaluation of the distorted geometry.
427: (However, the full distorted geometry is used in the evaluation of the Coulomb
428: interaction, Eqn.\ (3).)
429:
430: Using this approximation, and setting
431: $f_{\ell} = 0$, the self-consistent
432: equation for the equilibrium $\Delta_{\ell}$ is:
433: \begin{eqnarray}
434: \Delta_{\ell} = \left( \frac{ 2\pi\alpha t_0\lambda}
435: {\alpha -C_{\ell} t_0 \lambda} \right)
436: \left( \langle T_{\ell} \rangle -\Gamma - C_{\ell} a_0 \right),
437: \end{eqnarray}
438: where,
439: \begin{eqnarray}
440: C_{\ell} = \frac{U\beta} {2\alpha(1+ \beta(a_0 + \delta r_{\ell})^2)^{3/2}}
441: \langle (n_{\ell} - 1)(n_{\ell+1} - 1) \rangle.
442: \end{eqnarray}
443: We observe that, since the nearest neighbor density-density correlator
444: is negative, the Coulomb interactions tend to increase the bond
445: dimerisation.
446:
447: The calculations were performed for fixed chain lengths, which is
448: enforced by setting,
449: \begin{eqnarray}
450: \Gamma = \frac{1}{N-1} \sum_{\ell = 1}^{N-1} \left( \langle T_{\ell}
451: \rangle - C_{\ell} a_0 \right).
452: \end{eqnarray}
453:
454: To complete our discussion of the model we turn to its parametrisation.
455: There are three parameters in the model: $t_0$, $U$ and $\lambda$. An
456: optimal parametrisation for $t_0$ and $U$ was found in \cite{bursill98}
457: by fitting the Pariser-Parr-Pople model to the excited states of benzene.
458: Assuming that this parametrisation is transferable between all
459: $\pi$-conjugated
460: systems, we use them here, and set $t_0 = 2.539$ eV and $U = 10.06$ eV.
461: The remaining parameter, $\lambda$, is found by fitting the
462: vertical energies of the $1^1B^-_u$ and $2^1A^+_g$ states,
463: calculated from the Parsier-Parr-Pople-Peierls model,
464: to the 6-site linear polyene \cite{bursill99}. This gives
465: $\lambda = 0.115$. Finally, using $K = 46$ eV
466: $\AA^{-2}$ implies $\alpha = 4.593$ eV $\AA^{-1}$.
467:
468: \section{Solution of the Peierls Model}
469:
470: As originally recognised by Pople and Walmsley\cite{pople}, the
471: low lying excitations of the dimerised even $N$ site chain correspond to the
472: creation of two mid-gap states. These excitations are associated
473: with localised geometrical structures which lead to a reversal
474: of the lattice dimerisation, and were subsequently
475: termed solitons. The defect
476: states repel from each and are repelled from the ends of the chain.
477: Thus,
478: they reside
479: at approximately
480: $N/4$ and $3N/4$ along the chain (as may be seen in Fig.\ 5(b)).
481: Fig.\ 1 shows
482: a
483: schematic
484: energy
485: diagram of the molecular orbitals and defect states,
486: while Fig.\ 2 shows the energies of the $1^1B^-_u$ and $2^1A^+_g$ states
487: as a function of inverse chain length. It is clear
488: that the first excited even parity state lies above the odd parity
489: state. However, in the long chain, continuum limit, these states
490: are degenerate, with energy
491: $4\Delta_0/\pi = 0.12$ eV, using $\lambda = 0.115$,
492: $t_0 = 2.539$ eV and \cite{ssh}
493: \begin{eqnarray}
494: \Delta_0 = 8t_0 \exp\left[ - \left( 1 + \frac{1}{2\lambda} \right)
495: \right].
496: \end{eqnarray}
497: This gap is only a fraction of the experimentally measured gap of
498: approximately
499: 2.0
500: eV\cite{footnoteIII}. While a larger optical gap can be obtained
501: by increasing $\lambda$ and $t_0$, the energetic
502: ordering of the low lying
503: states would still be incorrect. As we see in the next section, it is
504: electronic interactions which primarily open the optical gap, and reverse
505: the energetic ordering of the states.
506: Furthermore, electronic interactions significantly modify
507: the soliton structures, as we show in section VII.
508:
509:
510: The non-interacting limit enables us to make a comparison between the
511: DMRG
512: methods and the
513: exact calculation. In Fig.\ \ref{comparison1} the energy difference
514: between the
515: exact results and DMRG calculations is shown for the $1^1B^-_u$ and
516: $2^1A^+_g$
517: states. We see that
518: for both states the accuracies of the infinite and finite lattice
519: algorithm calculations are close, so that both methods can be
520: used in the actual calculations. The accuracy
521: is better for the $1^1B^-_u$ state, but even for the $2^1A^+_g$
522: the error is about 0.002 eV for the 50 site chain in the
523: infinite lattice algorithm calculation. Other DMRG convergence
524: tests, confirming
525: the validity of the method, were
526: presented
527: in \cite{bursill99}.
528:
529:
530: \section{Solution of the Pariser-Parr-Pople model}
531:
532: The uniform chain in the limit of only on-site Coulomb interactions
533: is described by the Hubbard model. At half-filling, the spin
534: excitations are gapless in the infinite chain limit, whereas the charge
535: excitations are gapped.
536: Even though the Pariser-Parr-Pople model contains long range interactions,
537: the spin excitations still appear to be gapless in the uniform chain,
538: as shown in
539: Fig.\ 2.
540: The $2^1A^+_g$ state is also gapless, confirming the interpretation of it
541: as a pair of bound magnons.
542: The optical gap (E($1^1B_u^-$)) extrapolates to
543: approximately $1.6$ eV, and is
544: excitonic, lying approximately $1.0$ eV below the
545: charge gap for long chains. As discussed in section I, the energies of the
546: $2^1A_g^+$ and
547: $1^1B_u^-$ states
548: in polyacetylene thin films are believed to be at approximately $1.0
549: -1.8$, and
550: $2.0$ eV,
551: respectively. Approximately $0.3$ eV should be deducted from the
552: calculated $1^1B_u^-$ energy to account for solvation effects
553: \cite{yaron97}, indicating that the undimerised Pariser-Parr-Pople model
554: underestimates the optical gap by approximately $0.7$ eV and the
555: $2^1A_g^+$ energy by up to $1.8$ eV.
556:
557:
558: \section{Solution of the Pariser-Parr-Pople-Peierls model}
559:
560: Sections III and IV indicate that neither electron-lattice coupling nor
561: electron-electron interactions alone are sufficient to explain the
562: low energy excitations of polyene oligomers. A pure electron-phonon
563: model predicts degenerate $1^1B_u^-$ and
564: $1^3B_u^+$ states with the $2^1A_g^+$ state lying above them, while a
565: pure electron interaction model underestimates the optical gap,
566: has gapless spin excitations
567: and does not lead to a dimerised chain. We now turn to the DMRG
568: solution of the Pariser-Parr-Pople-Peierls model. We note that
569: an infinitesimally small electron-phonon coupling will open a gap in the
570: spin excitation spectrum for all electronic interaction strengths.
571:
572: We first
573: calculate the ground state energy and lattice geometry.
574: The normalised staggered
575: bond dimerisation is defined as,
576: \begin{eqnarray}
577: \delta_{\ell} \equiv (-1)^{\ell} \frac {(t_{\ell} - \bar{t}) } { \bar{t}},
578: \end{eqnarray}
579: where $\bar{t}$ is the average value of $t_{\ell}$ in the middle of the
580: chain. $\delta = 0.102$ in the center of the chain. Using
581: $\alpha = 4.593$ eV $\AA^{-1}$, this implies that the
582: bond length alternation of the
583: ground state in the middle of the chain is $0.056$ \AA,
584: in close agreement with the experimental result of $0.052$ \AA
585: \cite{kahlert87}.
586:
587: Using the ground state geometry, the vertical
588: energies (that is, the energies of these states with the ground state
589: geometry) ($E^{\text v}$) of the
590: $1^3B_u^+$,
591: $1^1B_u^-$ and
592: $2^1A_g^+$
593: states are calculated. These, as well as the relaxed energies
594: ($E^{\text{0-0}}$), are shown
595: in Fig.\ 4(a) as a function of inverse chain
596: length. The vertical energy of the $2^1A_g^+$
597: state lies approximately $0.3$ eV above that of the $1^1B_u^-$ state
598: in the long chain limit\cite{footnoteIV}.
599: The relaxation energy of the $1^1B_u^-$ state is modest, being
600: approximately
601: 0.2 eV
602: for $102$ sites.
603: By contrast, the relaxation
604: energies of the $1^3B_u^+$ and $2^1A_g^+$ states are substantial,
605: being approximately $0.8$
606: eV and $1.5$ eV, respectively, and converge rapidly with increasing
607: chain length. The energy of the relaxed $2^1A_g^+$ state lies 1 eV below
608: that of the $1^1B_u^-$ state.
609: We see in section VII that this strong relaxation is
610: associated with a large distortion of the ground state structure.
611:
612: In Fig.\ 4(b) we plot the charge gap,
613: \begin{eqnarray}
614: E(N+1) + E(N-1) -2E(N),
615: \end{eqnarray}
616: and the energy of the $1^1B_u^-$ state.
617: In the long chain limit the charge gap
618: represents the energy of an uncorrelated electron-hole pair, and
619: therefore represents the band edge. The relaxation energy of the
620: charge gap is roughly double that of the $1^1B_u^-$ state. This
621: is to be expected, as the two charges form independent polarons, whereas
622: the excitonic $1^1B_u^-$ state forms a single polaron, as discussed
623: in section VII.
624: We see that the single chain binding energy is $2.4$ eV. However, the
625: unbound pair is strongly solvated (ca.\ $1.5$ eV), while the exciton
626: is more weakly solvated (ca.\ $0.3$ eV) \cite{yaron97}.
627: This implies that the
628: bulk binding energy of the $1^1B_u^-$ state is ca.\ $1$ eV.
629:
630: The experimental
631: values of $E^{\text{0-0}}(1^1B_u^-)$ and
632: $E^{\text{0-0}}(2^1A_g^+)$ for short
633: polyenes
634: are also
635: shown\cite{kohler}. The $2^1A_g^+$ values are in excellent agreement
636: with our calculation. The
637: $1^1B_u^-$ values are approximately
638: $0.3$ eV lower than our predictions,
639: which is approximately the reduction expected by the solvation of the
640: chains in solution \cite{yaron97}. Thus, for short polyene oligomers,
641: the optimised parametrisation of the
642: Pariser-Parr-Pople-Peierls model gives remarkably good results.
643:
644: Kohler has analysed the experimental results for
645: $N = 6 - 16$\cite{kohler}.
646: For the $2^1A_g^+$ state the empirical relation,
647: \begin{eqnarray}
648: E^{\text{0-0}}(2^1A_g^+) = 0.96 + 20.72/N,
649: \end{eqnarray}
650: was derived. This relation appears to confirm the work of
651: \cite{halvorson93}, who
652: find a $2^1A_g^+$ state at $1.1$ eV in thin films. However, there is no
653: particular reason why a linear extrapolation in $1/N$ is valid.
654: Our calculation
655: for the Pariser-Parr-Pople-Peierls model shows a significant flattening
656: off of the
657: $2^1A_g^+$ energy for chain lengths of roughly $30$ or more sites. The
658: calculated converged energy of $1.74$ eV is in agreement with \cite{fann89}.
659:
660: This rapid convergence of energy with chain length is in contrast to both
661: the
662: Pariser-Parr-Pople and Peierls models. In the Peierls model the excitation
663: energies
664: are gapped, but the deviation from $1/N$ behavior is only evident for
665: long chains
666: (ca.\ $100$ sites). In the Pariser-Parr-Pople model a deviation from $1/N$
667: behavior is only evident in the long chain limit for the $1^1B_u^-$ state
668: and the
669: charge gap. In the Pariser-Parr-Pople-Peierls model, however, states
670: which form
671: pronounced solitonic structures, such as the $2^1A_g^+$ and triplet
672: states (as
673: discussed in section VII) self-trap once the chain length exceeds the
674: size of their
675: solitonic structures. It is possible that this self-trapping
676: is a consequence of the
677: adiabatic treatment of the lattice, and that a full treatment
678: involving quantum
679: phonons would change this prediction.
680:
681: Our understanding of self-trapping - and its validity or otherwise - is
682: complicated by
683: the discussion of the $1^1B_u^-$ state energy. Again,
684: an empirical relation,
685: \begin{eqnarray}
686: E^{\text{0-0}}(1^1B_u^-) = 2.01 + 15.60/N,
687: \end{eqnarray}
688: was derived by Kohler, which is in good agreement with the
689: thin film result. Our calculated value of $2.74$ eV is too high,
690: even when solvation effects (ca.\ $0.3$ eV) are deducted. Once again, the
691: $1^1B_u^-$ state is self-trapped, and the possible relaxation by lattice
692: fluctuations
693: would lead to a better agreement.
694:
695: However, since the phonon frequency of ca.\ $0.2$ eV is so small compared
696: to the
697: electronic energy scales, any corrections to the adiabatic limit
698: are expected to be
699: small, so we need to consider other possible reasons for the
700: discrepencies in the
701: long chain limit. One source is the possible renomalisation of the
702: Pariser-Parr-Pople-Peierls model parameters in the long chain limit;
703: another is
704: $\sigma$-electron screening.
705:
706:
707: \section{Photo-induced absorption}
708:
709: The photo-induced absorption spectrum of a
710: system, obtained while it is being pumped
711: at an energy above the optical gap, gives an insight into the excited
712: states of that system \cite{vardeny}. Typically the system is pumped
713: at $2.4$ eV, and photo-induced absorption peaks are observed at $0.43$ eV
714: and $1.35$ eV.
715:
716: The higher energy peak is believed to intrinsic and has been ascribed
717: to a bound soliton-anti-soliton pair \cite{orenstein}. A possible
718: interpretation is that excitations to states above the vertical
719: $1^1B_u^-$ state decay non-radiatively to the $2^1A_g^+$ state,
720: which subsequently relaxes. The
721: photo-induced absorption is then a vertical transition from the relaxed
722: $2^1A_g^+$ state to a $^1B_u^-$ state. We find
723: that the
724: energy of the $1^1B_u^-$ state in the relaxed geometry of the $2^1A_g^+$
725: state lies
726: 1.3 eV above the $2^1A_g^+$ state for $102$ sites. However, the dipole moment
727: is weak, being only $ 0.16 \langle \mu \rangle_{1^1B_u^-}$ (where
728: $\langle \mu \rangle_{1^1B_u^-}$ is the dipole moment between the ground state
729: and $1^1B_u^-$ state). A second possibility is that it is a
730: triplet-triplet ($T \rightarrow T^{\ast}$)
731: transition. We calculate this transition energy
732: to be $2.8$ eV, while the dipole moment is
733: $0.96 \langle \mu \rangle_{1^1B_u^-}$. Since the excited triplet
734: ($T^{\ast}$) is
735: a triplet-exciton (as opposed to a spin-density wave excitation) at
736: high energy, it is reasonably to assume that it will be strongly solvated,
737: reducing this transition energy by as much as $1$ eV.
738: Thus, a triplet to triplet transition is a possible explanation
739: for this absorption.
740:
741: We calculate the transition energy between the lowest polaron state
742: and the first dipole connected excitation to be $0.45$ eV at $102$
743: sites, and the
744: dipole moment
745: is
746: $0.88 \langle \mu \rangle_{1^1B_u^-}$,
747: suggesting that this is the origin of the lower peak.
748:
749:
750: \section{Soliton Structures}
751:
752: In Fig.\ 5(a) we plot, as a function of bond index from the
753: center of the chain, the
754: normalised staggered bond dimerisation, $\delta_{\ell}$, Eqn.\ (14).
755: We note that the $1^3B_u^+$ and
756: $2^1A_g^+$
757: states undergo considerable bond distortion, whereas
758: the $1^1B_u^-$ state shows a weak polaronic
759: distortion of the lattice, similar to the distortion associated with
760: a doped charge.
761: In \cite{bursill99} we showed that the $1^3B_u^+$ and
762: $1^1B_u^-$ states
763: fit a 2-soliton form
764: \cite{su95}, \cite{campbell}, \cite{bach},
765: whereas the $2^1A_g^+$ state fits a 4-soliton form.
766: The bond distortions of the non-interacting limit (the Peierls model)
767: are plotted in Fig.\ 5(b).
768: A comparison between these plots
769: illustrates the role played by the electronic
770: interactions
771: in
772: modifying the non-interacting picture:
773: \begin{enumerate}
774:
775: \item
776: The dimerisation in the ground state is enhanced by a five-fold factor, in
777: qualitative agreement with \cite{horsch81} and \cite{konig90}.
778:
779: \item
780: The $1^1B_u^-$ state evolves to an exciton-polaron, in agreement with
781: \cite{grabowski85}.
782:
783: \item
784: The $2^1A_g^+$ state, owing to its strong triplet-triplet contribution,
785: evolves to a
786: 4-soliton solution, in agreement with \cite{hayden86}.
787:
788: \end{enumerate}
789:
790: Further insight into the electronic structure of
791: polyenes and its relation to their geometry can be obtained from the
792: spin-spin correlation function, defined as,
793: \begin{equation}
794: S_i = -\langle S^Z_i S^Z_{N+1-i} \rangle.
795: \end{equation}
796: This function measures anti-ferromagnetic
797: correlations between sites symmetrically
798: situated with respect to the center of the chain.
799: As the correlation function shows unimportant
800: oscillations between even and odd site indices $i$,
801: we use the symmetrized function:
802: \begin{equation}
803: \tilde{S}_j = \frac{1}{2} ( S_{\frac{1}{2}(N-j)}+S_{\frac{1}{2}(N-
804: j)+1}),
805: \end{equation}
806: $j=0,4,8,...,N-2$, which measures the correlations between
807: pairs of doubly-bonded sites, with $j$ being the distance between
808: them.
809:
810: The spin-spin correlation
811: functions, calculated in the ground state
812: geometry, are shown in Fig.\ 6(a).
813: They show a monotonic decay
814: for the correlations in the $1^1A^+_g$ and $1^1B^-_u$ states,
815: but in the $2^1A^+_g$ state there is a small minimum at $j=8$
816: and a maximum at $j=16$.
817: This behavior of the spin-spin correlations in the $2^1A^+_g$ state
818: becomes clearer when we calculate it in the relaxed geometry for
819: this state. Here, the correlation function of the $2^1A^+_g$ state,
820: shown in Fig.\ 6(b), has a strong minimum at $j=8$,
821: where it changes sign, and a maximum at $j=20$. These features
822: strongly confirm the triplet-triplet character of this state. By
823: comparing
824: Fig.\ 6(b) to
825: the soliton structure shown in Fig.\ 5(a), we
826: see
827: that the
828: unpaired spins correspond to the positions of the geometrical solitons.
829:
830:
831:
832: \section{Discussion and Conclusions}
833:
834: We began this investigation of the electronic and geometrical structure
835: of linear polyenes by performing seperate studies of the
836: $U=0$ Peierls model and the $\lambda=0$ Pariser-Parr-Pople model.
837: These studies show that these two limits predict quite different
838: low-lying excitations. The Peierls model predicts mid-gap states
839: associated with geometrical defects. The dipole forbidden
840: $2^1A_g^+$ state lies above the degenerate singlet and triplet
841: $1B_u$ states. In contrast, the Pariser-Parr-Pople model predicts
842: gapless (or very small gapped) triplet and $2^1A_g^+$ states, with
843: the $1^1B_u^-$ state lying above them.
844:
845: When these two models are combined in the Pariser-Parr-Pople-Peierls model
846: we
847: see the effect of the inter-play of electron-electron and electron-phonon
848: interactions.
849: The lowest lying triplet
850: ($1^3B_u^+$) is a soliton-antisoliton pair; the lowest lying singlet
851: ($2^1A_g^+$) is an even-parity pair of
852: soliton-antisoliton pairs, owing to it being a bound pair of triplets;
853: and the lowest optically allowed state ($1^1B_u^-$) is an
854: exciton-polaron.
855: The soliton positions in the $2^1A_g^+$ state
856: is confirmed by the spin-spin correlation function.
857: Electron-electron interactions play the dominant role in opening the
858: optical gap and dimerising the lattice.
859:
860:
861: We find that the relaxation energy of the $1^3B_u^+$ and $2^1A_g^+$
862: states are
863: substantial,
864: whereas that of the $1^1B_u^-$ state is modest. The vertical energy of the
865: $2^1A_g^+$ state lies above that of the $1^1B_u^-$ state, but the relaxed
866: $2^1A_g^+$ state lies ca.\ $1.0$ eV below that of the $1^1B_u^-$ state.
867: The role
868: of electron-electron interactions are crucial and subtle in determining
869: these relative
870: positions. A larger electron-electron interaction leads to a more
871: dimerised ground
872: state, and this tends to raise the vertical energy of the $2^1A_g^+$ state
873: relative to that
874: of the $1^1B_u^-$ state. However, a larger electron-electron interaction
875: also leads to
876: a larger relaxation of the $2^1A_g^+$ state energy compared to that of the
877: $1^1B_u^-$ state, leading to a reversal of their energies.
878:
879: For short polyenes we find good agreement with experimental values.
880: However, in
881: the long chain limit the results (at least for the $1^1B_u^-$ state) become
882: more
883: qualitative. The experimental uncertainty in the position of the
884: $2^1A_g^+$ state
885: means that we cannot be sure of the validity of our prediction.
886: However, if we
887: assume that ca.\ $1.0$ eV is the correct relaxed energy of
888: the $2^1A_g^+$ state,
889: then our predictions are between $0.5$ to $1.0$ eV too high.
890: In section V we
891: discussed some of the possible origins of these discrepencies.
892: They include, the
893: neglect of lattice fluctuations in the adiabatic treatment of the
894: lattice, the possible
895: renormalisation of the $\pi$-model parameters in the long chain
896: limit, and the neglect of
897: the $\sigma$-bond screening.
898: We would expect that as a molecule gets larger the $\pi$ orbitals
899: will become more extended, as they mix with other orbitals. This will
900: reduce $U$ and $\alpha$ (and hence $\lambda$), and increase $t_0$, thus
901: reducing the excitation energies.
902: Work is currently in progress to
903: study these affects.
904:
905:
906:
907: \vspace{2 cm}
908:
909:
910: {\bf Acknowledgments}
911:
912: This work was started while W.\ B.\ was on Study Leave at the UNSW. He
913: thanks
914: the
915: Royal Society and the Gordon Godfrey Committee of the UNSW for
916: financial
917: support. R.\ J.\ B.\ was supported by the Australian Research Council
918: and
919: M.\ Yu.\ L.\ was supported by the EPSRC (U.K.) (GR/K86343).
920: We thank D.\ Yaron and Z.\ Vardeny for discussions.
921:
922:
923: \begin{references}
924:
925: \bibitem[*]{email}
926: Email address: W.Barford@sheffield.ac.uk.
927:
928: \bibitem[**]{leave}
929: Current address: The School of Chemistry, University of Bristol,
930: Bristol, U.\ K.\
931: On leave from Institute of Inorganic Chemistry, 630090 Novosibirsk,
932: Russia.
933:
934: \bibitem{footnoteI} As defined by the energy to add a particle and a hole
935: to the half filled system.
936:
937: \bibitem{hudson72} B.\ S.\ Hudson and B.\ E.\ Kohler, Chem.\ Phys.\
938: Lett.\
939: {\bf 14},
940: 299 (1972).
941:
942: \bibitem{schulten72} K.\ Schulten and M.\ Karpus, Chem.\ Phys.\ Lett.\
943: {\bf 14}, 305
944: (1972).
945:
946: \bibitem{tavan87} P.\ Tavan and K.\ Schulten, Phys.\ Rev.\ B {\bf 36},
947: 4337 (1987).
948:
949: \bibitem{refs} R. R. Birge, K. Schulten and M. Karplus, Chem. Phys. Letts.
950: {\bf 31}, 451 (1975); R. McDiarmid, J. Chem. Phys. {\bf 79}, 1565 (1983);
951: B. Hudson and B. Kohler, Synth. Met. {\bf 9}, 241 (1984).
952:
953: \bibitem{hayden86}
954: G.\ W.\ Hayden and E.\ J.\ Mele, Phys.\ Rev.\ B {\bf 34}, 5484 (1986).
955:
956: \bibitem{su95}
957: W.\ P.\ Su, Phys.\ Rev.\ Lett.\ {\bf 74}, 1167 (1995).
958:
959: \bibitem{wen97} G-Z Wen and W-P Su,
960: Relaxations of Excited States and Photo-induced Structural Phase
961: Transitions (p 121),
962: Springer-Verlag 1997; Synth.\ Metals {\bf 78}, 195 (1996).
963:
964: \bibitem{ovchinnikov} A.\ A.\ Ovchinnikov, I. I. Ukrainskii and
965: G.\ V.\ Kventsel, Sov.\ Phys.\ ---Usp.\ {\bf 15}, 575 (1973).
966:
967: \bibitem{grabowski85} M.\ Grabowski, D.\ Hone and J.\ R.\ Schrieffer,
968: Phys.\ Rev.\
969: B {\bf 31}, 7850 (1985)
970:
971: \bibitem{halvorson93} C.\ Halvorson and A.\ J.\ Heeger, Chem.\ Phys.\
972: Lett.\ {\bf
973: 216}, 488 (1993).
974:
975: \bibitem{vardeny} Relaxation in Polymers (p 174), ed. by T. Kobayashi.
976: World
977: Scientific (Singapore) 1993.
978:
979: \bibitem{fann89} W.-S. Fann, {\em et al.}, Phys.\ Rev.\ Lett.\,
980: {\bf 62}, 1492 (1989).
981:
982: \bibitem{kiess} D.\ Baeriswyl, D.\ K.\ Campbell and S.\ Mazumdar, in
983: {\em Conjugated Conducting Polymers}, edited by H.\ Kiess (Springer-Verlag,
984: Berlin, 1992).
985:
986: \bibitem{bursill98}
987: R.\ J.\ Bursill, C.\ Castleton, W.\ Barford, Chem.\ Phys.\ Lett.\ {\bf
988: 294},
989: 305 (1998).
990:
991: \bibitem{bursill99} R.\ J.\ Bursill and W.\ Barford,
992: Phys.\ Rev.\ Lett.\,
993: {\bf 82}, 1514 (1999).
994:
995: \bibitem{White}
996: S.\ R.\ White, Phys.\ Rev.\ Lett.\ {\bf 69}, 2863 (1992);
997: Phys.\ Rev.\ B {\bf 48}, 10345 (1993).
998:
999: \bibitem{book} Density Matrix Renormalisation, edited by I.\ Peschel, X.\
1000: Wang, M.\
1001: Kaulke and K.\ Hallberg, Springer, Berlin, 1999.
1002:
1003: \bibitem{horsch81} P. Horsch, Phys.\ Rev.\ B
1004: {\bf 24}, 7351 (1981).
1005:
1006: \bibitem{konig90} G.\ Konig and G.\ Stollhoff,
1007: Phys.\ Rev.\ Lett.\,
1008: {\bf 65}, 1239 (1990).
1009:
1010:
1011: \bibitem{takimoto89}
1012: J.\ Takimoto and M.\ Sasai, Phys.\ Rev.\ B {\bf 39}, 8511 (1989).
1013:
1014: \bibitem{gammel93}
1015: J.\ T.\ Gammel and D.\ K.\ Campbell, Synth.\ Met.\ {\bf 55}, 4638 (1993).
1016:
1017:
1018: \bibitem{yaron98}
1019: D.\ Yaron, E.\ E.\ Moore, Z.\ Shuai, J.\ J.\ Bredas, J.\ Chem.\ Phys.\
1020: {\bf 108}, 7451 (1998).
1021:
1022: \bibitem{fano98}
1023: G.\ Fano, F.\ Ortolani, L.\ Ziosi, J.\ Chem.\ Phys\ {\bf 108}, 9246
1024: (1998).
1025:
1026: \bibitem{ehrenfreund87}
1027: E.\ Ehrenfreund, Z.\ Vardeny, O.\ Barfman, B.\ Horovitz, Phys.\ Rev.\ B
1028: {\bf 36},
1029: 1535 (1987).
1030:
1031:
1032: \bibitem{pople} J.\ A.\ Pople and S.\ H.\ Walmsley, Molec.\ Phys.\ {\bf
1033: 5},
1034: 15
1035: (1962).
1036:
1037: \bibitem{ssh} A.\ J.\ Heeger, S.\ Kivelson, J.\ R.\ Schrieffer and
1038: W-P
1039: Su,
1040: Rev.\ Mod.\ Phys.\ {\bf 60}, 781 (1988).
1041:
1042: \bibitem{footnoteIII} However, the exact $1^1B^-_u$ transition energy
1043: obtained from Eq.\ (1) at 102 sites is
1044: $0.23$ eV, while the $2^1A^+_g$ state is at $0.33$ eV.
1045:
1046:
1047: \bibitem{yaron97}
1048: E. Moore, B. Gherman B. and D. Yaron, J. Chem.\ Phys.\ {\bf 106}, 4216
1049: (1997).
1050:
1051: \bibitem{kahlert87} H.\ Kahlert, O.\ Leitner and G.\ Leising, Synth.
1052: Met. {\bf 17}, 467 (1987).
1053:
1054: \bibitem{footnoteIV} Experimentally, this implies that the vertical energy
1055: of the $2^1A^+_g$ state lies ca.\ $0.6$ eV above the $1^1B^-_u$ state,
1056: because of the greater solvation
1057: energy of the latter state.
1058:
1059: \bibitem{kohler} B.\ E.\ Kohler, J.\ Chem.\ Phys.\ {\bf 88}, 2788 (1988).
1060:
1061: \bibitem{orenstein} J.\ Orenstein and G.\ L.\ Baker, Phys.\ Rev.\ Lett.\,
1062: {\bf 49}, 1043 (1982).
1063:
1064: \bibitem{campbell}
1065: D. K. Campbell and A. R. Bishop, Nucl. Phys. B {\bf 200}, 297 (1982).
1066:
1067: \bibitem{bach} M. A. Garcia-Bach, R. Valenti, S. A. Alexander and D. J.
1068: Klein,
1069: Croatia Chemica Acta {\bf 64}, 415 (1991).
1070:
1071:
1072: \end{references}
1073:
1074:
1075: \begin{table}[h]
1076: \caption{
1077: The classification of the relevant states.
1078: }
1079: \begin{tabular}{cccc}
1080: State & $^1A_g^+$ & $^1B_u^-$ & $^3B_u^+$ \\
1081: \hline
1082: Spatial Inversion Symmetry & + & $-$ & $-$ \\
1083: Spin & 0 & 0 & 1 \\
1084: Particle-hole symmetry & + & $-$ & + \\
1085: Character & Covalent & Ionic & Covalent \\
1086: \end{tabular}
1087: \label{states}
1088: \end{table}
1089:
1090: \begin{table}[h]
1091: \caption{
1092: The density-density correlator as a function of distance
1093: }
1094: \begin{tabular}{cc}
1095: $j$ & $\langle (n_i - 1)(n_{i+j} - 1) \rangle$ \\
1096: \hline
1097: 1 & $-0.308$ \\
1098: 2 & $+0.002$ \\
1099: 3 & $-0.021$ \\
1100: 4 & $+0.004$ \\
1101: 5 & $-0.011$ \\
1102: \end{tabular}
1103: \end{table}
1104:
1105: \begin{figure}[p]
1106: \caption{
1107: Energy level diagram for the key low-lying states in the non-interacting
1108: limit.
1109: }
1110: \label{MO}
1111: \end{figure}
1112:
1113:
1114: \begin{figure}[p]
1115: \caption{
1116: Transition energies for the $1^1B_u^-$ (squares), $2^1A_g^+$
1117: (diamonds),
1118: $1^3B_u^+$ (triangles) states and charge gap (circles)
1119: as a function of inverse chain length for the $U=0$ Peierls model
1120: (dashed lines and open symbols) and the $\lambda=0$
1121: Pariser-Parr-Pople model (solid lines and filled symbols).
1122: (In the Peierls model the $1^1B_u^-$ and $1^3B_u^+$ states are degenerate.)
1123: }
1124: \label{PPP}
1125: \end{figure}
1126:
1127: \begin{figure}[p]
1128: \caption{
1129: The difference between the exact calculation of the $2^1A^+_g$
1130: (diamonds)
1131: and $1^1B^-_u$ (squares) states in the non-interacting limit, and the
1132: results
1133: of DMRG calculations in the infinite and finite algorithms.
1134: Solid lines correspond to infinite lattice algorithm results,
1135: dashed lines to the finite lattice algorithm.
1136: }
1137: \label{comparison1}
1138: \end{figure}
1139:
1140:
1141: \begin{figure}[p]
1142: \caption{
1143: (a) Transition energies for the $1^1B_u^-$ (squares), $2^1A_g^+$
1144: (diamonds)
1145: and
1146: $1^3B_u^+$ (triangles) states as a function of inverse chain length.
1147: Vertical/relaxed transitions are indicated by dashed/solid lines
1148: and open/solid
1149: symbols.
1150: Experimental results for the relaxed $1^1B_u^-$ ($\times$) and
1151: $2^1A_g^+$ ($+$)
1152: state energies for polyenes in hydrocarbon solution
1153: \protect\cite{kohler}.
1154: (b) Transition energies for the $1^1B_u^-$ state (squares) and charge
1155: gap (circles) as a function of inverse chain length.
1156: }
1157: \label{energies}
1158: \end{figure}
1159:
1160: \begin{figure}[p]
1161: \caption{
1162: (a) The geometries (normalised
1163: staggered bond distortion $\delta_{\ell}$
1164: as a function of bond index $\ell$ from the
1165: center of the lattice) of various states of the
1166: Pariser-Parr-Pople-Peierls model: $1^1A_g^+$ (crosses),
1167: $1^1B_u^-$
1168: (squares),
1169: $1^3B_u^+$ (triangles) $2^1A_g^+$ (diamonds) and polaron (circles),
1170: for the $102$ site system.
1171: (b) The same as (a) for the $U=0$ Peierls model.
1172: }
1173: \label{geometries}
1174: \end{figure}
1175:
1176: \begin{figure}[p]
1177: \caption{
1178: Spin-spin correlation functions for $1^1A^+_g$ (solid squares),
1179: $2^1A^+_g$ (solid diamonds) and $1^1B^-_u$ (empty squares) states.
1180: (a) In the relaxed $1^1A^+_g$ geometry,
1181: (b) in the relaxed $2^1A^+_g$ geometry.
1182: }
1183: \label{correlations1}
1184: \end{figure}
1185:
1186:
1187: \end{document}
1188:
1189:
1190:
1191: