Sie sind auf Seite 1von 72

Excimers

J B BIRKS
The Schuster Laboratory, University of Manchester, Manchester M13 9PL

Abstract
Excimers are dimers with associated excited electronic states, dissociative ground states, and structureless emission spectra. Noble and other monatomic gases form atomic excimers. Aromatic molecules form excimers in fluid solutions, liquids, crystals and polymers, at crystal defects, and intramolecularly. Excimer interaction is attributed to configurational mixing of exciton and charge resonance states. The helium excimer and pyrene crystal dimer potential curves are compared. Studies of aromatic excimers discussed include: (i) solution fluorescence kinetics yielding thermodynamic and rate parameters ; (ii) excimer formation by dimer cation neutralization, molecular ion association, and triplet-triplet interaction; (iii) steric effects on photodimer, sandwich dimer and excimer conformations; (iv) environmental effects on excimer radiative and radiationless transitions ; (v) excimer exciton migration and interaction in crystals; and (vi) evidence for triplet excimers. Related photophysical studies on atomic excimers are considered. Saturated amines, which exhibit vapour and solution excimer fluorescence, provide a link between atomic and aromatic excimers. Aromatic molecules form complexes, exciplexes or mixed excimers with different molecules, and noble gas atoms form complexes or exciplexeswith different atoms. This review was completed in January 1975.

Rep. Prog. Phys. 1975 38 903-974 60

904

J B Birks

Contents

1. Introduction . 1.1. What is an excimer ? , . 1.2. The occurrence of excimers 1.3. Summary . 2. Excimer interaction 2.1. Exciton resonance states . 2.2. Charge resonance states . 2.3. Nature of the aromatic excimer state . 2.4. Configurational mixing . 3. Excimer potentials . 3.1. The potential diagram . 3.2. The helium excimer 3.3. T h e pyrene crystal dimer , 4. Excimer reaction kinetics 4.1. Photostationary kinetics 4.2. Transient kinetics . 4.3. High-temperature kinetics 4.4. Thermodynamics . 4.5. Diffusion-controlled processes . . 5. Different modes of excimer formation 5.1. Dimer cation neutralization . 5.2. Molecular ion association . 5.3. Triplet-triplet interaction . 6. Photodimers, sandwich dimers and excimers 7. Photophysical processes in excimers . 7.1. Radiationless processes . 7.2. Radiative processes . 8. Excimer excitons in crystals . 9. Triplet aromatic excimers . 9.1. Theoretical considerations . 9.2. Benzene . 9.3. Naphthalene . 9.4. Pyrene . 9.5. Triplet concentration quenching . 9.6. Halobenzenes 10. Atomic excimers . 11. Other excimers . 12. Exciplexes, DA complexes and mixed excimers 12.1. Exciplexes and DA complexes . 12.2. Mixed excimers . . 12.3. Atomic exciplexes and complexes 13. Conclusion

Page

Acknowledgments Appendix . References .

905 905 907 910 911 911 912 913 914 915 915 917 918 921 921 923 925 926 927 928 928 930 931 935 938 938 939 943 944 944 946 947 948 948 949 950 956 957 957 958 959 960 960 961 968

Excimers

905

1. Introduction

I . I . What is an excimer?
There are certain analogies between the spectroscopic properties of aromatic hydrocarbons like benzene and naphthalene and those of the noble gases. On the perimeter-free-electron-orbital model of Platt (1949) the cata-condensed aromatic hydrocarbons have 'closed-shell' f4 v electron configurations in the ground state which are similar to the ground-state closed-shell electron configurations of the noble gases. I n consequence the ground electronic state of an aromatic hydrocarbon molecule or a noble gas atom is a singlet state So, and their excited electronic states are either singlet states Si, S2, . . ., S p or triplet states Ti, T2, . . ., T,. T h e atomic fluorescence spectrum of a noble gas at low pressures consists of a series of lines corresponding to allowed radiative transitions between different electronic states. An increase in the pressure causes (a) a decrease in the atomic fluorescence quantum yield due to self (pressure) quenching, and (b) the appearance of one or more structureless bands or continua in the fluorescence spectrum at lower energies than the quenched atomic fluorescence. T h e molecular fluorescence spectrum of a condensed aromatic hydrocarbon in dilute fluid solution consists of a series of vibronic bands corresponding to radiative transitions from the first T electronic excited singlet state SI of the molecule to different vibrational levels of its ground singlet state So. For most condensed aromatic hydrocarbons (exceptions are discussed in 93.1) an increase in solution concentration causes (a) a decrease in the molecular fluorescence quantum yield due to self (concentration) quenching, and (b) the appearance of a structureless band in the fluorescence spectrum at lower energies than the quenched structured molecular fluorescence. T h e concentration dependence of the fluorescence spectrum of pyrene in cyclohexane solution is shown in figure 1. T h e concentration dependence of the fluorescence of the noble gases and the aromatic hydrocarbons may be summarized as follows. (i) T h e fluorescence (hvM) of the singlet excited atom or molecule (1M")

1M"+lM + h v ~ (1) is subject to self-quenching. (ii) T h e self-quenching is due to the collisional interaction of lMx with an unexcited atom or molecule (1M) 1M" +lM+lDQ (11) which yields a singlet excited dimer (ID"). (iii) This excited dimer is dissociated in the ground state, so that its fluorescence emission (hvD)is structureless,
(111) T h e noble gases are monatomic, so that their dimers are clearly dissociated in the ground state. T h e absorption spectra of the aromatic hydrocarbons in solution are

1 D " d M + 1M + hvD.

906

J B Birks

independent of concentration, showing that the excited dimers responsible for the structureless emission (111) are not present in the ground state. T h e term excimm was introduced by Stevens and Hutton (1960) to describe ID+ and to distinguish it from the excited state of a stable dimer. An excimer was defined as a dimer which is associated in an excited electronic state and dissociated in its ground electronic state. This definition is adequate for intermolecular excimers in fluid media, but it requires slight revision for excimers in rigid media or for intramolecular excimers, where there are environmental and/or steric restraints on molecular motion. To include these cases Birks (1975a) has redefined an excimer as 'a dimer

Figure 1. Fluorescence spectra of pyrene solutions in cyclohexane. Intensities are normalized

to a common value of the molecular fluorescence quantum yield (DFM. Concentrations:A,10-2M;B,7*75~10-~M C ,;5 * 5 ~ 1 0 - 3 M ; D , 3 - 2 5 ~ 1 0 - 3 M ; 10-3M; E, F, 10-4 M (after Birks and Christophorou 1963a).

which is associated in an excited electronic state and which is dissociative (;e would dissociate in the absence of external restraints) in its ground electronic state'. T h e revised definition includes such entities as the pyrene crystal dimer, the anthracene sandwich dimer in rigid solution, and intramolecular excimers, all of which were excluded by the original definition. T h e term 'excimer', unless qualified, is taken to refer to a singlet excimer ID* in its lowest-energy associated state. Collisional interaction between a triplet excited atom or molecule (3M") and an unexcited atom or molecule (1M)

3M*f1M+3Da

(IV)

may yield a triplet excimer 3D+ in its lowest-energy associated state. Higher excited states of ID* and 3D* can also exist, and they are described as excited (singlet) excimers ID++and excited triplet excimers 3D++respectively.

Excimers

907

Rayleigh (1927) observed the ultraviolet emission spectrum of high-pressure mercury vapour which has a broad continuum with a 330 nm wavelength maximum superimposed on the partially quenched atomic line spectrum. The continuous emission was assigned to a transition from an excited dimer state Hg,** to the dissociated ground state (Hg+ Hg). Hg,**, which results from the interaction of Hge*(3P1) and Hg(lSo), corresponds to an excited triplet excimer state 3D**. There is a lower-energy excited dimer state Hg,*, resulting from the interaction of Hg* (3Po) and Hg (I&) and responsible for the 480 nm wavelength emission continuum in mercury vapour, which corresponds to the lowest triplet excimer state 3D+ (Herzberg 1950). The first singlet excimer to be identified was that of helium (He). Hopfield (1930a,b) observed an intense continuum between 60 and 100 nm wavelength in the vacuum ultraviolet emission spectrum of a high-pressure helium discharge and attribbuted it to a transition from an excited dimer He? to the dissociated ground state (He+He). Het is the lowest associated state (12:) of the helium singlet excimer. Mulliken's (1932) calculations of the potential curves of the helium excimer initiated an extensive series of such calculations. The ultraviolet emission continuum from a high-pressure xenon (Xe) lamp, which is commonly used as a spectroscopic source, originates from the xenon excimer Xet. The excimer fluorescence spectra of the other noble gases, neon (Ne), argon (Ar) and krypton (Kr), which extend into the vacuum ultraviolet, have also been observed, but they have attracted little attention until recently. Unlike He; and Hg,**, they were omitted from Herzberg's (1950) compilation of 'all known diatomic molecules'. They are discussed in $10. An emission continuum in the vapour phase does not necessarily originate from an excimer. The molecular hydrogen (Hz) discharge spectrum has an intense continuous emission extending from 160 to 500 nm wavelength. This corresponds to the transition from the lowest excited triplet state of the Hz molecule (32c,') to the dissociated repulsive triplet state (32:) of two normal hydrogen atoms (Winans and Stueckelberg 1928). The optical selection rules allow this transition, but they forbid the transition to the stable singlet ground state (I&) of the Hz molecule, which is of lower energy than 3X:. Other causes of continuous spectra have been discussed by Finkelnburg (1938) and Herzberg (1950).

1.2. The occurrence o f excimers 1.2.1. Gases. The first excimer to be identified was that of mercury (Hg). Lord

1.2.2. Fluid solutions. Excimer fluorescence in an aromatic hydrocarbon was first identified by Forster and Kasper (1954, 1955), who observed the concentration dependence of the fluorescence spectrum of pyrene in fluid solution (figure 1) and interpreted it in terms of processes (1)-(111). The observation of excimer fluorescence from a fluid solution depends on the magnitude of the molar fluorescence yield ratio K1=@FD/@FM['M] (1.1) where @FD and @FM are the ID" and 1M" fluorescence quantum yields respectively at a molar concentration [IM]. The parameter K1 depends on the solute, solvent and temperature. Pyrene solutions have high values of K1: for pyrene in cyclohexane solution at 20 "C K1= 2-4x 103 M-1 (Birks et a1 1963b))so that (DFD exceeds (DFM at [1M] > 4.2 x 10-4 M (figure 1). Because of the high values of K1, the molecular and excimer fluorescence of pyrene solutions has been extensively studied. The thermodynamic parameters of excimer

908

J B Birh

formation and dissociation (11) have been evaluated from observations of K1 as a function of temperature (Doller and Forster 1962b, Stevens and Ban 1964) and of pressure (Forster et al 1963). The rates of excimer formation and dissociation and of molecular and excimer relaxation (processes 1-111) in pyrene solutions at different temperatures have been determined from observations of the 1M" and ID* fluorescence response functions following nanosecond flash excitation of 1M*, and also by phase and modulation fluorometry (Birks et al1963b). These appear to have been the first dynamic studies of a bimolecular process on a nanosecond timescale. Similar observations on pyrene in several different solvents over a wide range of temperature and viscosity were used to determine the frequency factors and activation energies of the different rate processes (Birks et al 1964c,d). They also showed the changes in enthalpy and entropy of the pyrene excimer to be independent of the solvent. The solvent viscosity q influences the rate parameter KDM of excimer formation (11). The latter process is diffusion-controlled, and observations of KDM and q have provided a means of testing the validity of the Stokes-Einstein molecular diffusion relation (Alwattar et al 1973). Most aromatic compounds have much lower values of K1 than pyrene, and for several years the excimer fluorescence of pyrene solutions appeared to be exceptional. The situation changed dramatically in 1962 when several other aromatic hydrocarbons, including benzene (Ivanova et al 1962), naphthalene (Doller and Forster 1962a) and 1:2-benzanthracene (Birks and Christophorou 1962a,b) and many of their derivatives, were also observed to exhibit excimer fluorescence in solution. I t soon became clear that excimer formation by aromatic hydrocarbons is the rule rather than the exception, although care and ingenuity is sometimes required to observe the excimer fluorescence.

1.2.3. Pure liquids. The maximum value of (DFD for an aromatic compound in solution is restricted by its solubility, which limits the concentration [IM]. Above the melting point of the compound the solvent can be omitted, and [1M] is then a maximum for the pure liquid. The fluorescence spectrum of an aromatic liquid just above its melting point often includes a strong excimer component which can be observed more readily than in solution (Ivanova et al 1962, Stevens and Dickinson 1963a, Birks and Aladekomo 1964). Perylene is a striking example: the excimer fluorescence is dominant in liquid perylene at 270 "C (J Langelaar and G J Hoytink 1968, private communication); in a saturated solution of perylene in benzene at 20 "C it is only just detectable (Birks and Christophorou 1964). With optical excitation the fluorescence spectrum of an aromatic liquid commonly contains both molecular (1M") and excimer (ID") components, with the former often obscuring the latter. Carter et al (1967) found that if the liquid is excited with an intense beam of medium-energy (4-36 keV) electrons, the 1M" emission is strongly quenched and the IDx emission is dominant. The excimer fluorescence spectrum of p-xylene, which has not been detected with optical excitation (Birks et al 1965), has been observed with intense electron beam excitation (Christophorou et a l 1968). The reasons for the dependence of the emission spectrum on the mode of excitation are discussed in 55.1.
1.2.4. Photodimers. I n some aromatic molecules the excimer interaction is so strong that a stable photodimer M2 is formed from the excimer ID*. We may write lM"+lM+1D"+

M2

(VI

Excimevs

909

where lD* is the excimer intermediate (Birks 1970a, Ferguson and Mau 1974). Photodimerization, which is an extreme case of excimer formation, has been known since the 19th century to occur in anthracene solutions exposed to sunlight, and many anthracene derivatives behave in a similar manner (Calas and Lalande 1959, Lalande and Calas 1959, 1960, Calas et a Z 1960, 1965, Greene 1960). Photodimerization has also been observed in oxygen-free solutions of the higher polyacenes, tetracene and pentacene (Birks et aZ1963a).
1.2.5. Sandwich dimers. Condensed aromatic hydrocarbons are planar, and their most stable excimer conformation is usually that of a sandwich dimer (M IIM), corresponding to a symmetrical pair of parallel molecules (see $2.1). Figure 2 shows the structure of

Figure 2. Structural diagrams of (i) the anthracene (R = H) sandwich dimer, and (ii) its photodimer. 9-substituted anthracenes (R= CH3, CN, C1, etc) form photodimers and sandwich dimers in the trans conformation, as shown (see $6) (after Tomlinson et a1 1972).

(i) the sandwich dimer (M IIM) of anthracene, consisting of a parallel pair of anthracene molecules belonging to the Dah symmetry point group, and (ii) the photodimer (Ma), dianthracene, which is formed by a pair of covalent 0 bonds joining the meso (9 to 9', 10 to 10') positions in the two molecules. T h e photodimer structure (ii) shows that the (M IIM) conformation (i) is that of the interacting molecules prior to the final step in (V), ie that ID* corresponds to 1(M llM)*. Irradiation of dianthracene or other photodimers with photons of sufficient energy (eg Hg 253.7 nm radiation) causes photolysis, and the photodimer reverts to the original pair of individual molecules. Such photolysis in frozen solutions was introduced by Chandross (1965) as a method for the production of sandwich dimers Mz+hv-+(MIIM) (VI) in which the molecular conformation of figure 2(ii) reverts to that of figure 2(i). The fluorescence of the excited sandwich dimer 1(M jl NI)* in rigid solution resembles that of the excimer ID*, which is not observable in fluid solution because of process (V).

1.2.6. Single crystals. Stevens (1962) classified aromatic molecular crystals into type A or B, dependent on their structural and fluorescence properties. I n pyrene, perylene and other type B crystals the molecules are arranged in sandwich dimers (M IIM) in the crystal lattice, and the crystals emit pure excimer fluorescence (Sangster and Irvine 1956, Ferguson 1958, Birks and Cameron 1959). The maximum of the structureless excimer fluorescence spectrum lies about 6000 cm-1 below the edge of the

910

J B Birks

structured molecular absorption spectrum. The interaction potentials, force constants and zero-point energies of the excited and unexcited states of the pyrene crystal dimer have been determined from an analysis of the temperature dependence of the pyrene crystal fluorescence spectrum (Birks and Kazzaz 1967, 1968). Excimer exciton migration and excitation transfer have been observed in pyrene crystals (Birks et a1 1966a, Klopffer and Bauser 1970). These topics are discussed in 993.3 and 8 respectively.

1.2.7. Defects in crystals. I n naphthalene, anthracene and other type A aromatic molecular crystals the overlap between neighbouring parallel molecules is small and/or the spacing is relatively large. Perfect type A crystals emit only structured molecular (1M") fluorescence, which is the approximate mirror image of the So-Sl absorption spectrum. However, if the crystal has a high defect concentration due to either microcrystallinity (Perkampus and Pohl 1964) or to prior subjection of the crystal to a high pressure (Jones and Nicol 1965), an excimer emission band may appear in the crystal fluorescence spectrum. The excimers are formed at crystal defects, where the molecules have more freedom to interact than in the perfect crystal lattice. Such defects act as traps for the molecular excitons, and thus enhance the excimer fluorescence yield. I n crystals of anthracene and its 9-substituted derivatives, photodimers are also formed preferentially at defects (Craig and Sarti-Fantoni 1966, Williams and Thomas 1972).
1.2.8. Polymers and intramolecular excimers. Intramolecular excimer fluorescence due to the interaction of two identical aromatic groups in the same molecule, eg 1,3diphenylpropane, in solution was discovered by Hirayama (1965). He used it to explain the excimer fluorescence observed from polystyrene and other aromatic polymers (Yanari et a1 1963, B a d e 1964, Vala et aZl965, Fox et a1 1972), although the interacting aromatic groups in the latter may belong either to the same or to different polymer chains. Intramolecular excimer fluorescence also occurs in the dinucleotides of thymine (TPT) and cytosine (cpc) and in polycytosine (poly c) (Eisinger et a1 1966). The thymine excimer is probably the precursor of the thymine photodimer which plays a key role in the radiation damage of DNA. A comprehensive review of intramolecular excimers has been provided by Klopffer (1973).
1.2.9. Triplet excimers. Excimer (3D") phosphorescence has been observed from several halobenzenes : from crystals of 1,4-dibromobenzene, 1,3,5-tribromobenzene and 1,3,5-trichlorobenzene (Castro and Hochstrasser 1966); from rigid solutions of chlorobenzene, bromobenzene and iodobenzene (Lim and Chakrabarti 1967); and from crystals of 1,4-dichIorobenzene and 1,2,4,5-tetrachlorobenzene (George and Morris 1970). Triplet excimer emission from naphthalene and phenanthrene solutions has been reported by Langelaar et a1 (1968), and from pyrene crystals by Gijzeman et a1 (1970). The latter assignments are questionable (Birks 1975a) and will be discussed in $9.

1.3. Summary
Singlet and triplet excimers of the noble and other monatomic gases have been observed in the vapour phase. The scintillations observed from liquid and solid

Excimers

911

noble gases excited by ionizing radiation probably also correspond to excimer emissions (Birks 1964a). Excimers of aromatic molecules have been observed in several condensed-phase environments: in fluid solutions, in pure liquids, in single type B crystals, at defects in type A crystals, in polymers, as sandwich dimers, and as intramolecular excimers. Aromatic molecular excimer fluorescence has yet to be observed in the vapour phase, although Stevens and McCartin (1960) have obtained indirect evidence of excimer formation in aromatic vapours. Excimer fluorescence has been observed from some saturated amines in the vapour and solution phases (Halpern and Maratos 1972, Halpern 1974), and this is discussed in $11. Table A1 (see Appendix) summarizes the conditions under which excimer fluorescence or photodimer formation has been observed in aromatic compounds. A representative bibliography is given for each compound. Molecular and excimer fluorescence data for many of the compounds listed have been published by Birks (1970a,b) and Stevens (1971), and aromatic excimers have also been reviewed by Forster (1962, 1963, 1969) and Birks (1964b, 1975a,b). I n its extension from physics into chemistry and biology, the excimer has increased in size, complexity and diversity. Large aromatic molecules have replaced simple atoms as the excimer constituents. Fluid solutions and liquids, polymers and intramolecular systems, have replaced the vapour phase as the excimer environment. These factors have extended the range of excimer phenomena, but they have not changed their physical nature. The excimer remains a physical entity. The excimer is not a chemical entity, since it exists only while it possesses electronic excitation energy. On dissipation of this energy, the system reverts to a dissociated pair of molecules. T h e following aspects of the physics of excimers are discussed in subsequent sections: excimer interactions ( $2); excimer potentials ( $3); excimer reaction kinetics ($4); different modes of excimer formation ( $ 5 ) ; photodimers, sandwich dimers and excimers ( $6); photophysical processes in excimers ( $7); excimer excitons in crystals ($8); triplet excimers ($9); atomic excimers ($10); other excimers ($11); and exciplexes, donor-acceptor complexes and mixed excimers ( $12).

2. Excimer interaction
There are two types of molecular interaction which can contribute to excimer formation: ( a ) exciton resonance (ER), due to dipole-dipole (or multipole-multipole) interaction between excited and unexcited molecular states; and (b) charge resonance (cR), due to Coulombic interaction between positive and negative molecular ion states. The excimer states are considered to originate from mixtures of ER and CR states.

2.1. Exciton resonance states


An aromatic hydrocarbon molecule or noble gas atom has a singlet ground state So and several excited electronic singlet states S , and triplet states T, of zero-point energies S," and T," respectively. T h e magnitude of the interaction between an excited and unexcited molecule of the same species depends on their separation distance r

912

J B Birk

and relative orientation. Dipole-dipole interaction splits each state S , into a pair of singlet exciton resonance (ER) states of energies

lEp = Si T (mp2/r3)(COS CY. - 3 COS

el COS

(2.1)

where m, is the So-S, electric dipole transition moment, CY. is the angle between the transition moments in the two molecules, and 81 and 82 are the angles between the dipoles and the line of centres. If the two molecules are parallel (CY. = 0, 8 = 81= &), ( 2 . 1 ) simplifies to

l(Ep)g, = Si T (mp2/r3)( 1 - 3 cos2 8)

(2.2)

where the negative and positive signs refer to the even (g) and oad (U) parity states respectively. For a given value of r, the minimum and maximum values of lEp at 8=Oare
1(

l(Ep)g = Si + 2mP2/r3.
The minimum and maximum values of lEp at 8 = 90" are

E , )

=S E - 2mP2/r3

(2.3~) (2.3b) (2.4~) (2.4b)

For spherical atoms or molecules the lowest-energy ER state originating from S , is l ( E p ) u at 8= 0 , given by ( 2 . 3 ~ ) This . applies to the excimers of the noble gases. For planar aromatic hydrocarbon molecules the closest distance of approach between the molecular centres is much less in the symmetrical sandwich (M [IM) conformation (8=90") than in the coplanar one (8=0). Under this condition the . lowest-energy ER state originating from S, is l(Ep)gat 8=90", given by ( 2 . 4 ~ ) The excimers of the noble gases and the aromatic hydrocarbons thus differ in energy and parity because of the differences in molecular shape. The zero-order wavefunctions of the singlet ER states of the (M jlM) excimer are

4( l ( E P ) P ) = $ l ( S , )
where
$1

$2(SO)

+ $1(SO) $2(S,)

( 2 .Sa)

and

$2

(2,Sb) $ (l(EP>u>= $ l ( S P ) M S O ) - $l(SO) + 2 ( S P ) refer to molecules 1 and 2 respectively (Murre11 and Tanaka 1964).

2.2. Charge resonance states


Charge resonance (CR) states are due to Coulombic interaction between positive negative (ZM-) molecular ions. I n the absence of orbital overlap between the two molecules, the four CR states are degenerate with a common energy for the (M 11 M) excimer of 1, 3(R)g,u = I - A - C (r) (2 6 )
(2M+) and

where I is the molecular ionization potential, A is the electron affinity, and C ( r ) is the Coulombic interaction potential which is approximately

C (r)=eZ/er
where e is the electronic charge, and
E

(2 * 7)

is the dielectric constant. The zero-order

Excimers

913

wavefunctions of the singlet CR states of the (M 11 M) excimer are (Murrell and Tanaka 1964) $ (1(R)g)=1($1(2M+) $z(~M-) + $I(~M-)$z(~M+)) (2.8a)

$ ( l ( R ) u ) = '($I(~M+)$J~(~M-) - $I(~M-)$Q(~M+))*
2.3. Nature .f the aromatic excimer state

(2.8b)

On the Platt (1949) model of the cata-condensed hydrocarbons, the square mp2 of the transition moment from SO= 1A to S, = 'Lb, 1L, and lB,, b increases approximately in the ratio of 1 : 10 : 100 respectively. 'Lb and 1L, correspond to the SI and/or S2 states, and 1B, and lBb are higher-energy states (Birks 1970a, 1973). These considerations led Forster (1962, 1963) to propose three conditions for excimer formation by an aromatic molecule : (i) SI= 'Lb, to give a weak transition moment ml, a long S1 radiative lifetime, and hence a long 1M" lifetime to facilitate process (11); (ii) S2 = 1L,, to give a relatively strong transition moment ma; and (iii) a small S2-S1 energy gap 4 1 2 . Forster equated the excimer state to the 1L, exciton resonance state 1(E2)g, which lies below the energy S! when m22lr3 > 4 1 2 , ie when
Y

< (m22/612)1/3.

(2.9)

Many aromatic hydrocarbons satisfy (i) and (ii), and those that also meet condition (iii) generally form excimers. There are, however, several groups of aromatic compounds which exhibit excimer fluorescence without satisfying Forster's conditions. They include the following: (a) benzene and its alkyl derivatives, toluene, o-xylene, m-xylene, p-xylene, mesitylene, ethyl benzene and isopropyl benzene (Ivanova et al 1962, Birks et al 1965, Lumb and Weyll967) ; (b) 9-methylanthracene, 9,lO-dimethylanthracene and other meso-substituted (9-acetoxy, 9-methoxy, 9-n-propyl, 9-carboxylic acid, 9,10-di-n-propyl, 9-methyl 10-methoxy) derivatives of anthracene (Birks and Aladekomo 1963, Bazilevskaya and Cherkasov 1965a,b, Cherkasov and Bazilevskaya 1965, Barnes and Birks 1966); (c) perylene in the crystal (Tanaka 1963), solution (Birks and Christophorou 1964) or liquid (J Langelaar and G J Hoytink 1968, private communication) phases; and ( d ) 2,5-diphenyloxazole (Berlman 1961) and other oxazoles and oxadiazoles (Lami and Laustriat 1968, Horrocks 1969). Group (a) satisfy conditions (i) and (ii), but not (iii), since 4 1 2 is too large for (2.9) to be satisfied at any reasonable value of r. Groups (b) and (c) have SI = 1L,, in contradiction to conditions (i) and (ii), Group ( d ) has an S1 radiative lifetime in the nanosecond range, equivalent to SI= 1B. As an alternative to the exciton resonance model of the excimer, Ferguson (1958) and Slifkin (1963) proposed that the excimer is a charge resonance state. Calculations by Konijnenberg (1963) and Azumi and McGlynn (1964) showed both models to be inadequate to explain the spectroscopic data on aromatic excimers. This is not surprising, since the ER and CR models each neglect orbital overlap between the two molecules. They can therefore only properly be applied at values of Y exceeding the excimer separation distance. As pointed out by Murrell and Tanaka (1964), the repulsive forces which keep two molecules apart only come into play when there is

914

J ' B Birks

some overlap of their occupied orbitals. This means that in the stable excimer conformation, in which the forces of attraction and repulsion are just in balance, overlap effects cannot be neglected.

2.4. Conjigurational mixing


T o allow for the orbital overlap of the two molecules, Konijnenberg (1963) introduced configurational mixingt of the ER and CR states in calculations of the naphthalene and pyrene excimers. Each ER state '(E) is mixed with the CR state '(R) of the same symmetry species to yield a pair of singlet excimer states '(E T R) of which the '(E - R) state has the lower energy. Azumi et a1 (1964) and Murre11 and Tanaka (1964) have made configurational mixing calculations of the naphthalene, anthracene, pyrene and perylene excimers. Each molecule belongs to the D2h symmetry point group. T h e excimer conformation is taken as an overlapping sandwich dimer (M IlM) ,which also belongs to the D2h point group. T h e symmetry of the 'Lb molecular state is lBsU, and its ER states are 1BZg and lBgu. T h e symmetry of the 'La molecular state is lBzU,and its ER states are 'Bag and lBzU. The lowest-energy singlet CR states also have 1B3, and lBzu symmetries, and configurational mixing of these with the ER ('La) states yields the ~ B G , 1BZ, IBZ', and 'B& excimer states in order of increasing energy (Azumi et a1 1964). There is no mixing of the ER ('Lb) and lowest CR states because of their different symmetries. T h e ER ('La) and lowest CR states belong to the same symmetry species for any conformation of two parallel molecules, but the ER (lLb) states belong to different 'non-mixing' symmetry species in all the more symmetrical D2h, C2h and cg, dimer conformations (Azumi and McGlynn 1965). For each molecule considered by these authors ID" is equated to 'BG, the lower ER ('La) state stabilized by configurational mixing with the corresponding CR state. This assignment is likely for anthracene and perylene, where S1=lLa, and for naphthalene, where 1=1Lb, Sz='La, and A12 is small. Pyrene also satisfies the latter conditions, but the failure to take account of the lower ER ('Ba) state of 1B3, symmetry represents a possible source of error (Birks and Kazzaz 1968) to be discussed in 53.3. benzene molecule are lBzU, lBlu and 'Elu T h e 'Lb, 'La and 'B states of the respectively. For the symmetrical D6h(M IIM) benzene excimer, the ER ('Lb) states are lBlg and lBzU,the ER ('La) states are 1Bzg and lBIU, the ER (1B) states are 'Elg and 'ElU, and the CR states are lBlg, 1BIU, lBzg, lBzU, ' E l u and 'El, (Azumi and McGlynn 1965). Thus configurational mixing can occur between each of the ER states and the corresponding CR state of the same symmetry. T h e calculated energies of the '(E- R) states of the benzene Deh excimer as a function of interplanar separation Y are plotted in figure 3 (Vala et a1 1966). At r >2.8 A the order of the benzene singlet excimer states is the same as that of the molecular states from which they originate: Big, Bzu (from lBzu), B2g, B1, (from lBlu), Elg, Elu (from 'Elu). T h e benzene excimer fluorescence maximum of Dm=31 300 cm-1 (Birks et a1 1965) corresponds to a ID* equilibrium separation of Ym= 3.3 A. It is concluded that the lowest singlet excimer state ID* of benzene corresponds to the lower ER ('Lb) state lBlg, strongly stabilized by configurational mixing with the CR 1B1, state.

t I am grateful to Professor R S Mulliken for suggesting 'mixing' to replace the conventional term 'interaction', which is ambiguous in the present context.

Excimers

915

Similar results have been obtained by other theoretical methods. Chesnut et a1 (1965) have applied an extended Huckel treatment to the benzene excimer, and the same method has been used for other excimers by Polak and Paldus (1966). Azumi and Azumi (1967) and Chandra and Lim (1968) have used simple Huckel-type dimer orbitals to treat the Dah excimers. The results have been shown to be equivalent to the configurational mixing treatment for the D2a naphthalene excimer (Chandra and Lim 1969).

2.5

3.0

3.5

4.0

Ring separation

(11

Figure 3. Singlet states of the benzene excher. Theoretical energies of the Dah sandwich dimer as a function of interplanar separation r. Low-energy excimer states of 1Bzu, lBlu and lEiu molecular parentage (after Vala et aZ1966).

3. Excimer potentials
3.1. The potential diagram Figure 4 is a schematic potential diagram of the ground state (1M+1M) and first excited singlet state (IN"+ 1 M S D " ) of a pair of molecules in an (M jlM) sandwich dimer conformation as a function of the intermolecular distance r. At r = CO where there is no molecular interaction, the ground- and excited-state energies are zero and MO( =8 : ) respectively. The ground-state potential is the resultant of a longrange attractive van der Waals potential A (r),which may be neglected at small r, and a short-range repulsive potential R ( r ) . There is a similar repulsive potential R'(r) in the excited state. The nuclear and electronic coordinates of the pair of aromatic molecules are practically identical in

916

J B Birks

the ground and excited states, apart from those of one excited 7~ electron, and hence

R'(r)=R(r). (3 * 1) This approximation does not necessarily apply to noble gas excimers. There is also an attractive excimer interaction potential V'(r)in the excited state, corresponding to the ID" energy considered in 92, where R'(r) was ignored. T h e total ID" energy is thus D'(r) = V'(r)-IR'(r).

r-

0 ) '

Figure 4. Schematic diagram of the potential energy W of a pair of parallel molecules 1M* and 1M as a function of the intermolecular separation r . R, R' are repulsive potentials in the ground (1M) and excited (lM*) states. V' is the excimer interaction potential. D'(= V'+R') is the resultant excimer energy. M O is the molecular 0-0
R,, RL) V L and DL are potentials at the equilibrium excimer separation
the excimer binding energy (after Birks 1970a). fluorescence transition.

D, is the peak of the excimer fluorescence transition.


Y,.

B is

T h e system forms a bound excimer state, provided that the parameter

B (r)= MO - D'(r) > 0


and has apositive maximum of

(3.3)

There are two groups of aromatic hydrocarbons which do not form excimers or photodimers, and which do not exhibit fluorescence self-quenching in fluid solution. T h e first group includes unsubstituted compounds like phenanthrene and chrysene in which there is no steric hindrance to excimer formation. I n these molecules SI= 1Lb,

corresponding to the excimer binding energy B at the ID" equilibrium separation of r=rm. Dm is the energy of the ID" fluorescence maximum, and Rm, Rk, Vk and Dk are the potentials at r = rm.

Excimers

917

but the splitting of any higher molecular state S p due to exciton and/or charge resonance is insufficient to span the S& energy gap Alp, so that V ( r )and hence D'(r) remain too large to satisfy (3.3). The second group includes molecules like 9,lOdiphenylanthracene in which there is steric hindrance to the close approach of 1M" and 1M, so that R'(r) and hence D'(r) are too large to satisfy (3.3).

w c

Figure 5. The helium excimer. Potential energy curves (without rotation) for selected states of Hez. The energy in cm-1 is expressed relative to the N = 0, v = 0 level of the 3 Z : state. The energy in eV is expressed relative to the lowest level of the x 2 Z t state of He$. As far as possible the observed vibrational levels are indicated by short horizontal lines at the edges of the appropriate curves (Ginter 1970a) [reproduced by
permission from Rosen (1970)l.

3.2. The helium excimer


The many calculations of different states of the helium excimer include those of Mulliken (1932, 1964a,b, 1966, 1969)) Buckingham and Dalgarno (1952), Brigman et a1 (1961)) Poshusta and Matsen (1963), Browne (1965), Scott et al (1966)) Gupta and Matsen (1969) and Guberman and Goddard (1972). The theoretical and related spectroscopic data have been collected by Ginter (1970a). The ground-state potential of He2 is mainly repulsive with a shallow van der Waals minimum of 0.0009 eV at r = 2.9 A (Buckingham 1958). The helium excimer states are bound with respect to the energies of the separated atoms, and their potential curves have minima at r m = 1.0-1.1 A. There are sufficient theoretical and experimental data to enable reasonably precise potential energy curves to be constructed for many states of the helium excimer. Figure 5 shows a set of such curves constructed

918

J B Birh

by Ginter (1970a), in which the estimated errors of different portions of the curves are indicated. From theoretical considerations and spectroscopic data, Mulliken (1964a) concluded that (a) there are no purely repulsive states of the helium excimer, other than the ground state, below the energy of He++ He, and (b) many of the helium excimer states have high maxima in their potential curves at r > rm. Most of the potential maxima are attributed to the quantum mechanical noncrossing rule, which forbids the intersection of potential curves of different electronic states of the same symmetry species (Mulliken 1964a,b, 1966). Subsequent calculations, including those of Browne (1965), Gupta and Matsen(1969) and Guberman and Goddard (1972), have confirmed the existence and the interpretation of the potential maxima, some of which are shown in figure 5.

3.3. Thepyrene crystal dimer


The pyrene crystal is dimeric in structure, with the molecules arranged in (M I/M) sandwich pairs which form the elementary lattice units. The molecules of the unexcited dimer have their planes and axes parallel at an interplanar distance of ro = 3.53 8. When one molecule is projected on the plane of the other, the long molecular axes coincide, but the short molecular axes are separated by one C-C bond length (figure 6). This arrangement maintains a high degree of symmetry (Czh), and minimizes the intermolecular repulsion, since no C-C bonds are opposite to each other (Birks et aZ1966a).

Figure 6 .The pyrene crystal dimer. Projection of one molecule on the plane of a parallel neighbour. Interplanar distance ro = 3 m53 A.

The pyrene crystal absorption spectrum is structured, corresponding to the M +1M*. However, because Franck-Condon excitation of individual molecules 1 TO is less than the range of the attractive excimer interaction, an excimer ID* is rapidly formed by 1M*and its unexcited neighbour lM, so that the crystal fluorescence spectrum consists solely of excimer emission. Figure 7 shows the fluorescence

Excimers

919

4K

'0

I O-

D m

Yl

I 1.20

Figure 8. Experimental potential energy diagram of the pyrene crystal dimer and excimer. Energy W against intermolecular separation Y and displacement q from excimer equilibrium. Notation as for figure 4 (after Birks and Kazzaz 1968). 61

920

J B Birks

spectrum of a single pyrene crystal as a function of temperature. These and other data have been analysed to evaluate the potential diagram of the pyrene crystal dimer shown in figure 8 (Birks and Kazzaz 1967, 1968). The energy W i s plotted as a function of r, the intermolecular spacing, and of q ( = r - r m ) , where rm is the equilibrium spacing of the excimer, and q is the displacement from this equilibrium. The potential diagram is analogous to that of a luminescence centre in an ionic crystal, and the relations derived for the latter by Williams and Hebb (1951) can be applied. Treating the excimer as a harmonic oscillator along the q axis, the classical expression for the fluorescence spectrum is (3 * 6 ) where E is the photon energy at displacement q, k is the vibrational force constant of the excimer, and the exponential factor is the probability of the system having a displacement q. I n a more exact treatment the classical distribution is replaced by the sum over the quantized vibrational levels of the excimer

P ( E )= (k/27rkT)1/2 exp (- kq2/2kT) dq/dE

p(d=C i

( + i ( d ) 2 exp

[-(2i+W/TI

(3.7)

where k 8( = &AV) is the zero-point energy of the harmonic oscillator of frequency U, and $&) is the normalized wavefunction of its i t h vibrational state. Replacement of the exponential distribution in (3.6) by p ( q ) and renormalization is equivalent to replacing T by an effective temperature

Te = 8 coth (8/T)
so that

(3.8)

(3.9) The vibrational levels of the ground-state dimer are also quantized, but the latter can be closely approximated to a classical harmonic oscillator, provided the displacement qo( = r o - r m ) of the excimer and ground-state dimer potential minima is large, so that the fluorescence occurs into high vibrational levels of the ground state. The pyrene crystal excimer satisfies these conditions. The fluorescence spectrum at 0 K is given by

P ( E ) = (k/2~kTe)1/2 exp ( -Kq2/2kTe) dq/dE.

p o ( q q $ o ( q ) ) 2dq/dE=(+)l/2

exp (-

4) dq/dE

(3.10)

where +&) is the zero-point vibrational wavefunction of the excimer, and ol=K/Au. From (3.9) the full width at half-maximum of P ( E ) at temperature T is

AP (E) = APo(E) [coth (e/T)]1/2.

(3.11)

These relations have been used to analyse the pyrene crystal fluorescence spectra (figure 7). The parameters MO( = 26 600 cm-1) and Dm( = 20 900 cm-1) are obtained from the crystal absorption and fluorescence spectra. The excimer binding energy B (=3230 cm-I), obtained from solution data, is assumed to be the same in the crystal phase. Hence we obtain Dk(=MO - B = 23 370 cm-1) and Rm( = DL - Dm = 2470 cm-1). Comparison of the fluorescence spectral data with (3.11) gives the zeropoint energy KO( = 90 cm-I), vibrational frequency U( = 5.4 x 1012 Hz), and force constant k( = 1.93 x 105 dyn cm-1) of the pyrene crystal excimer, and hence the excimer potential

D(r)= Dk + ikp2.

(3.12)

Excimers

921

Comparison of the observed fluorescence spectrum at 4 K, which approximates closely to Po(E), with (#o(q))2 using (3.10) gives the transition energy E and hence R ( = D& - E ) as functions of q. For a harmonic oscillator

R (r)= Qk (YO - r)2 = +k (40- 4)'

(3.13)

where k is the vibrational force constant of the ground-state dimer, and q=qo at r=ro. The experimental values of against q give a linear plot, consistent with (3.13), extrapolating to R=O at q0=0.16 A, corresponding to rm-3.37 A for the equilibrium separation of the pyrene crystal excimer. The ground-state dimer has a zero-point energy k0 (= 135 cm-I), vibrational frequency v( = 8.1 x 1012 Hz), and force constant k ( = 4.2 x 105 dyn cm-1). The experimental pyrene excimer interaction potential V'(r) (=D'(r)- R ( r ) ) does not agree with the calculations (92.4) of the pyrene ER (lLB)state stabilized by mixing with the corresponding CR state (Konijnenberg 1963, Azumi et al 1964, Azumi and Azumi 1967, Murre11 and Tanaka 1964). It is instead found that V'(r)= S E - mp2/r3 (3.14) with Sg=51 330 cm-1 and mp= 15 D (Birks and Kazzaz 1968). This is consistent E and transition with the lower ER state originating from a molecular state of energy S moment mp,and the 1B, state of pyrene appears to satisfy these criteria. The simple assignment of the pyrene excimer to the lower ER (lBB) state of 1B3, symmetry is, however, unsatisfactory, since it violates the non-crossing rule ( 93.2). The pyrene ID* state probably results from configurational mixing of the ER ('Ba), ER ('La) and CR states of 1B3, symmetry. Relation (3.14) indicates that the ER (lBB) state is dominant in determining V'(r) at small r. The pyrene crystal excimer fluorescence is predominantly polarized along the molecular a axis, consistent with the above assignment (Hochstrasser and Malliaris 1965, Chaudhuri and Ganguly 1970).

4. Excimer reaction kinetics


4.1. Photostationary kinetics I n the notation of Birks et aZ(1963b) the rate parameter describes the B c A process, where A is the initial excited species and B is the product species or radiation, is the sum of those of all unimolecular processes, The rate parameter K A = other than dissociation, operating on A. The (A, B) subscript notation is as follows: M=lM*; D = ID*; T=3M*; N=3D*; F = fluorescence ; P = phosphorescence. G = 1 M or 1M+1M;
Square brackets denote molar concentrations. Figure 9 shows the rate processes and parameters involved in excimer association and dissociation in solution. The latter processes are indicated by horizontal arrows, The competing radiative and radiationless processes are indicated by solid and broken downward arrows respectively. ~ D M [ I Mand ] kMD are the rates of ID* formation and dissociation respectively, and (4.2) are the total rates of 1M* and ID* decay by other means respectively. KNT[~M] and
'

AM =JZFM + ~ T + M GM k~ =~ F + DAND + ~ G D

(4.1)

922
~ T are N

J B Birks
the rates of 3D+ formation and dissociation respectively, and

KT =J ~ P T +~ G T kN = k m 4-kGN

(4.3) (4.4)

are the total rates of 3M+ and 3D+ decay by other means respectively. At infinite dilution the 1M" lifetime T M = l / k ~and , the 1M+quantum efficiencies of fluorescence, inter-system crossing and internal conversion are QFM =~ F M / ~ M , QTM = ~ T M / ~and M QGM = ~ G M / ~ respectively. M By analogy we define at infinite concen, the ID+ quantum efficiencies of fluorescence, tration the ID+ lifetime T D = 1 / k ~and QND , = ~ N D / ~ and D inter-system crossing and internal conversion as QFD = ~ F D / ~ D ~ G= DR G D / ~ D respectively.

+ $
'M f hu,
'M 'M thu,
2'M t hu,,

2'M

Z'Mthv,

Figure 9. Molecular and excimer rate processes and parameters.

For steady excitation with light of intensity lo einsteins litre-1 s-1, only 1M+ is initially excited, and the rate equations are d['M+]/dt=lo- ( ~ M ~ ~ D M [ [lM"] ~ M+ ]~ ) MD[~D*] d [1D"]/dt = KDM [1M] [1M+]- (AD + ~ M D [1D+]. )

(4.5) (4 6)

Under photostationary conditions d [lM+]/dt= d [lD+]/dt = 0, and the following relations are obtained. The 'effective' molar equilibrium constant of process (11) is Ke=[lD"]/[lM+] [ ~ M ] = ~ D M / ( ~ D + ~ M D ) . The 1M+ and ID+ fluorescence quantum yields are respectively

(4.7) (4 * 8)
(4.9)

@ F M = ~ FS M K /[ (l~ M])=k~x/P
where
@FD=qFD/(1 f

1/K [ ~ M ] > = ~ Ke F D[lM]/P

K= l / [ l M ] h = ( A ~ / kKe ~)

(4 * 10)

is the Stern-Volmer self-quenching coefficient, its reciprocal [lM]h is the 'half-value' concentration at which OFM = $qFM and @)FD= frqFD, and

P = ~ 4MkD&

['MI

(4.11)

Excimers

923

is the 'weighted-mean' decay parameter of the system. The ratio of the fluorescence yields is
@FD/@FM = qFD[lM]/qFM[ l M ] h = k FD Ke [lM]/k FM = K1 ['lvr].

(4.12) (4.13)

Elimination of K [1M] from (4.8) and (4.9) gives


@FM/QFM+ @FD/QFD=

1.

This relation has been used by Hirayama and Lipsky (1969b), who observed linear plots of @FD against @FM for benzene and toluene solutions at constant temperature from their intercepts. and variable [IM], and evaluated QFM and ~ F D , [lM]h and K1 can be obtained from fluorescence The parameters ~ F M ,~ F DK, quantum yield (@FM, @ED) measurements as a function of [1M] (Doller and Forster 1962a,b, Birks et aZ1963b, 1964c,d, 1965). Fluorescent lifetime (transient) observations are required to determine the individual rate parameters (see $4.2). The 1M" and ID+ quantum yields of inter-system crossing to 3M+ and 3Dx are respectively (4.14) @TM = ~ T M /1 ( K [lM]) = K T M / ~ @ N D = Q N D / (1/K ~ [~M])=~N Ke D['MI/&

(4.15)

The total triplet quantum yield of 3M+ and 3DB, which normally dissociates rapidly (KTNBKN) into 3M" and 1M, is The 1M+ and ID* quantum yields of internal conversion to 1M and 1M + 1M are respectively (4.17) @GM =QGM/( 1+ K [lM]) = KGM/S
@ G D = ~ G D / ( ~1/K [ ~ M ] ) = ~ G Ke[lM]/F. D

@T= @TM

+ @ND=

(~TM

+AND Ke [lM])/P.

(4.16)

(4.18)

Observations of the concentration dependence of QFM, @)FD(Parker and Hatchard 1963) and @T (Medinger and Wilkinson 1966, Heinzelmann and Labhart 1969) for pyrene and 1:2-benzanthracene solutions have been analysed to evaluate QFM, ~ T M , GM (EO), ~ F D ,qND and ~ G D (Birks 1970a, Birks et al 1971). Observations of K M and KD then enable all the individual rate parameters of (4.1) and (4.2) to be determined.

4.2. Transient kinetics


Excitation of a solution system by a 6 function light flash at t=O produces an of excited molecules. At a time t after the excitation the initial concentration [~M"]o rate equations are m [ID"] - X [IM"] (4.19) d [1M"]/dt = k d [1D"]/dt where

DM [1M] ['M"] - Y [ID"]

(4.20) (4.21) (4.22)

X=KM+KDM [lM]
Y = h + km.

From (4.19) and (4,20), applying the initial conditions [IM"] = [1M"]o, [ID"] = 0,

924

J B Bivks

at t=O, we obtain (Birks et al 1963b) the 1M* fluorescence response function

~M(~)=~FM[~M']/[~M*]O=[~FM(A~-X)/(XZAI)] [exp (- h t ) + A exp (- h t ) ]


and the ID* fluorescence response function

(4.23)

~D(~)=KFD[~D*]/[~M*]O=[~FD~DM[~M]/(X~X I ) ] [exp (- hlt)-exp (-ht)]


where
Ai,z=&

(4.24)
(4.25) (4.26)

{X+Y T [ ( Y - X ) ' + ~ ~ M kDM D [1M]]1'2}


A = (XAI)/(
A2

and

-X).

The relative fluorescence response functions can be determined by pulse fluorometry, phase and modulation fluorometry, or photon sampling fluorometry (Birks et aZ1963b, 196413, Birks and Munro 1967), and the parameters hi, A2 and A evaluated therefrom. Two methods have been used to determine the four rate parameters

u;ol

io7
5 -

IO8
5

10'
5
2

L M I (mole lid1

Figure 10. Pyrene solutions in cyclohexane. Double logarithmic plots of decay parameters A 1 and A2 against molar concentration [MI at temperature T : x experimental

data from molecular fluorescence; 0 experimental data from excimer fluorescence. Theoretical curves from relation (4.25) with appropriate choice of parameters E 1963b). (after Birks et a

kM, kD, kDM and kMD. In the first method A 1 and A2 are determined as a function of [IM], and theoretical curves from ( 4 . 2 5 ) with suitably chosen rate parameter values
are fitted to the experimental data. Figure 10 shows such data for solutions of pyrene Z 1963b). The following limiting in cyclohexane at various temperatures (Birks et a properties of A1 and A2 aid the curve-fitting.

Excimers

925

A second method of evaluating the rate parameters involves measurements at only two concentrations (Birks et a Z 1964b). hl( =AM) is observed at low [IM], and X ,I A2 and A are obtained from iM(t) and iD(t)at a single high concentration [IM]. The results are then analysed using successively the following relations obtained from (4.21), (4.22), (4.25) and (4.26): (i) (ii) (iii) (iV) (v)

X= (Ah2 + AI)/(A 1) ~DM['M] = X- k~


~MD=(X XI) - (h2-X)IkDM ['MI KD = Y - KMD.
Y=h1+hz-X

The parameters AM, k ~ DM , and k m depend on the nature of the solute and solvent and on the temperature, but they are generally assumed to be independent of [IM]. This assumption appears to be valid except at very high values of [IM] > 1 M, where the rate parameters may be influenced by changes in the viscosity, refractive index and other properties of the solvent (Aladekomo and Birks 1965, Birks and King 1966). The parameters ~ F M T , M, ~ F D and TD for concentrated solutions and pure liquids may therefore differ from those for more dilute solutions.
N

4.3. High-temperature kinetics


T h e kinetic relations of $54.1 and 4.2 are considerably simplified in the 'hightemperature' or dynamic equilibrium region (Birks et aZ1965), where

~DM[IM], ~ M D B KD. ~M,

(4.27)

This condition is satisfied by all solutions above some critical temperature Tc, which depends on the excimer binding energy and the solvent viscosity. For most compounds in low-viscosity solvents Tc is below room temperature. For pyrene Tois at or above room temperature (Birks et al 1964c,d). Above Tc (4.7) becomes

Ke+ (Ke)o=KDM/kMD

(4.28)

where (Ke)o is the true molar equilibrium constant of process (11), under conditions where the competing ID* dissociation processes represented by KD are negligible. The dynamic equilibrium conditions (4.27) simplify the fluorescence response functions &(t) and iD(t). The parameter
A2 = ~ M D

+ KDM [lM]

is very large, and exp ( - A$) is negligible, except at very small t. Equations (4.23) and (4.24)become (4.29) iM(t)=mkFM exp ( - At)
i D ( t ) =d

k exp ~ (~ At)

(4.30) (4.31) (4.32) (4.33)

where

h = mkM + dkD

m=l/(l+(Ke)o [lM]) d=1-m.

926

J B Birks

The 1M' and ID" fluorescences decay exponentially with a common lifetime T ( = l/X), At [1M] = [lM]h (4.34) T = Th = 4 ( T M f T D ) a relation used to determine the temperature dependence of
et aZl968).
Q

for toluene (Greenleaf

The parameters m and d are the fractions of excited molecules in the 1M" and ID* states respectively (Birks and Conte 1968). From (4.31) the total decay rate X of the excited system is the weighted mean of the 1M" and ID* decay rates, K M and K D respectively. The mean rate parameter k x of a specific photophysical process of the excited system-eg internal conversion, inter-system crossing, impurity quenching, energy transfer to another molecular species-is similarly the weighted mean of m respectively, so that the corresponding 1M* and ID* rate parameters, K X M and k

kx = mkxM -k dkxD.

(4.35)

Under dynamic equilibrium conditions the reaction kinetics of the system can be treated as if only one excited species were present with mean rate parameters defined by(4.35).

4.4. Thermodynamics
From the general thermodynamic relations for process (11), the change in free energy AG= -RTIn(Ke)o=AH-TAS (4.36) where R is the gas constant, and AH and A S are the enthalpy and entropy changes in the reaction respectively. At T > Tc we obtain from (4.12), (4.18) and (4.36) that

K~=(DFD/~FM[~M]=(KFD/RFM) ( K ~ ) ~ = ( ~ F D /exp ~ F (AS/R) M) exp (- AH/RT).


(4.37)

R F D / ~ F M for pyrene has been evaluated from fluorescence lifetime and quantum yield measurements, and it has been found to be independent of T and of the solvent (Birks et aZ1964d). Observations of K1 as a function of T at T > Tc enable AH to be determined from the gradient of In K1 against 1/T, and if K F D / ~ F M is known, A S can also be evaluated. Such observations on pyrene (Birks et aZ 1964d), 9-methylanthracene and 9,lO-dimethylanthracene (Barnes and Birks 1966) in various solvents show that K F D / K F M , AH and A S are all independent of T and of the solvent. The experimental determination of the rate parameters ($4.2) as a function of T enables the frequency factors and activation energies of the various rate processes to depends on the be determined. The rate parameter of molecular fluorescence ~ F M solvent refractive index n, JZFM = n2 (KFM)O (4.38)
but it is otherwise normally independent of T (see $7.2). The observations (Birks et aZ 1964d) that K F D / K F M is independent of T and of the solvent indicate that JZFD obeys a similar relation (see $97.1 and 7.2),
~ F = Dn2( ~ F D ) o .

(4 39)
9

Excimers
The 1M" internal quenching rate K I M ( = KTM

927

where K k is the temperature-independent component, R h and WIM are the frequency factor and activation energy respectively of the temperature-dependent component A&, and K is Boltzmann's constant. Similarly the ID" internal quenching rate AID( =k m KGD) obeys the relation

KIM = k&

+K&

= K!M

+K;,

+AGM)

obeys the relation

exp ( - WIM/KT)

(4.40)

km = k&

+K& =k!D +K;D exp ( - Wm/KT )

(4.41)

where is temperature-independent, and KfD is temperature-dependent with a frequency factor kiD and activation energy WID. The relation of k&, and kyD to the alternative radiationless decay modes, KND and KGD, is discussed in $7.1. The excimer formation (KDM) and dissociation ( K M D ) rate parameters are temperature-dependent, DM = exp ( - WDM/KT) (4.42)
KMD =

exp ( - WMD/kT). exp [(WMD- wDM)/kT]*

(4.43) (4.44) (4.45) (4.46)

Their ratio is
KDM/k~=(K,)O=(K~,/kLD)

Comparison with (4.37)shows that


K L M / k L D = exp

(AS/R)

WM D W D M = B =- AH/N where B is the excimer binding energy, and N is Avogadro's number.


4.5. Diffusion-controlledprocesses

Excimer formation in solution is a diffusion-controlled collision process, the rate parameter of which is generally described by the Sveshnikov (1935, 1937) relation

KDM=~TN'D$R [1 + R ( ~ T M ) - ~ /(M-l ~]

S-l)

(4 * 47)

derived from Einstein-Smoluchowski diffusion theory, where N' =N x 10-3, D = D M +DG is the sum of the diffusion coefficients DMand DG of the reactants 1M" and 1M respectively, R = T M + Y G is the sum of their interaction radii T M and TG, and p ( < 1 ) is the reaction probability per collision. Alternative forms of (4.47) have been discussed by Noyes (1961) and Alwattar et uZ(1973). The term in square brackets in (4.47) is usually close to unity, so that the relation then simplifies to
kDM = 4rrN'DpR.

(4.48)

Direct observations of molecular diffusion coefficients are sparse, and hence the Stokes-Einstein relations

DM= kT/6rrqa~,

DG=k T / 6 n q a ~

(4 * 49)

where U M and UG are the molecular radii of 1M" and 1M, and q is the solvent viscosity, are often used to evaluate DMand DG. If it is further assumed that uM = aG = r M = YG, then substitution of (4.49) in (4.48) gives
kDM=

8RTp/3000q.

(4.50)

928

J B Birks

The experimental values of KDM for cyclohexane solutions of pyrene, 1:2benzanthracene and three of its methyl derivatives at 20 "C agree with (4.50) for values of p between 0.4 and 1.0 (Birks et a1 1963b, 1964b). The activation energy WDMof excimer formation by pyrene in six different solvents agrees satisfactorily with the activation energy W, of viscous diffusion, evaluated from the temperature dependence of T/q (Birks et a2 1964d). Forster et a2 (1963) have studied the molar fluorescence yield ratio K1 (see equation 1.1) of pyrene and 1:2-benzanthracene solutions as a function of applied pressure P. For pyrene in toluene solution at 20 "C, where Kl is approximately proportional to ~ D M K1 , decreases by a factor of 10 as P is increased from 1 to 5000 atm, in a similar manner to the decrease of l/v. The various studies show that KDM is diffusion-controlled and related to 1/77, but they also indicate that the Stokes-Einstein relation (4.49) is only approximately valid. Thus for pyrene in a series of solvents at 20 "C Doller and Forster (1962b) observed that kDMcCrj-0'55. (4.51) The analysis was extended to include further solvents by Alwattar et a1 (1973), who obtained J ~ D M ~A' / T+ = b'qo.6 (4.52) in place of the constant value predicted by (4.50). They also discussed the probable reasons for the deviations from Stokes-Einstein behaviour.

5. Different modes of excimer formation


Excimers can be found by various processes other than the direct interaction of excited (1M* or 3M*) and unexcited (1M) molecules by processes (11) or (IV).

5.1, Dimer cation neutralization


The interaction of a molecular cation 2M+ with a neutral molecule yields a metastable dimer cation 2Df: 2M++ 1M -+ 2D+. (VW Such dimer cations are formed in ionized noble gases (Herzberg 1950). Their formation in y-irradiated solutions of benzene, naphthalene, anthracene and other aromatic hydrocarbons in low-temperature glasses has been observed by electron spin resonance (Lewis and Stringer 1965, Howarth and Fraenkel 1966, 1970) and by optical absorption (Badger et a1 1967, Badger and Brocklehurst 1968, 1969a,b). The excimer fluorescence bands observed in the thermoluminescence or infrared-stimulated luminescence spectra of concentrated glassy solutions of naphthalene and other aromatic hydrocarbons irradiated at 77 K (Badger et a1 1967, Badger and Brocklehurst 1968, 1969a,b, Brocklehurst and Russell 1967,1969) are attributed to the neutralization of 2D+ by thermally released electrons, 2D++ 2e-f ID" L 3D+ leading to ID+ fluorescence via process (VIII). The quantum yields (DFM and (Dm of the 1M* and ID+ fluorescences of a liquid alkyl benzene at room temperature depend on the mode of excitation. With ultra-

Excimers

929

violet (uv) excitation > (Ivanova et al 1962, Birks et al 1965, Lumb and and Weyl 1967). Decrease of the excitation wavelength he, causes a decrease in , :@ : but @ $ /@ ;:: is independent of hex for A e x a 190 nm (Birks et al1968b). With 2 @; :/@$$, and @iD/@iM increases with the ionizaionizing (I) irradiation @iD/@iM tion density (Carter et a1 1967, Christophorou et al1968). With low-intensity ionizing (LI) irradiation @$:/@;if= @ ~ ~ (Birks / @ et~ al ~ 1964a). With high-intensity (HI) electron beam irradiation @FA/@:&$ ;@ ;/ :@ ; and OF&. An explanation of these effects has been proposed (Birks 1970c) in terms of processes (VII) and (VIII), the ionization quenching of lMx (Birks 1951, 1964a), and the reaction kinetics of Ss4.1 and 4.3. For initial excitation of 1M" (uv)

@ ; A $

[Q]) ( 5 * 1) where kQD is the rate parameter of ID" quenching by a quenching species of molar concentration [Q]. With excitation of the liquid by ionizing radiation (I), the high molar concentration [1M] favours process (VII), leading to ID" formation by process (VIII). For initial excitation of ID"
FD/

@UV @UV

F M = ~ F D ~ D ['M]/~FM M (KMD+~D+~QD

@~D/@~M=~D(~D +Ki\.I+ M [ ~ kQM[Q])/kFMKMD M] (5.2) where KQM is the rate parameter of 1M* quenching by [Q]. For liquid alkyl benzenes at room temperature, KFM N KFD 2: 106 s-1, K M 2: KD N 2 x 107 S-1, KDM 2: 1010 M-1 S-1, K M D 2: 1011 S-1, KQM, KQD < 1010 M-1 s-1, and [1M]2: 10 M (Birks et al 1965, Birks and Conte 1968, Greenleaf et a1 1968, Gregory and Helman 1972). Hence, provided [Q] Q [IM], ie with uv or LI irradiation,
N

KDM[~M], ~ M D B ~ M [Ql, + ~ ~Q DM + & D [QI and (5.1) and (5.2) reduce to the common relation

(5.3) (5.4)

@ $ /@ ;::

@g:/@gg =~

F D DM

[lM]/kr~ ~

M D

irrespective of whether 1M" or ID" is initially excited. Under HI irradiation a high local concentration [Q] of ionic species is produced, and the local concentration [1M] of unexcited molecules is reduced from [ ~ M ] o towards zero due to 2M+ and 2D+ formation. Under these conditions ([Q] $ [IM]) (5.2) becomes (5.5) Experimentally @FA/@F&B @$:/@&, and hence from ( 5 . 5 ) and (5.4) KQM[Q] B ~ D M [ ~ M assuming ]o, that the other rate parameters are not modified in the local ionic environment. Ionization quenching of 1M" with a rate parameter (KQM) faster than that of a diffusion-controlled process ( K D M ) has been previously proposed to account for the dependence of the scintillation efficiency of an organic scintillator on the specific energy loss of the ionizing particle (Birks 1951, 1964a). Under HI irradiation the IDx dissociation is practically irreversible, because of the strong ionization quenching of 1M" and the high depletion of [IM]. Under uv and LI irradiation the 1M" and ID" fluorescences decay with a common lifetime 72120-40 ns (Birks et al 1965, Greenleaf et al 1968, Wagner et al 1969). Under HI irradiation the IDx fluorescence has a predicted lifetime of 10-11 s and an estimated quantum yield of 10-5. The original observations (Carter et a1 1967, Christophorou et a1 1968) that @FA/@;&%@;:/@:$ have been confirmed by
%/@:&=~FD
~ Q [ M Q ] / ~ F M~ M D .

930

J B Birks

Joneleit (1969) for several liquid alkyl benzenes, but they have been questioned by Horrocks (1970), who irradiated liquid benzene with intense 600 keV electron pulses of 3 ns duration and observed @;=/Diu= . ;@ :/ :@ : Horrocks (1970) recorded the fluorescence spectrum 20 ns after the electron pulse, in order to exclude a large fast emission component, with a lifetime equal to the electron pulse, which he attributed to Cerenkov emission. It is considered that this fast component also includes the predominant ID" fluorescence observed in the total emission spectrum by Carter et al (1967) and Christophorou et al (1968), whose electron excitation energy (100 keV) was below the Cerenkov threshold. The fluorescence observed by Horrocks (1970) originates from the small fraction of excited species which survive the initial intense ionization quenching and then rapidly establish dynamic equilibrium. West and Nichols (1970) have shown that a radiolysis product, possibly biphenyl, also contributes to the emission spectrum of liquid benzene under LI electron or proton irradiation, but this can be eliminated by the use of a flow-cell technique.

5.2. Molecular ion association


The collisional interaction of molecular cations and anions can yield the following products: (XII j (XIII) Hoytink (1968) has discussed the thermodynamics and kinetics of these processes in aromatic hydrocarbon solutions. Chandross et a1 (1965) studied the electrochemiluminescence (ECL)produced by the alternating-current electrolysis of aromatic hydrocarbon solutions in polar solvents, such as acetonitrile and dimethylformamide, and they interpreted the behaviour in terms of processes (X) and (XII). Tetracene, rubrene and 9,lO-diphenylanthracene solutions have structured ECL spectra ascribed to 1M" fluorescence from (X), and perylene and 3 :4-benzopyrene solutions have structureless ECL spectra attributed to IDx fluorescence from (XII). 9,lO-dimethylanthracene solutions have both structured and structureless ECL emission bands attributed to 1M* and ID* fluorescence from (X) and (XII) respectively. This assignment is supported by subsequent studies (Parker and Short 1967) showing that the ratio of the excimer to the molecular ECL yields exceeds the corresponding ratio of the normal photofluorescence yields. As discussed in 95.3, this observation indicates that ID* is formed directly by (XII), rather than by the subsequent association of 1M" and 1M by process (11). Further studies of the ECL emission of 9,lO-dimethylanthracene have been made by Werner et al(1970). Anthracene and phenanthrene solutions also exhibit both structured and structureless ECL emission bands, initially attributed to 1M* and ID* fluorescence from (X) and (XII) respectively (Chandross et a1 1965). The structureless ECL emission of anthracene solutions has since been shown to originate from anthranol, produced by and then excited by energy transfer the decomposition of the anthracene cation (2M+) from anthracene (1M") to emit at lower energies than the 1M" fluorescence (Faulkner and Bard 1968, Werner et al 1970). The structureless band in the ECL spectra of phenanthrene solutions is probably also due to a decomposition product, since phen-

Excimers

931

anthrene solutions do not emit excimer fluorescence or exhibit concentration quenching of molecular fluorescence at room temperature (Birks and Georghiou 1968, Birks 1970a).

5.3. Triplet-triplet interaction


The collisional interaction of two triplet-excited molecules or triplet excitons can yield the following products :

M " fluorescence of type A crystals, such as anthracene, is due to the The delayed 1 homofusion process (XIV) in competition with the quenching process (XV) (Sternlicht et a Z 1963, Avakian and Merrifield 1968). The delayed ID" fluorescence of type B crystals, such as pyrene, is due to the homofusion process (XVI) in competition with (XV) (Avakian and Abramson 1965, Peter and Vaubel 1973b). Triplet-triplet and other exciton interactions in organic molecular crystals have been comprehensively reviewed by Swenberg and Geacintov (1973). Delayed 1M* fluorescence due to triplet-triplet interaction (XIV) was first identified in concentrated rigid solutions of naphthalene (Czarnecki 1961) and 3 :4benzopyrene (Muel 1962). The delayed 1M" fluorescence of anthracene and phenanthrene in dilute fluid solution, resulting from process (XIV), was observed by Parker and Hatchard (1962a,b). The latter authors (1963) also observed the delayed lM* and IDx fluorescences of pyrene in fluid solutions, resulting from processes (XIV) and (XVI) respectively. Parker (1964a, 1967) introduced the term 'P-type (pyrene-type) delayed fluorescence' to describe the emission resulting from triplettriplet interaction, in order to distinguish it from the 'E-type (eosin-type) delayed fluorescence' exhibited by many dye molecules, which is due to thermal repopulation of the excited singlet state S1 from the adjacent excited triplet state TI. P-type delayed 1M" and ID* fluorescences have been observed in fluid solution and/or the liquid phase from naphthalene and its 1- and 2-methyl derivatives, anthracene and its 9-methyl, 9-phenyl and 9,lO-dimethyl derivatives, 1:2-benzanthracene and its 5-methyl derivative, pyrene and 3 :4-benzopyrene (Tanaka et aZ1963, Parker 1963a,b, Stevens and Walker 1964, Birks et al 1966b, 1968c, Moore 1966, Birks and Moore 1967, Moore and Munro 1967, Stevens and Ban 1968, Holzman and Jarnagin 1969, Langelaar et aZ1971b). The initial kinetic analysis of a fluid solution system exhibiting P-type delayed 1M* and ID* fluorescences (Birks 1963, 1964c) was extended to consider the effect of the solvent viscosity and temperature on the quantum yields of the prompt and delayed 1M" and ID* fluorescences and the 3M* phosphorescence (Birks et aZ1968c). The delayed fluorescence decay kinetics have also been considered (Birks 1970a,b). After establishment of an initial photostationary triplet concentration [3M*lS during the excitation period, the triplet concentration [3M*] decreases as follows during the dark period:

13M*]= [3M*]s exp (- k~t)/{l + [3M*]s

(~TT/~T [I )

- exp (-

k~t)])

(5.6)

932

J B Birks

where AT is the total first-order 3M+ decay rate, including process (XV), and ATT is the total rate parameter of all 3M*-3Mn annihilation processes. The 3M+ phosphorescence response function ip(t) is proportional to [3M*], and the 1M+ and IDn delayed fluorescence response functions i$(t) and i$(t) are proportional to [3M+]2. Two limiting cases of (5.6) may be distinguished:

(A) ~ T ~ A T T [ ~ M corresponding *], to small [3M*Is and/or large t ; and (B) K T < ~ T T [3Mn], corresponding to large [3MxIs and/or small t.
I n case (A) (5.6) reduces to

[3M"]=[3M*ls exp (-&)


so that
ip( t )CC exp

(5.6A)
( 5 .7A) ( 5 . SA)

( -k ~ t )

itf(t),i$(t)cc exp ( - 2 A T t ) .

The 3Mn phosphorescence decays exponentially with a lifetime of TT( = l / k ~ ) and , the lMn and ID* delayed fluorescences decay exponentially with a common lifetime of $ 7 ~ .This behaviour has been verified experimentally, and observations of i&(t) and &t) have been used to determine TT from ( 5 . SA) at higher temperatures where the phosphorescence yield is negligible (Stevens and Walker 1963, 1964, Birks et al 1968~). I n case (B) (5.6) reduces to

[3M+] = [3M+]~/( 1+ ATT [3M"]st).


The 3M+ phosphorescence yield
@PT

(5.6B) (5.SB)

is negligible, and

i&(t), i$(t)cc[3M*]2

have a common non-exponential bimolecular decay. Type B behaviour has been observed in intensely irradiated pyrene solutions, and it has been analysed to determine the temperature dependence of the relative triplet quantum yield @T (Birks and Moore 1967). From a kinetic analysis (Birks 1963, 1970a) the ratio of the quantum yields @&, and @iM of the delayed IDn and lMn fluorescences respectively is given by

where

(5 9)
(5.10)

and ADTT and kMTT are the rate parameters of processes (XVI) and (XIV), yielding IDx and lMx 1M respectively. If KDTT = 0, a: = 0, (5.9) reduces to

@ F D / @ F M = k FD kDM ['M]/AFM (AD -k A M D ) = Kl [lM]

(4.12)

which is the relation describing the prompt fluorescence. I n general a is the ratio of the probabilities of the initial formation of ID+ and 1M+ by a particular process, and (5.9) can be applied to any such process, irrespective of the mechanism. For P-type delayed fluorescence it is observed that
@&/@&a > @FD/@FM

(5.11)

Excimers

933

demonstrating that 01 > 0 and that (XVI) and (XIV) are parallel, and not consecutive, processes. Similar observations on the ECL spectra of 5410-dimethylanthracene solutions (Parker and Short 1967), where (5.11) also applies, have shown that (XII) and (X) are parallel, and not consecutive, processes ($5.2). The ID+ emission is more prominent in the delayed fluorescence spectrum than in the prompt fluorescence spectrum ( 5 . l l ) , and the delayed and prompt emissions can therefore be distinguished spectroscopically. Birks and Seifert (1965) have used this feature to show that the ruby-laser-excited fluorescence of pyrene solutions is due to biphotonic excitation of lMQ,yielding prompt fluorescence, and not to monophotonic excitation of 3MQ,yielding P-type delayed fluorescence. T h e parameter 01 and its temperature (T) dependence have been observed for 2 pyrene, 1:2-benzanthracene and other compounds in various solvents (Birks et a 196613, Moore and Munro 1967, Stevens and Ban 1968, Labhart and Wyrsch 1971). For pyrene in ethanol 01 has a limiting value of 1.8 at T > 270 K ; it decreases rapidly with decrease in temperature to 01=0.5 at T= 160 K (Stevens and Ban 1968). It has been proposed that the ID" formation process (XVI) is a short-range diffusioncontrolled collisional process of rate parameter kDrT =pDkdiff, and the 1M+ 1M formation process (XIV) is due to a longer-range electron exchange interaction, which has a temperature-independent component and a temperature-dependent, diffusioncontrolled component, corresponding to a total rate parameter ~ Y T = T kgTTf p M k d i f f . Hence E = p D kdiff/(k&TT +pM kdiff) (5.12)

WherepD andpM are the reaction probabilities per encounter (see $44,and
kdiff = 8RT/30007.

(5.13)

p ~ / p 1.8, ~ =kLTT/pM=8 x 107 M-1 s-1 (Stevens and Ban 1968). Evidence for

Relation (5.12) provides a good fit to the experimental data for pyrene in ethanol for

has been obtained from studies on pyrene in propylene glycol (Birks et al 1968c), and from observations of P-type delayed 1M" fluorescence from rigid solutions (Czarnecki 1961, Muel 1962, Azumi and McGlynn 1963a,b, Parker 1964b). The observation of P-type delayed 1M" fluorescence from 9,1O-diphenylanthracene,a molecule which does not exhibit IDx formation due to steric hindrance ( $3.1), shows the medium-range ( 10-15 A) nature of the interaction responsible for KMTT, process (XIV) (Parker and Joyce 1967). Stevens (1969) has proposed that the primary 3M+-3M+interaction is the mediumrange process 3M++ 3MQ+ 1M: + 1Mo (XW where the subscript 0 identifies the two species originating in the same annihilation event, and that this is followed by
N

kgTT

1M: + lMo-+lDx (114 a re-encounter association of the annihilation partners to yield ID" with probability 8'. This occurs in competition with 1MZ unimolecular decay of rate K M and with the normal ID+ formation process 1M: +1M ---f ID+ (IIb) of rate KDM [IM], where 1M is an unexcited molecule other than 1Mo. On this model
01 = ~

DTT/~MT =T P I / ( 1-

P).

(5.14)

934

J B Birks

For pyrene, where p = 1 for process (IIa), 8' = p, the 1M,*-1Mo re-encounter probability. The latter, which depends on T / r and the 1M,*lifetime, was evaluated using a one-dimensional random-walk model. At low T / r , p=O, a=O, and at high Tly, p=g, a=2, as observed. The model also agrees with experiment at intermediate T/y, and at > 0-3 it approximates to (5.12). The lower a values observed (Birks et al 1966b) for 1:Zbenzanthracene ( a = 0-9) and 5-methyl 1:2-benzanthracene ( a = 0.86) in room-temperature cyclohexane solutions are explained by the lower values of p ( = /3'/8) of 0.4 and 0.6 respectively for these compounds (Birks et al1964b). T h e available data on a a n d p and their dependence on the solvent and temperature are unfortunately much too sparse to provide an adequate test of the re-encounter model. T h e application of a magnetic field H to an aromatic molecular crystal modifies the rate parameter kMTT of 3M"-3M" interaction and the resultant delayed 1M" fluorescence yield O& (Swenberg and Geacintov 1973). Faulkner and Bard (1969) : N of P-type delayed 1M" fluorescence found that a magnetic field reduces the yield O of anthracene in N,N-dimethylformamide solution by about 4.5 % on increasing H from 0 to 8 kG, and they showed that the effect was due solely to the influence of
4 2

-2

500

1000 H(G1

1500

Figure 11. 1:2-benzanthracene in dilute ethanol solution. Relative changes of intensities of phosphorescence (@PT), delayed molecular fluorescence (@&) and delayed excimer fluorescence as a function of magnetic field strength H : 0 (DPT at - 133 " C ; 0 @& at -133 "C; 0 @gx at -107 "C; @.ID at -107 "C (after Wyrsch and Labhart 1971a).

H on kMTT. These results were confirmed by Avakian et al (1971), who observed similar behaviour for anthracene in ethanol solution and proposed a model to explain the monotonic decrease of kMTT with increase in H. Wyrsch and Labhart (1971a) have studied the influence of a magnetic field on the delayed emission of 1:Zbenzanthracene in fluid ethanol solutions from - 70 to - 160 "C. The solutions have three long-lived emissions: phosphorescence (yield OPT),delayed lMx fluorescence (yield OiM), and delayed IDx fluorescence (yield O$,). Typical results are shown in figure 11. With relatively low magnetic fields ( H <1.65 kG), OPTis observed to be independent of H. At - 107 "C @& increases by about 2.2% to a maximum at H=0.4 kG and then decreases steadily to about its zero-field value at H = 1.5 kG, while OiD increases by about 4% to a maximum at and O$, behave in a H=0.4 kG and then remains constant up to H= 1-65kG. is between 0.5 and 2.2% similar manner at other temperatures. The increase in at H = 0.4 kG, and between 0 and - 1.0% at H = 1.5 kG, between 80 and - 150 "C. The increase in O$D is between 1.6 and 4% at H = 0 . 4 kG, and it is constant from H = 0 - 4 to 1-5kG, between -70 and - 110 "C.

Excimers

935

T h e reaction kinetics of a solution system exhibiting delayed 1M* and ID* fluorescence are rather complicated (Birks 1970a,b). and @gD depend on the and ~ D T T ,and on the concentration [IM], on the excitation light intensity, on ~ M T T other 1M*, ID* and 3 M * rate parameters, which are influenced by the solvent and the temperature. For a specific solution system (given solvent, solute, concentration and temperature) we may write (Birks 1970a)
@&a = a[kDkMTT fkMD (kMTT -k kDTT)]
@$D
b[kMkDTT

+kDM ['MI

(5 * 15) (5.16)

(kMTT fADTT)]

where a and b depend on the other rate parameters, the concentration, and the excitation light intensity, which are independent of H . For dilute ethanolic solutions of 1:%benzanthracene below -70 "C (Wyrsch and Labhart 1971b)

k ~ ~,D B ~ [lM], D Mk m
so that (5.15) and (5.16) simplify to
@&=
@$D

(5.17)
(5 15a)
9

akDkMTT

=b k ~ k ~ ~ p

(5.16~)

The magnetic field dependence of @iM and @gD (figure 11) thus corresponds to that of kMTT and ~ T respectively. T The difference in the behaviour of ~ M T T and ~ D T T shows that the two processes are distinct, and do not occur via a common excimer intermediate. Relation (5.12) and the molecular re-encounter model of Stevens (1969) are not necessarily inconsistent with the magnetic field effects, although they do not explain them. A quantum mechanical treatment of processes (X1V)-(XVIII) and the spinspin interactions between them, analogous to that developed for process (XIV) in aromatic molecular crystals (Swenberg and Geacintov 1973), is required. The model will need to take account of the anisotropies of the different processes involved in the molecular encounter. The medium-range electron exchange interaction responsible for process (XIV) is approximately isotropic and extends beyond the van der Waals radius. I n contrast the short-range process (XVI) is highly anisotropic, due to the anisotropy of the exciton resonance interaction ( $2.1) and to the steric limitations on the possible excimer conformations. Just as observations of KDM have led to a deeper understanding of diffusion-controlled processes (Alwattar et al 1973), the interpreand kDTT and their dependence on solute, solvent, tation of observations of ~ M T T temperature and magnetic field could provide a means of elucidating bimolecular photophysical interactions.

6. Photodimers, sandwich dimers and excimers


The cata-condensed aromatic hydrocarbons may be divided into two groups : (i) the majority in which SI = 1 L b , which have only a weak So-Sl transition moment m l ; and (ii) anthracene and the higher polyacenes in which S1= 1L8,which have a relatively strong So-Sl transition moment ml. Group (i) includes benzene and its alkyl derivatives, in which the configurational mixing with the charge resonance state ($2.4) is sufficient to stabilize the 1(E1), state
62

936

J B Birks

originating from SI (figure 3). I n naphthalene and higher compounds in group (i) ID* corresponds to a l(Ep)g state originating from a higher excited molecular state S,, with a larger transition moment m , , as discussed in $2.4. The l(Ep)g energy is a minimum in the symmetrical 1(M IIM)" conformation, but it is only associative when r Q Y D = (mpz/A1,)l/3 (2.9P) where A l p = S ~ - S and ~ , YD is the effective range of the excimer exciton interaction (Birks 1967). <S , O for all For molecules in group (ii), where S1=1L,, the energy of ~(EI), values of r, ie the range of the excimer exciton interaction is unlimited. The resultant attractive potential is a maximum in the symmetrical 1(M 11 M)" conformation ( 2 . 4 ~ ) and is further enhanced by charge resonance configurational mixing. I n photoexcited fluid solutions of anthracene, tetracene and pentacene (Birks et al 1963a) the attractive potential aligns 1M* and 1M in a close symmetrical 1(M llM)* conformation (figure 2(i)). Here they interact chemically to form two covalent bonds linking the meso positions of the two molecules, the regions of high electron density in the lLa state, and thus form the photodimer 1M2 (figure 2(ii)). The four valence electrons in each of the four meso carbon atoms change from an unsaturated (sp2 hybrid, T") electron configuration to a saturated (sp3 hybrid) electron configuration. Ferguson and Mau (1974) have made an elegant study of the photophysics of the anthracene sandwich dimers 1(M /IM), formed by the photolysis of dianthracene molecules 1M2 in a dianthracene crystal by process (VI). Photoexcitation of 1(M 11 M) then leads to the following processes:

The quantum yields of ID* fluorescence, @FD( = ~ F D / ~ D )and , of 1M2 formation, @'AD( =KAD/~D), are observed to sum to unity at all temperatures from 4 to 300 K. At low temperatures @FD= 1.0 and AD=^, and at room temperature OFD tends to zero and @'AD tends to unity. I n the relation

kiD = 7.6 x 109 s-1

and photodimerization are the only significant processes competing for the ID" excitation energy, and they confirm that the excimer is the intermediate state in photodimer formation. Meso substitution in anthracene introduces steric hindrance to photodimer and excimer formation (Birks 1970a). I n fluid solution the 9-substituted anthracenes exhibit both excimer fluorescence and photodimerization. Birks and Aladekomo (1963) proposed that the excimer fluorescence of 9-methylanthracene in solution occurs from the cis (head-to-head) excimer conformation ID,*, in which the steric hindrance by the methyl groups to the close approach of the two molecules is a maximum, and that the photodimer is formed from the trans (head-to-tail) excimer conformation ID:, in which the steric hindrance is a minimum. Other solution and crystal studies support this hypothesis. All the 9-substituted anthracenes studied in room-temperature solution yield photodimers with a trans conformation (Applequist et al 1959, Calas et al 1960, 1965, Greene 1960, Chapman and Lee 1969)

k m =kkD exp ( - WAD/KT) (6.1) and WAD = 590 cm-1. The results show that excimer fluorescence

Excimers

937

(see figure 2). The trans sandwich dimer of 9-cyanoanthracene (CNA), formed by the photolysis of the trans photodimer in rigid solution, exhibits only weak trans excimer (ID:) fluorescence, and reverts rapidly to the trans photodimer on exposure to light (Chandross and Ferguson 1966b). This contrasts with the strong cis excimer (ID:) fluorescence of CNA in concentrated solutions or crystals. In topochemically controlled reactions in crystals the lattice structure usually determines the geometric conformation of the photodimer (Schmidt 1967). CNA and other 9-substituted anthracenes are striking exceptions to the general pattern. They crystallize in a cis 1(M 11 M) conformation and exhibit strong cis excimer(lD,*) fluorescence, but only trans photodimers are formed within the crystal (Craig and SartiFantoni 1966, Cohen 1969, Cohen et aZ1969, 1971, Kawada and Labes 1970). Craig and Sarti-Fantoni (1966) proposed that in CNA crystals photodimerization only occurs at defects or surfaces or in zones disordered by photodimerization. Cohen (1969) and Cohen et a Z (1969, 1971) showed that the CNA crystal photodimerization occurs preferentially at dislocations emerging at the bc and ac crystal faces. Kawada and Labes (1970) studied the photodimerization of single CNA crystals by observing the growth of four-sided photoetched pits as a function of time, light intensity and temperature. They proposed that the cis excimer ID,* is formed initially, leading to exciton migration which culminates in ID,* fluorescence or internal quenching or in a ratedetermining, thermally activated, reorientation to the trans excimer ID;, at the perimeter of the etch pit, followed by rapid formation of the trans photodimer 1M2: '(M II M)c + ~ V D By analogy with the behaviour of pyrene crystals (58), it is probable that excimer (ID:) exciton migration occurs in CNA crystals. The 1 D , * 4 D Fprocess involves intermolecular (or inter-excimer) energy transfer at the perimeter of the etch pit, rather than 180" internal rotation of a given dimer pair. Aladekomo (1973) observed the photodimerization quantum yield @)AD and excimer fluorescence quantum yield @FD of 9-methylanthracene in toluene solution as a function of concentration and temperature. I n fluid solution, where crystal restraints are absent, (XX) simplifies to (XIX). The 9-methylanthracene solutipn data were analysed in terms of a reaction scheme similar to (XIX), yielding
a relation which was verified experimentally. The value of WAD=1850 cm-1 for = 590 cm-1 for anthracene (Ferguson and 9-methylanthracene exceeds that of WAD Mau 1974) because of the steric effects of the methyl groups. Ferguson et aZ(l973) have studied the absorption spectra and fluorescence spectra and lifetimes of the symmetrical sandwich dimers of anthracene and its 9-methyl, 9-phenyl, 9-chloro, 9-bromo and 9-cyano derivatives. The sandwich dimers were formed by photolysis in a rigid glass matrix at 77 K. The glass was then allowed to soften to study the effects of dissociation of the sandwich dimer. Initially the fluorescence spectrum shifts slightly to higher energies, and the lifetime decreases from ~ 2 0 to 0 100 ns. There is then an abrupt spectral broadening and the appearance of vibrational structure, and the lifetime drops to the molecular value of - 5 ns. Anthracene and its 9-methyl, 9,lO-dimethyl and 9,lO-diphenyl derivatives each have SI= 1L,, with similar excimer attractive potentials mlZ/rS in the symmetric
N

938

J B Birh

(M 11 M) conformation. T h e difference in their ability to form photodimers or excimers in fluid solutions is due to the steric factors which determine the minimum intermolecular spacing rmin. rmin is least for the sandwich dimers of anthracene and of 9-methylanthracene in the trans conformation, and these systems photodimerize. rmin is larger for the sandwich dimers of 9-methylanthracene in the cis conformation and of 9,1O-dimethylanthracene, and these systems form excimers, but not photodimers. rmin is largest for the sandwich dimer of 9,lO-diphenylanthracene (the phenyl groups are at 60" to the anthracene plane), and this system does not form photodimers or excimers. It does, however, give P-type delayed fluorescence by process (XIV), showing that the range of the latter interaction exceeds that of the excimer interaction (Parker and Joyce 1967).
7. Photophysical processes in excimers 7.1. Radiationlessprocesses
There are four possible modes of ID" decay: fluorescence (rate KFD), inter-system crossing (rate KND), internal conversion (rate KGD), and dissociation into lMx and 1M (rate AND) (figure 9). As described in $2, KFD, KMD and KID ( = K N D + ~ G D ) can be determined for a given solution system from observations of the fluorescence response functions and quantum yields. For molecules such as pyrene (Birks et al 1964d), and 9-methyl and 9,lO-dimethyl anthracene (Barnes and Birks 1966),

KFD

=f i 2 ( K p ~ ) o

(4.39)

depends on the solvent refractive index n, but is otherwise independent of the temperature T. RMD is strongly temperature-dependent, (4.43) where WMD is the sum of the IDx binding energy B and the diffusion activation energy W D ~(4.46). I I n general (4.41) is the sum of a temperature-independent component kyD and a temperature-dependent component ATD. It is of interest to relate these to the fundamental processes of intersystem crossing ( K N D ) and internal conversion (KGD). Table 1 lists experimental data on K F D , ( K F D ) ~ , KITD (at 20 "C), A;,, and WID for pyrene in various solvents (Birks et al 1963b, 1964d, 1971, Parker and Hatchard 1963, Smith-Saville 1968, Hauser and Heidt 1970), and for pyrene crystals (Birks et a1 1966a). Also listed are the values of K N D and KGD for pyrene in ethanol and nhexane solution at 20 "C, obtained from observations of the concentration dependence of the total triplet yield (DT (Medinger and Wilkinson 1966, Heinzelmann and Labhart 1969). Several conclusions may be drawn from the data (Birks et a1 1971). ( a ) The values of K F D are consistent with (4.39) with a mean value of ( K F D ) ~ = 5.5 x 106 s-1. (b) For the pyrene crystal excimer K F D = ( K F D ) ~ , corresponding to an apparent refractive index n = l . This result is probably fortuitous. The crystal excimer is limited to 1(M j j M)" conformations, while the solution system includes non-sandwich excimer conformations, which will tend to reduce its mean radiative lifetime.

Exciws
lo7s-1; W values in eV.

939

Table 1. Pyrene excimer parameters in various solvents at 20 "C (Birks et aZ1971). k values in
Solvent Propylene glycol Isopropanol n-nonane Acetone Ethanol Cyclohexane n-hexane Crystal

k m
0.95
1-1 1.0
1.0

(kFD)O

k ! D

' k
0 0.39

k m

0.49

0.55

1 e07 1e 1 6 1 -05

0.50 0.56 -0 0.55 0.3

2.4 0.48

0.58 0.57
0.55

-0.6 1.0 0.23 0.65

-0

0.09 0.15

0.30 0.70 0.33

0.10 0.10 0.25 0.63 0.1 0.14 0.1 0.13 0.15 0.70 4.7 0.07 -

~ G D

1500 2700 220 50 40 20

WID 0.23 0.22 0.15 0.1

0.22 0.23

W,,

(c) For the two solvents for which data are available, the temperature-independent =AND), and the temperatureradiationless transition is inter-system crossing dependent radiationless transition is internal conversion (K:, = KGD). (d) The thermally activated internal conversion is attributed to molecular motion which yields non-sandwich excimer conformations and leads to increased intermolecular vibrational interaction. This motion is subject to viscous constraint, and there is satisfactory agreement between the values of WIDand W,,, the activation energy for viscous motion (table 1). When 7 2 4.7 CPthe molecular motion and the thermally activated internal conversion are inhibited (Birks et a l 1971). (e) I n the pyrene crystal excimer the thermally activated radiationless transition AyD is attributed to inter-system crossing to 3D", which dissociates into 3M*+1M. The observation of pyrene crystal 3M" phosphorescence (Peter and Vaubel1973b) and delayed fluorescence due to 3M1-3M* interaction (Avakian and Abramson 1965) supports this conclusion. Internal conversion via a non-sandwich excimer conformation is inhibited in the crystal lattice. On the experimental potential diagram of the pyrene crystal excimer (figure 8) the 3M*+lM potential curve crosses the ID* curve at 0.1 eV ( WID) above its minimum, consistent with thermally activated inter-system crossing. Cundall and Pereira (1972a) determined the excimer fluorescence quantum efficiency QFD and lifetime TD( = 1 / k ~ for ) 1-methylnaphthalene in ethanol from - 30 : , = 5 x 106 to 60 "C. They found that KFD is independent of temperature, and that k s-1, K;,=3.2 x 1011 s-1 and W 1 ~ = 0 * 2eV. 9 From the magnitudes of and KID they concluded that the former corresponds to inter-system crossing and the latter to internal conversion, as in the pyrene solution excimer.
N N

7.2. Radiative processes


I t is observed that KFD/KFM is independent of the solvent and temperature T for solutions of pyrene (Birks et aZ1964d) and 9-methyl and 9,lO-dimethyl anthracene (Barnes and Birks 1966), and independent of T (at TG300 K) for solutions of benzene (Gregory and Helman 1972) and toluene (Greenleaf et al 1968). For solutions of 1-methylnaphthalene KFD is practically independent of T (Cundall and Pereira 1972a). The experimental values of KFM and ~ F for D fluid solutions and for rigid solutions of molecules and their sandwich dimers are listed in table 2. For molecules with SI='Lb, ~ F M 106 s-1, and for those with SI=1L,, ~ F M108 s-1. For most excimers and sandwich dimers K F D N 106 s-1, but for the excimers of pyrene and of 1:2benzanthracene and its derivatives KFD- 107 s-1, which exceeds k ~ n ffor the parent molecule.
N N

940
Compound Benzene

J B Birks
Observers Solvent k m k m Methylcyclohexane 2 -52 0.65 Gregory and Helman (1972) (298 K) Methylcyclohexane 1e66 0 - 4 2 Gregory and Helman (1972) (195 K) Cyclohexane (293 K) 2 * 14 0.76 Cundall and Robinson (1972a) Cyclohexane (313 K) 2-28 1.12 Cundall and Robinson (1972a) 2.4 1.6 Greenleaf et aZ(1968) n-hexane (293 K) 95% ethanol 2.3 0.9 Selinger (1966) 1.1 Selinger (1966) 95% ethanol 2.9 95% ethanol 3.4 1.1 Selinger (1966) Methylcyclohexane 50 1-35 Ferguson et a Z (1973) glass (77 K) Methylcyclohexane 78 1.5 Ferguson et aZ(1973) glass (77 K) Methylcyclohexane 83 1.5 Ferguson et aZ(l973) glass (77 K) Methylcyclohexane 100 0.94 Ferguson et aZ(1973) glass (77 K) Methylcyclohexane 100 0.47 Ferguson et aZ(1973) glass (77 K) Methylcyclohexane 91 0.70 Ferguson et aZ(1973) glass (77 K) 1.5 11.6 Birks et aZ(1963b) Cyclohexane 4.2 Cyclohexane 11 Birks et aZ(1964b) 4.7 Cyclohexane 5.3 Birks et aZ(1964b) 2.5 Cyclohexane 8.5 Birks et aZ(1964b) 4.6 Cyclohexane 6.6 Birks et aZ(1964b)

Table 2. Radiative molecular ( k m ) and excimer ( k m ) rate parameters (units of lo6 s-l)

Toluene Naphthalene 1-methylnaphthalene 2-methylnaphthalane Anthracene sandwich dimer


(SDI

9-methylanthracene (SD) 9-phenylanthracene (SD) 9-chloroanthracene (SD) 9-bromoanthracene (SD) 9-cyanoanthracene (SD) Pyrene 1:2-benzanthracene 5-methyl 1:2-benzanthracene 6-methyl 1:2-benzanthracene 10-methyl 1:2-benzanthracene

The radiative transition from the symmetrical 1(M If M)" excimer conformation to the ground state is symmetry-forbidden. Several authors (Konijnenberg 1963, Azumi and Azumi 1966a, b, Chandra and Lim 1968, 1969) have therefore attempted to explain the finite values of kFD in terms of rotational distortion of the 1(M IIM)" conformation. This hypothesis is both incorrect and redundant, since the pyrene crystal excimer (table 1) and the symmetrical sandwich dimers of anthracene and its derivatives (table 2), which have 1(M /I M)" conformations, possess kFD values similar to those of solution excimers. The Sl-So fluorescence transition in molecular benzene is symmetry-forbidden, and the finite value of FZFM is due to vibronic coupling to a higher excited molecular state S 4 = 1E1,) (Murre11 and Pople 1956). By analogy it is proposed (Birks 1970a) that the finite value of KFD for the fluorescence transition from the ID" state is derived from vibronic coupling to one or more higher excited excimer (ID"") states., Further evidence for this is considered later. The relation K F M = ~ ' (kFM)o (4.38) is an abbreviated form of the theoretical expression (Strickler and Berg 1962, Birks and Dyson 1963), derived from the relation between the Einstein A and B coefficients in a medium of refractive index n,

Excimers

941

where e( 3) is the molar extinction coefficient at wavenumber 3, (Ij-3>F is the mean value of (3)-3 over the fluorescence spectrum, and the integral extends over the first absorption band system. This relation has been verified experimentally for several condensed aromatic hydrocarbons and other molecules (see Birks 1970a). Cundall and Pereira (197213) determined K F M for several aromatic hydrocarbons in solution as a function of temperature and solvent. For aromatic hydrocarbons, other than benzene, the observed temperature dependence of KFM is entirely due to that of n2. Substitution in (4.38) yields temperature-independent values of (KFM)O, which are practically independent of the solvent. For benzene the temperature dependence of n2 is insufficient to account for that of K F M , and substitution in (4.38) yields values of ( K F M ) O which depend on the temperature and solvent. I t is therefore necessary to distinguish the radiative properties of benzene and its derivatives from those of the polycyclic aromatic hydrocarbons. The validity of (4.38) and (7.1) for the polycyclic aromatic hydrocarbons (Birks 1970a, Cundall and Pereira 1972b) and the observations that KFD/KFM is independent of the solvent and temperature represent evidence for the relation (Birks et aZl965) (4.39) which has been verified for the pyrene solution excimer (table 1). It is proposed that KFD is given by an expression analogous to (7.1). This hypothesis could be tested by observations on sandwich dimers in rigid solutions, where e( 3) can be determined (Ferguson et al 1973). e(;) cannot be observed for excimers in fluid solutions, since they are dissociated in the ground state. Relation (4.39) indicates that the solution excimers of polycyclic aromatic hydrocarbons, such as naphthalene, anthracene and pyrene, have a definite equilibrium steric conformation, and this is equated to the symmetrical 1(M // M)" sandwich conformation. The temperature dependence of KFD for the benzene and toluene excimers differs from that of the excimers of the polycyclic aromatic hydrocarbons. Hirayama and Lipsky (1969a, b) analysed data on the concentration and temperature dependence of @FM and @)FD of benzene and toluene solutions using relation (4.13), and they concluded that qFD and KFJJ increase markedly with increase in temperature. Gregory and Helman (1972) observed a 55% increase in K)FD for benzene in methylcyclohexane solution over the 103 "C interval from 195 to 298 K, although the ratio KFD/KFM increased by only 2%. Greenleaf et a Z (1968) concluded that K F D / K F M is practically independent of temperature for toluene in n-hexane over the 135 "C interval from 188 to 325 K. This conclusion is not invalidated by their assumption (probably incorrect) that KFM for toluene obeys relation (4.38). Cundall and Robinson (1972a, b) observed qFD = 0.0227, T D = 30 ns at 293 K, and qFD = 0,0276, TO = 24.7 ns at 313 K, for benzene in cyclohexane solution, corresponding to a dramatic 47% increase in KFD and a 38% increase in K F D / K F M over the 20 "C interval (table 2). The experimental values of K F M for benzene in different solvents as a function of temperature are not consistent with relation (4.38), and they are a factor of 2-3 less than the theoretical values of K F M obtained from (7.1) (Cundall and Pereira 197213). The derivation of (7.1) by Strickler and Berg (1962) involves the expansion of the electronic dipole transition moment integral Mlu(Q)as a power series with the normal molecular coordinates Q :

942

J B Birks

Relation (7.1) is derived on the assumption that the zeroth-order term in the expansion (7.2) is dominant, and that the higher-order terms can be neglected. StricMer and Berg (1962) suggested that for forbidden or weakly allowed transitions the higherorder terms in (7.2) may not be negligible. Cundall and Pereira (1972b) have proposed that these higher-order terms are responsible for the strong temperature dependence of K F M for benzene. This hypothesis is probably invalid. T h e value of KFM( = ( K F M ) ~ )for equilibrated benzene vapour (n= 1) at pressures above 10 Torr agrees satisfactorily with (7.1). This indicates that the anomalous behaviour of K F M for benzene in solution is due to a solvent perturbation, and not to the intrinsic properties of the benzene molecule. I n benzene solutions the nature of the solvent influences the intensity of the symmetry-forbidden So-Sl, 0-0 transition (Ham band) and the quantum yield of S3-S1 internal conversion (Lawson et a2 1969), showing that the vibronic coupling between the S3(1ElU)and S1(1BZu) states, responsible for K F M , the Ham band, and the S3-S1 internal conversion, depends on the solvent (Koyanagi 1968, Birks 1970a). At room temperature K F M for benzene in solution approximates to K F M ( = ( J Z F M ) ~ ) for equilibrated benzene vapour, and Cundall and Ogilvie (1975) have suggested that (7.1) with n = 1 may be valid in solution. The agreement of K F M and (KFM)O at room temperature is fortuitous: at lower temperatures an unrealistic value of n < 1 would be required to fit (7.1) to the experimental JZFM values. T h e observed temperature independence of IZFD/KFM for solutions of benzene (Gregory and Helman 1972) and toluene (Greenleaf et aZ1968) over wide temperature ranges shows that ~ F behaves D in a similar manner to K F M . Just as K F M is derived from vibronic coupling between the lBzu and lElu states of the benzene molecule (Murre11 and Pople 1956), it is proposed that KFD is derived from vibronic coupling between the B1 and E1 states of the benzene singlet excimer, which originate from the molecular 1Bzu and lElu states (figure 3). Solvent perturbation of this vibronic coupling accounts for the unusual solvent and temperature dependence of K F M and K F D . Some change in K F D / ~ F M might be expected with increase in temperature, due to the increased thermal excitation of ID", but it is difficult to reconcile the 0.02 deg-1 increase in K F D / K F M for benzene solutions between 293 and 313 K reported by Cundall and Robinson (1972a,b) with the 0.0002 deg-1 increase between 195 and 298 K observed by Gregory and Helman (1972). The latter result appears to be the more reliable. A broad structureless absorption band at 515 nm has been observed in the nanosecond pulse radiolysis of liquid benzene and toluene (Cooper and Thomas 1968). The band, which decays with the ID" lifetime TD, has been assigned (Birks 1968a) to the allowed E l u c B 1 benzene excimer absorption transition (figure 3), which has the appropriate energy at im = 3.3 A, the equilibrium ID" separation. The temperature dependence of the transient absorption intensity is similar to that of the excimer fluorescence intensity, consistent with this assignment (Cooper and Thomas 1968). The transient absorption of the benzene excimer has also been observed by nanosecond flash photolysis, using the frequency-doubled (28 810 cm-1) photons from a 15 ns pulsed, Q-switched ruby laser for the two-photon excitation of liquid benzene at 57 620 cm-1 (173.5 nm) (Richards and Thomas 1970). The transient absorption spectrum of the pyrene excimer in solution has been studied between 400 and 600 nm using laser flash photolysis (Goldschmidt and Ottolenghi 1970) and modulation excitation spectrometry (Slifkin and Al-Chalabi 1973). A single structureless peak is observed with a maximum at 20 400 cm-1 and a band-

Exciws

943

width of 2800 cm-1. Goldschmidt and Ottolenghi (1970) compared their observations with the configurational mixing calculations of Azumi et aZ(1964), and assigned the transient absorption to the allowed 1B&,elB, transition in the pyrene excimer (see 52.4).

8. Excimer excitons in crystals


Singlet exciton (1M") migration in anthracene and other type A molecular crystals occurs by a hopping mechanism at normal temperatures (Birks 1970a). I n pyrene and other type B molecular crystals, which crystallize in 1(M llM) pairs, ID" lies ~ 0 . eV 4 below 1M". The migration of 1M* is therefore inhibited, and it might be expected that the singlet excitation energy would remain trapped in local ID* states. The fluorescence lifetime TD of pyrene crystals decreases with crystal size, and this is associated with the appearance of monomer defect bands in the fluorescence specZ (1966a) proposed that ID" exciton trum. T o account for this effect, Birks et a migration can occur to surface monomer defects. T h e IDx exciton migration has been confirmed and studied by observations of singlet-singlet energy transfer from pyrene host crystals to various guest molecules (Northrop and Simpson 1956, Hochstrasser 1962, Kawaoka and Kearns 1964, 1966, Tomkiewicz and Loewenthal 1969, Klopffer and Bauser 1970, Chu et a Z 1971, Tomura and Takahashi 1971, Klopffer et aZ1972). For crystal pyrene the 1M* and ID" energies are M0=26 000 cm-1 and 0;=23 370 cm-1 respectively at absolute zero (Birks and Kazzaz 1968). Klopffer et aZ(1972) observed that energy transfer does not occur from crystal pyrene to any guest molecule whose singlet excitation energy lY* lies between 25 400 and 26 700 cm-1 (anthracene, picene, l:Zbenzanthracene, 9,10-diphenylanthracene, tetraphenylethylene and 1:2: 5 :6-dibenzanthracene). Energy transfer does occur for guest molecules in which lY+ < 23 350 cm-1 (coronene, perylene, anthanthrene, ovalene, tetracene, dinaphthopyrene, violanthrene and isoviolanthrene). The results show that energy transfer only occurs if ID" > lY", and does not occur if 1M*3 1Y* > ID*. This shows that 1M" migration and transfer processes are absent in pyrene crystals, and that the host-guest energy transfer involves ID* migration and transfer. Klopffer et aZ(1972) used thin crystal specimens and chose guest molecules with weak absorption in the pyrene fluorescence region in order to minimize the probabilities of long-range radiative and radiationless (dipole-dipole) transfer from ID* to lY*. With coronene as guest they observed energy transfer at 298 K, but not at 77 K, showing that the ID* exciton migration requires thermal activation. Two nonfluorescent guests, anthraquinone and acenaphthylene, were used as quenchers in further studies. The hopping time Th( = 5 x 10-11 s at 298 K) and activation energy WDD( = 0.055 eV) of ID* exciton migration were determined from the temperature dependence of the fluorescence quantum yields of pure and doped pyrene crystals. The results show that ID* exciton migration occurs by a thermally activated hopping process, and that in the particular mixed crystals studied it terminates in collisional transfer to lY*. Goldschmidt et a Z (1968) observed a decrease in the ID* fluorescence quantum yield of pyrene crystals at high laser excitation intensities, which they attributed to lDX-lD* exciton interaction. Masuhara and Mataga (1970) observed a reduction in the fluorescence lifetime of the pyrene solution excimer at high excitation intensities,

944

J B Birks

similarly consistent with 1D*-1D* interaction. Inoue et a1 (1972b) have studied the temperature dependence of the rate parameter kDD of 1D*-1D* exciton annihilation in pyrene and perylene crystals. Using a 4 ns pulsed nitrogen laser for excitation they observed the fluorescence response function at different excitation intensities. The mole fraction CD of ID* excitons decays as follows,

dCD/dt = - k D C D - kDDCD2
integration of which gives

(8.1)

is the initial ID* mole fraction. As predicted by (8.2), a linear relation where (CD)O is observed between l/m and exp (kDt). K D D / ~ D is evaluated from its intercept, and kD is determined from the fluorescence decay at low excitation intensity. For pyrene and perylene crystals at 293 K, values of kDD = 9 x 10-15 and 8 x 10-14 cm3 s-1 are obtained respectively. From the temperature dependence of kDD, values of WDD= 0.041 and 0.036 eV are obtained for the activation energies of ID* exciton migration in pyrene and perylene crystals respectively. Klopffer et a1 (1972) considered three possible mechanisms of ID* exciton migration in pyrene crystals, only one of which is consistent with the experimental potential diagram of the pyrene crystal excimer (figure 8). This mechanism involves thermal activation of the excimer to its ground-state equilibrium intermolecular separation (TO = 3.53 A), and energy migration therefrom to a neighbouring ground-state dimer. The 410 nm absorption band, observed by Fischer et a1 (1973) in crystal pyrene at low temperatures and attributed to the Do-DI absorption of the unexcited crystal dimer, is consistent with the pyrene crystal potential diagram (figure 8) and with this model of ID" exciton migration.

9. Triplet aromatic excimers


9.1. Theoreticalconsiderations
T h e transition from the singlet ground state So to an excited triplet state T, of a molecule is spin-forbidden, ie the electric dipole transition moment m,-O. I n a molecuIe containing heavy atoms of high atomic number 2, spin-orbit coupling introduces a component of singlet character into the wavefunction of T, and a comso that m, acquires a reasonable ponent of triplet character into the wavefunction of SO, magnitude. I n an aromatic hydrocarbon where 2,<6 the spin-orbit coupling is small, though finite, so that mq2- 10-*mp2, where m p is the transition moment of an allowed So-S, transition (Birks 1970a). Substitution of a high-2 atom (eg C1, Br, I) in an aromatic hydrocarbon increases mq2 and the probabilities of So-T, and TI - SO transitions, a phenomenon which is known as the internal heavy-atom effect (McClure 1949). A similar enhancement of mq2, known as the external heavy-atom effect, is obtained by using a liquid or glass containing high-2 atoms (eg carbon tetrabromide, ethyl iodide) as a solvent for the unsubstituted aromatic hydrocarbon (Kasha 1952). The triplet exciton resonance (ER) states originating from T, in the (M IIM) conformation are given by relations similar to ( 2 . 4 ~and ) (2.4b) :

Exciws
The corresponding triplet charge resonance (CR)states are, from (2.6),

945

l,l=I-A - C ( r ) . (9.2) The triplet excimer states originate from configurational mixing ($2.4) of triplet ER and CR states of common symmetry. I n the benzene molecule the TI, Tz and T3 states are 3Lp, 3B and 3La respectively (Birks et al 1968a), with 3Blu, 3Elu and 3Bzu symmetries respectively. Figure 12 plots configurational mixing calculations of the transitional energies of the 3(E - R) states of the benzene D,jh excimer as a function of interplanar separation (Hillier et aZ1966). The lowest triplet excimer state 3D+ at r > 2.6 A is the Bzg ER state, originating from TI( = 3L, = 3Bzu), and strongly stabilized by configurational mixing with the Bzg CR state.
3(R)g,

Ring sparution(; 1

Figure 12. Triplet states of benzene excimer. Theoretical energies of D 6 h sandwich dimer as a function of interplanar separation r . Low-energy excimer states of 3Blu, 3 E ~ and u
3 ! bu

molecular parentage (after Hillier et aZ1966).

I n general, in the aromatic hydrocarbons TI= 3L, (see Birks 1970a). In molecules belonging to the D2h symmetry point group (naphthalene, anthracene, pyrene, perylene, etc) the symmetry of the 3L, molecular state is 3Bzu, and its ER states in the (M IjM) D2h conformation are 3B3, and 3Bzu. The lowest-energy triplet CR states also have 3Bsg and 3Bzu symmetries, and configurational mixing of these with the ER states yields the 3B&, 3B;, 3B2; and 3B& excimer states in order of increasing energy. Thus the 3D+ state of these molecules corresponds to the 3B& excimer state. Whether the 3D+ state is stable or not depends on the relative magnitudes of the 3D+ excimer interaction potential V&(Y) and the intermolecular repulsive potential R(Y).By analogy with 53.1, the total 3D+ energy is

Dk(r)= V&) +R(r).

(9.3)

946

J B Bi7ks
= BN(~ ) T:-Dk(7)

The system forms a bound 3Dx state, provided that the parameter has a positive maximum of

BN=T?--D(9.6) corresponding to the 3D* binding energy BN at the 3De equilibrium separation of 7=7". T: is the 3M+ energy, D " is the energy of the 3D+ phosphorescence maximum, and R " , RLN, VANand D& are the potentials at 7 = r " . In the aromatic hydrocarbons the 3D+ interaction is much less than the ID+ interaction. The 3Dm ER interaction is negligible compared with the ID+ ER interaction, because mq2<m,2. Configurational mixing with the CR states increases the stabilization of the 3D+ ER states (figure 12), but the effect is less than for the ID+ ER states (figure 3) because of the increased energy gap A ( =3(R)g-3(E1)g) between the interacting states. The configurational mixing is proportional to A-2. It is therefore concluded (Birks 1975a) that for the aromatic hydrocarbons BN<B TmN > rm (9.7)
= T,O-D"-R&

and the spectroscopic shift

AENT( T,O-D") < AEDM( = S,O- Dm). (9.9) These criteria may be used in the assessment of the spectroscopic evidence for triplet excimers.

9.2. Benzene
For the benzene singlet excimer Dm=31 300 cm-1, A E ~ ~ = 5 7 0 cm-1 0 (Birks et a Z 1965) and B=2800 cm-1 (Cundall and Robinson 1972a), corresponding to Rk=2900 cm-1. Comparison of D m with the calculated energy of the B1, singlet excimer state (figure 3) gives r,=3.3 A. The calculated Bzs triplet excimer energy ? ; . Extra(figure 12) gives AENT(Ym)= 1900 cm-1, which is 1000 cm-1 less than I polation of R'(7) indicates that it may exceed AENT(~) at all values of 7 , in which case benzene does not form stable triplet excimers. If it does so, it is estimated that r " > 3.5 A, AENT < 1000 cm-l and BN< 200 cm-1 (Birks 1975a). The 2,2'-paracyclophane molecule consists of two benzene rings, joined by (CH& chains at their para positions and forming a slightly distorted (M /IM) conformation, with the ring separation distance Y varying from 2.8 to 3.09 A (Klopffer 1973). The fluorescence spectrum and the corresponding absorption spectrum are broad bands, with vibrational fine structure, due to the intramolecular singlet 'dimer'. The phosphorescence spectrum is a similar broad band with a maximum at -21 000 cm-1, which agrees satisfactorily with the calculated benzene B2 e: triplet excimer energy at 7 = 3.0 A (Hillier et aZl966, figure 12). The high value of AENT 8500 cm-1 for the triplet 'dimer' is due to the short 'intermolecular' links. I n 4,4'-paracyclophane the two benzene rings are joined by (CH2)4 links and are separated by 7=3.73 A (Klopffer 1973). The absorption spectrum is structured and shows negligible ground-state 'dimer' interaction. The fluorescence spectrum includes an intense structureless band (Dm=29 400 cm-1) due to intramolecular singlet

Excimers

947

excimers produced by the closer approach of the benzene rings towards r m z 3.3 di. In contrast, the phosphorescence spectrum is structureless and has a bathochromic shift of only 500 cm-1 relative to the 3M+ phosphorescence of ethyl benzene (Hillier et a1 1966). Despite the favourable (M I/M ) conformation, the negligible ground-state interaction, and the formation of singlet excimers, no intramolecular 3D+ phosphorescence is observed. A structureless emission band at 22 000 cm-1 observed in the spectrum of solid benzene excited by 1 MeV electrons has been attributed to benzene 3D* phosphorescence (Phillips and Schug 1969). This assignment is probably incorrect, since it would correspond to AENT 6000 cm-1 AEDM,which is inconsistent with (9.9). The assignments of the structureless emission bands at 19 000-20 000 cm-1, observed in the spectra of six liquid alkyl benzenes excited by an intense electron beam, to 3D* phosphorescence (Christophorou et a1 1968) are also considered to be incorrect. 8000-9000 cm-1, which exceeds the These assignments would correspond to AENTN n values of A&~=35Oc)-4800 cm-1 observed for the alkyl benzenes (Birks 1970a), i contradiction to (9.9). The emissions are probably due to radiation products.
N N

9.3. Naphthalene
The syn and anti 2,2-paracyclonaphthanes are the analogues of 2,2-paracyclophane in which two naphthalene molecules are joined by dimethylene groups at their a positions in symmetric and displaced sandwich conformations, respectively, with a separation of r z 3 A (Froines and Hagerman 1969). The syn and anti isomers , = 21 650 cm-l, AEDM=9250 cm-1, have structureless fluorescence spectra with D , =24 900 cm-l, AEDM = 6000 cm-I respectively. The fluorescence spectrum and D of the anti isomer agrees with that of the 1,4-dimethylnaphthalene excimer, but this is fortuitous since the latter has a symmetric (M IjM) conformation and a wider separation T m N 3.3 A. The phosphorescence spectra of the syn and anti isomers are structured with bathochromic shifts of 2700 and 1700 cm-1 respectively relative to the 3M+ phosphorescence of 1,4-dimethylnaphthalene. Comparison of the structured phosphorescence spectra with the structureless dimer phosphorescence spectrum of 2,2-paracyclophane (Hillier et aZl966) indicates that naphthalene has less tendency to form triplet excimers than benzene. This conclusion is confirmed by the experiments of Chandross and Dempster (1970a), who prepared the intramolecular photodimer of 1,3-di-a-naphthylpropaneY and then photolysed it in solid solution at 77 K to produce the intramolecular sandwich dimer. This consists of two naphthalene groups in a symmetric (M IlM) conformation, and linked by a propane chain at their a positions. The sandwich dimer has a fluorescence spectrum similar to that of the naphthalene excimer, and the absence of any molecular fluorescence shows that all the naphthalene groups are arranged in sandwich pairs. In contrast, the phosphorescence spectrum is structured with only a small bathochromic shift of 250 cm-1 relative to the 3M+ phosphorescence spectrum of the same specimen after destruction of the sandwich structure by thawing and subsequent recooling. It is concluded that naphthalene does not form triplet excimers unless these have a quite different conformation from the l(M IIM) singlet excimer, which is most unlikely. A structureless emission band at 18 600 cm-1 observed in the spectrum of naphthalene in ethanol soIution at 160 K has been attributed to naphthalene 3D+ E 1968, Langelaar 1969). This assignment, which phosphorescence (Langelaar et a
N

948

J B Birh

corresponds to AENT N 2600 cm-1, compared with AEDM = 6500 cm-1 (Aladekomo and Birks 1965), conflicts with the results of Chandross and Dempster (1970a), described above. A similar emission band observed in the spectra of phenanthrene solutions and originally attributed to phenanthrene 3D* phosphorescence (Langelaar et al 1968) was subsequently found to be due to an impurity (Langelaar 1969), and the naphthalene emission may be due to a similar cause.

9.4. Pyrene
T h e pyrene crystal structure with its (M /IM) conformation and negligible groundstate interaction provides ideal conditions for the observation of any triplet excimer phosphorescence. Various long-lived emissions observed from pyrene crystals have been attributed to triplet excimers (Gijzeman et al 1970, 1971, Gijzeman 1972). These emissions have, however, since been shown to be due to impurities (Peter and Vaubel 1973a,b). A pure pyrene crystal emits only excimer (ID*) fluorescence, delayed ID* fluorescence due to 3Mx-3M* interaction, and molecular (3M*) phosphorescence, showing that the triplet excimer is not associated at r=3.53 A. The cumulative spectroscopic evidence indicates that benzene, naphthalene, pyrene and other aromatic hydrocarbons do not emit excimer phosphorescence.

9.5. Triplet concentration quenching


T h e molecular triplet (3M") lifetimes 7 T of o-xylene (Cundall and Voss 1969), naphthalene, anthracene, phenanthrene and pyrene (Langelaar 1969, Langelaar et al 1971a) in fluid solution are observed to decrease with increase in molar concentration [IM] in the following manner: where (RT)o = APT + RGT is the 3M* decay rate at infinite dilution, and kQT is the rate parameter of concentration quenching of 3M*. The triplet concentration quenching has been attributed (Langelaar 1969, Cundall and Voss 1969, Langelaar et al 1971a) to the formation (rate RNT [IM]), dissociation (rate RTN), and decay (rate RN = k p + ~ R G N ) of a triplet excimer (3D") as follows: 3M* + 1M c,$ 3Dx
kNTrl&Il
kTN

(9.10)

1M+ 1M
(9.11)

1kN

On this model

~ Q= TRN~NT/@TN

+RN) =@NT

where RNT is the diffusion-controlled rate parameter, and p is the quenching probability per encounter. Values of p from 5 x 10-5 to 3 x 10-4 are observed in 20 "C solutions (Langelaar 1969, Langelaar et al 1971a). Since RTN = RN (1 -p)/p % R N , the results indicate a low binding energy BN for the triplet excimers of the aromatic hydrocarbons. The triplet concentration quenching of phenanthrene in solution (Langelaar 1969, Langelaar et al 1971a) is unusual, since phenanthrene does not form singlet excimers or exhibit concentration quenching of its fluorescence (Birks and Georghiou 1968).

Excimers

949

If the phenanthrene contained a mole fraction f of its usual impurity, anthracene (IA), quenching would then occur by diffusion-controlled triplet-triplet energy transfer
SM"
I + 1A-~ A T P A1M +3A+

= w ( (9.12)

Since KATNKNT, f N p = 10-4 for phenanthrene (Langelaar 1969), which is a typical impurity content. The observed values of p (5 x 10-5-3 x 10-4) for naphthalene, anthracene and pyrene are of similar magnitude to the typical impurity content f of these compounds. Thus process (XXII) needs to be considered as an alternative to (XXI) in discussing the triplet concentration quenching of all the compounds (Birks 1975a). Observations on solutions of compounds of high purity (f<10-5) are required to establish the validity of process (XXI), under conditions in which process (XXII) is unimportant. Even if (XXI) is valid, the 3Do entity in (XXI) need not correspond to a triplet excimer with a finite binding energy BN, an (M ]jM)conformation, and a detectable excimer phosphorescence yield. It is more likely to correspond to a 'contact quench complex' of possible non-sandwich conformation and binding energy <AT, such as occur in other quenched solution systems (Birks 1970a, 1975b).

9.6. Halobenzenes
Several halogenated benzenes emit excimer phosphorescence. Such emission has been observed from crystals of 1,4-dichIorobenzene, 1,3,5-trichlorobenzene, 1,2,4,5tetrachlorobenzene, 1,4-dibromobenzene and 1,3,5-tribromobenzene (Castro and Hochstrasser 1966, George and Morris 1970), and from concentrated rigid solutions of chlorobenzene, bromobenzene and iodobenzene (Lim and Chakrabarti 1967). Three factors contribute to the stability and phosphorescence of the triplet excimers of the halobenzenes. (i) The energy of the charge resonance states 3(R)g, is reduced due to the decrease in I - A on halogen substitution in benzene. (ii) Heavy-atom substitution and the resultant increased spin-orbit coupling increase the So-T, transition moment m,, and thus increase the splitting of the triplet T h e heavy-atom substitution also reduces the exciton resonance states 3(Ea)g, spin forbiddenness of the excimer phosphorescence transition. (iii) The (M 11 M) crystal structure maximizes the halogen-halogen and 7r-7r intermolecular overlap, and it provides the appropriate conformation for triplet excimer formation. Values of h E ~ ~ ~ 4 cm-1 8 0 0are observed in the crystal phase (Castro and Hochstrasser 1966), and of AENT N 8900 cm-1 for the monohalobenzenes in concentrated EPA solutions at 77 K (Lim and Chakrabarti 1967). There appear to be no comparative for these compounds. data on AEDM The excimer phosphorescence of the halobenzenes is partly the result of the internal heavy-atom effect ( 49.1), although other factors, described above, also contribute. The excimer phosphorescence of benzene and other unsubstituted aromatic hydrocarbons might therefore be induced in a suitable heavy-atom solvent by the external heavy-atom effect.

950
10. Atomic excimers

J B Birk-s

There are two main classes of excimer: (i) excimers of simple atoms, and (ii) excimers of complex aromatic molecules. Atomic excimers (i) have been mainly studied in the gas phase, but they have also been observed in the liquid and solid phases. Aromatic molecular excimers (ii) occur in fluid solutions, pure liquids, pure crystals and polymers, at crystal defects, and as sandwich dimers and intramolecular excimers. Their fluorescence has not m been observed in the vapour phase, due to the rapid excimer dissociation rate k at the elevated temperatures of high-pressure molecular vapours, although the pressure quenching of the vapour molecular fluorescence provides indirect evidence of excimer formation (Stevens and McCartin 1960). The spectroscopic detection of the excimer fluorescence of aromatic vapours should not present insuperable difficulties for the more volatile compounds with modern laser techniques. Spectroscopic and theoretical studies of the helium excimer in the gas phase ($3.2) have yielded the detailed potential curves of figure 5 (Ginter 1970a). There are no corresponding data for other atomic excimers, and few definitive state assignments (Rosen 1970). Studies of the emission continua of noble gases at high pressure have been largely stimulated by two applications: their use in discharge lamps as spectroscopic sources in the near and vacuum ultraviolet (Wilkinson 1965 and references theiein); and their use as scintillators for the detection of ionizing radiation (Birks 1964a, chap 14 and references therein). Northrop and Nobles (1956a, b) discovered that the noble gases scintillate in the liquid and solid state with an efficiency similar to that in the high-pressure gaseous state, and they observed the scintillation decay times of liquid and solid Ar, Kr and Xe to be less than 10 ns. Spectroscopic studies showed that the scintillations correspond to excimer emissions (Jortner et aZ1965, Basov et a Z 1970, Packard et aZ1970, Cheshnovsky et a Z 1972a,b). A new area of research on atomic excimers was thus initiated. Noble gas excimer emission has been observed under the following conditions.

(1) Pure noble gases (pressure 2 100 Torr) excited by: (a) an electric discharge (Wilkinson and Tanaka 1955, Wilkinson and Byram 1965, Huffman et aZ1965, Prince and Robertson 1966, 1967, Kenty 1967, and others); (b) protons (Nichols and Vali 1968, Hurst et a Z 1969, Stewart et aZ1970); (c) 01 particles (Strickler and Arakawa 1964, Jortner et aZ1965, Takahashi et aZ1966, Z 1972a); Kubota et aZ1968, Cheshnovsky et a ( d )fast electrons (Basov et aZ1970); and (e)non-ionizing monochromatic photons (Freeman et aZ1971).
(2) Gas mixtures (total pressure 100-1000 Torr) consisting of a host noble gas and 102-104 ppm of a guest noble gas excited by: (a) an electric discharge (Gedanken et a Z 1972); and (b) CY particles (Cheshnovsky et aZ1972a). Excited guest atoms are produced by energy transfer from the host, and guest excimers are produced in a three-body collision with unexcited guest and host atoms [see discussion of processes (XXIII) and (XXV) below]. Electric discharge excitation of Xe/Ar and Xe/Kr gaseous mixtures yields Xe,* emission, and that of Kr/Ar mixtures yields Krz emission. Northrop and Gursky (1958) have studied the scintillation properties of binary noble gas mixtures excited by 01 particles.

Excimers

95 1

(3) Pure liquid and solid noble gases excited by: (a) 01 particles (Jortner et a1 1965, Packard et a1 1970, Cheshnovsky et a1 1972a,b); and (b) fast electrons (Basov et a1 1970).

(4)Liquid mixtures, consisting of a host noble gas and 50-104 ppm of a guest noble gas, excited by 01 particles, yield guest excimer emission (Cheshnovsky et a1 1972a,b). The mechanism is similar to that in gas mixtures (2).
Cheshnovsky et a1 (1972a) compared the Ar and Xe excimer emission spectra, excited by 01 particles, in different environmental conditions. The Ar,* spectrum (128 nm band) is practically identical in pure Ar gas at 90 K, in liquid Ar at 87 K, and in solid Ar at 80 K. The Xe,* spectrum (172 nm band) is identical at constant temperature in pure Xe gas (1000 Torr), in Xe/Ar gas mixtures (total pressure 1000 Torr for various Xe concentrations), in Xe/Ar liquid mixtures (Xe concentration 200 ppm), and in Xe/Kr liquid mixtures (Xe concentration 5000 ppm). The excimer emission band broadens with increase in temperature (cf figure 7). The data have been analysed in a similar manner to those for the pyrene crystal excimer described in 33.3. Vibrational frequencies of 140 and 310 cm-1 are thus obtained for Xe,* and A r t respectively (Cheshnovsky et a1 1972a). Aromatic molecular excimers normally obey Kasha's (1950) rules in that the emission occurs only from the lowest excited state of a given multiplicity, ie IDx or 3D*. This is due to the high density of vibrational molecular and excimer states, and the resultant rapid internal conversion within the excited singlet and triplet excimer manifolds (see Birks 1970a). An atom has no vibrational states, and it exhibits all radiative transitions allowed by the selection rules. An atomic excimer has the simple vibrational and rotational modes of a symmetric diatomic molecule. Its emission spectrum is similar to that of such a molecule (Herzberg 1950)) apart from the continua due to transitions to the dissociated ground state. Kasha's rules do not apply to atomic excimers. Subject to the selection rules transitions can occur from IDx and IDxx states and from 3Dx and 3Dra states to the dissociated ground state yielding structureless emission bands, and from higher ID** and 3Dxx states to lower ID** (or ID") and 3Dxx (or 3Dx) yielding emission bands with vibrational and rotational structure. The multiple emissions of the atomic excimer states provide a rich source of spectroscopic data (figure 5), but the spectral analysis is complex, particularly with excitation in a gas discharge or by ionizing particles. Of the many states of He2 shown in figure 5, only the 1 X : state has an allowed transition to the dissociated, singlet, even (g) parity, ground state, and it is responsible for the ID" emission. T h e excited states of other noble gas excimers are similar to those of He2. Increase of the atomic number 2 increases the spin-orbit coupling, which produces some relaxation of the multiplicity selection rule. Wilkinson and Tanaka (1955) have proposed that the xenon emission continuum, as excited in a discharge, is composite with contributions from the 3 2 $ , 31Tu, I C : and In,, excited states of Xe2. These states are probably those responsible for the other noble gas excimer emissions, with the triplet excimer emissions becoming more probable as 2 is increased. T h e relative intensities of the different excimer bands depend on several factors. (a) The mode of excitation determines the initial population density of the excited states of the atom. Excitation by a spark discharge or by ionizing particles populates many excited singlet and triplet states of the atom, and it also yields doublet ions and
63

952

J B Birks

electrons. Low-energy electron excitation in a spark discharge enhances the yield of excited triplet states; high-energy particle excitation favours the initial formation of excited singlet states. The subsequent recombination or interaction of these entities changes the population distribution and generates further species, including excimers and dimer ions. These processes are further influenced by the presence or absence of an electric field. These various factors explain the major differences observed in the emission spectra of He, Ne, Ar, K r and Xe high-pressure gases when excited by 4 MeV protons and when excited in a gas discharge (Stewart et all970). One major problem with excitation by a spark discharge or ionizing radiation is the formation of M,f ions. Such ions are metastable and emit continua which may obscure, or be mistaken for, excimer emissions. The ideal solution is to use nonionizing radiation for excitation. Individual atomic states, normally singlet, can be populated by monochromatic photons, and individual triplet or singlet atomic states can be excited by monoenergetic electrons. Only one emitting excimer state is then produced, and the system is more amenable to precise study (Freeman et al 1971). (b) The environmental conditions (state, pressure, temperature, atomic mass of gas, electric field, etc) determine the rate of subsequent collisional processes which lead to energy redistribution and changes in the excited state population, and to excimer formation and stabilization. ( c ) The potential curve of each excimer state determines its stability and whether there is a barrier to excimer formation. The helium excimer has many bound states, but the formation of several of these is inhibited by a potential barrier (figure 5). (d) The lifetime of an excited atomic state influences the probability of its forming an excimer. This factor favours the triplet atomic states and the formation of triplet excimers, but not necessarily their emission. ( e ) Radiative transitions to the dissociated ground state occur only from singlet, odd (U) parity, excimer states, except in the high-Z atomic excimers, where spinTable 3. Maxima (in nm) of noble gas emission continua, Probable excimer emissions shown in italics. Origins of other emissions uncertain
Atom State Ne Gas (50-200 Torr) Liquid (25 K) Solid (14 K) Solid (6 K) Ar Gas (50-2000 Torr) Gas 50 1400Torr) Gas t90-K) Liquid (87 K) Ar/Ne solid soln. (6 K)
Excitation
18 keV B s 18 keV p s

4 MeV ps

Kr

Xe

Gas (50-400 Torr) Kr/Ar gas mixture (100 K) Li uid (120 K) KJAr liquid mixture (80 K) Kr/Ne solid soln. (6 K) Solid Solid 110 K) Solid 16 K) Gas (50-300 Torr) Gas 500Torr Gas [lo0 To?] Xe/Ar as mixtures Xe/Ar fquid mixtures Xe/Kr liquid mixtures Liquid (165 K) Xe/Kr solid soln. 110 K) Xe/Ne solid soln. { 6 K)

5 MeV as 5 MeV as 4Me V ps 5 MeV as 5 MeV as 5 MeV as 5 MeV as 18 keV Bs 5 MeV as 4 MeV ps 5 MeV as 5 MeV U S 5 MeV US 5 MeV US 300-900 keV es 5 MeV as 5 MeV as 4 MeV ps 147 nm photons 148 n m hotons 5 Me$ as 5 MeV as 5MeV a s 300-900 keV es 5 MeV US 5 MeV as 300-900 keV es 151 nm hotons 5 Me8 as

83 98 77.4 90 77.4 89 77 I30 128 I28 128 128 128 127 I28 143 147 148 148 148 147

Observers Stewart et a1 (1970) Packard et a1 (1970) Packard et a1 (1970 Gedanken et a1 (1923a) . . Strickler and Arakawa (1964) Hurst et a1 (1969) Cheshnovsky et aZ(1972a) Cheshnovsky et aZ(1972a) Gedanken et al(1973a) Cheshnovskv et a 2 f1972a) Packard et d (1970) Gedanken et aZ(1973b) Stewart et al (1970)

148.4 168 168 172 I72 172 172 176 I72 172.5 I75 176 175.5

148 _.

160

Stewart et a2 (1970) Freeman et a1 (1971) Fink and Comes 1975) Cheshnovsky et U! 19728) Cheshnovsky et aZ11972a) Cheshnovsky et a 1 (1972al Basov et aI (1970) Cheshnovsky et al (197313) Gedanken et aZ(1973a) Basov et a Z (1970) Brodmann et a 2 (1974) Gedanken et a Z (1973b)

Excimers

953

orbit coupling relaxes the multiplicity rule and transitions from triplet u-states can also occur. Table 3 summarizes data on the emission continua observed in the noble gases excited optically or by high-energy ionizing particles. For corresponding data on the gas discharge spectra reference should be made to Wilkinson (1965), Rosen (1970) and Stewart et a1 (1970). Hurst et a1 (1969) assigned the 183 and 214 nm continua in Ar, which do not appear in the gas discharge spectrum, to excimer states with binding energies of about 4 eV. They attributed the 110 and 128 nm bands, which do occur in the discharge spectrum, to 'dissociative radiative continua', ie radiative two-body collisions (see below). However, the observation of the 128 nm continuum in the gas, liquid and solid phases shows that it is definitely due to an Ar,* state (Cheshnovsky et a1 1972a). The optical excitation of the 168 nm band in Xe demonstrates that it originates from an Xe,* state (Freeman et al 1971). Other emissions listed in table 3 probably originate from excimers, but some may be due to molecular ions or radiative two-body collisions. I n contrast to the detailed picture of the helium excimer (figure 5), four decades of conventional spectroscopy of the other noble gases have yielded surprisingly little information about the origins of their emission continua. The data have come mainly from less conventional studies. I n the process M*+M+D* (+ B ) (IIB)

D" formation occurs only if the whole or part of the excess energy B is removed during the short collisional time T C in which the potential energy is less than at infinite separation (figure 4). The time TC( 10-13 s) is of the order of the vibrational period of D". I n a condensed medium the excess energy B is rapidly dissipated as thermal energy to the environment, thus stabilizing D". This mechanism is not available in the gas phase. Here the removal of energy may occur in two ways (Herzberg 1950): (i) radiation of the whole or part of the total energy during the collision time (radiative recombination by a two-body collision); or (ii) collision with a third atom or molecule during the collision time (recombination by a three-body collision). Radiative recombination in a two-body collision (i) results in a continuous spectrum (the recombination continuum). It is, however, a very rare process, since the D" radiative lifetime TFD( 2 10-8 s) 9 TC( 10-13 s), the collisional time, so that radiative recombination occurs for only a very small fraction ( Q 10-5) of the collisions. The process has been observed in halogen and tellurium vapours. The emission spectrum of iodine (12) vapour at high temperature includes both discrete bands and a continuum (Kondratjew and Leipunsky 1929). Thermal dissociation ( 1 ~ ~ 2 yields 1) iodine atoms (I), some of which are electronically excited. I n a small fraction of the cases the atoms recombine by a radiative two-body collision (i), instead of by a threebody collision (ii), and the former gives rise to the continuum emission. The continuum emitted by tellurium (Tez) vapour in an intense discharge, in which strong dissociation occurs, probably originates in a similar manner (Rompe 1936a,b). The three-body collision process (ii)

M* +M+M+D++M

(XXIII)

is a much more probable process than (i) in gases at medium and high pressures (Herzberg 1950), and it is the process primarily responsible for D" formation and

954

J B Birks

emission in the noble gases. Colli (1954) observed the luminescence response function of the A r l (128 nm) emission as a function of pressure P (in Torr). He found it to be of the form (10.1) iD(t )K exp ( - t/71) - exp ( - t / 4 which is similar to relation (4.24). The parameter 72( = 3.4 p )is independent of P and is equated to the D" lifetime m, while ~ 1 = O . l l P - 2 corresponds ~ to the D" formation time. The inverse quadratic dependence of 71 on P is consistent with the three-body process (XXIII). BouciquC and Mortier (1970) analysed the emission from Townsend discharges initiated by a particles in Ar, K r and Xe at P= 50-600 Torr in terms of relation (10.1). For Ar;, 72( = m) =3.7 ps, AI( = 1/71) = 6.9P2 s-1; for Kr:, TD = 1.7 ps, A1=46P2s-l; and for Xe;, -=Os5 ps, Al=(26P2+7500P) s-1, the linear term in P being attributed to two-body Xe" quenching collisions. Thonnard and Hurst (1972) studied the time dependence of the 110 and 125 nm emission continua of Ar excited by 250 keV electron pulses. The 110 nm emission decays exponentially, consistent with a two-body radiative collision. The 125 nm Ar; emission agrees reasonably with (10.1) with 71N 0.1 P2 s, 72 = 2.8 p, although a detailed analysis suggests that two Ar? states, produced by independent channels, may contribute to the emission. Oka et aZ(1974) used pulse radiolysis, with excitation by 10 ns pulses of 500 keV electrons, to study the time dependence of the transient excimer absorption in Ne, Ar and Kr at P= 200-1400 Torr. The results agree with relation (10. l), with ~ 1 c c P - 2 and 72(=7~)=6*6,3.2 and 0.35 ps for Ne:, Ar; and Kr; respectively. The values of T D are consistent with the assignment of the long-lived Ne;, Arz and Kr; states to the excimer state (Oka et al 1974), and a similar assignment is commonly assumed for the Xe; state. The decrease of m with increase in 2 is attributable to increased spin-orbit coupling. The corresponding IC: states, which have much shorter radiative lifetimes, emit in the same spectral region as the 3 C z states, and nanosecond time-resolved spectroscopy is required to distinguish them. Unless this is done, the practice of Jortner and co-workers in describing the unresolved excimer spectra as originating from 1?3X$ states is preferable. Much smaller values of Q have been observed for the emission from gaseous and liquid He ( T D = 14-25 ns), and for other excimer emissions from gaseous Ne ( ~ ~ = 5ns *5 at P=760 Torr) and liquid and solid Ar, Kr and Xe ( T D < 10 ns) (see Birks 1964a). These short-lived emissions are attributed I C : ) states, or to mixed 193D" states containing a large ID* component. to ID* ( The 172nm Xe,* emission band has attracted recent attention as a potential laser source. Koehler et aZ(1972) observed stimulated Xe,* emission from Xe gas at P > 200 psi when excited by intense pulses of 0.6-2 MeV electrons. The emission bandwidth narrows from 15 to 1.7 nm, the pulse width decreases from 50 to 3 ns, and the lifetime is reduced from ~ ~ = to 2 0 - 2 ns. Wayne Johnson and Gerard0 (1973) studied the decay of the Xe; emission excited by 3 ns pulses of 0.6 MeV electrons. At low excitation intensities T D - ~ increases linearly with P and extrapolates to a zero-pressure value of m = 5 0 ns. These values of ~ = 2 0 - 5 0ns are much less than that of T D = 500 ns observed by BouciquC and Mortier (1970), and they suggest the presence of two adjacent Xez states of mainly singlet and triplet character respectively. Theoretical calculations by Weihofen (1974) yield a value of m =23 k 8 ns for Xe,*, in good agreement with the experimental values for the shortlived state. The calculations indicate that the latter is a mixture of $ singlet and

Excimers

955

triplet, so that its description as 19 3E$ is more appropriate than that of 3EC,+which is commonly used. The long-lived Xe,* state might then be described as 3 * 1 l 7 . The emission spectrum of Hg vapour at medium pressure, excited by 3 ns pulses of 500 keV electrons, has three continua centred at 350, 485 and 540 nm respectively (Eckstrom et aZl973). The 350 and 485 nm continua are due to the transitions from the 3Hg;" (31u) and 3Hg: (30;) states to the dissociative (IC:) ground state. The 540 nm band, which exhibits vibrational structure under certain conditions (Stupavsky et al 1972), is attributed to the collision-induced 32,-330; excimer transition (Phillips 1973). As P i s increased, the intensities of the 350 and 540 nm bands decrease relative to that of the 485 nm band, and the band maximum of the latter shifts to 457 nm, possibly due to more complete vibrational relaxation (Eckstrom et a Z 1973). The radiative lifetime of the 457 nm emission is 14 k 3 p, independent of P from 0.1 to 10 atm, and the 350 and 540 nm emissions have similar lifetimes. At high P the 457 nm emission decay becomes non-exponential due to the bi-excimer collisional deactivation of Hg;, Hgg Hgg-'Hg,*" 2Hg

analogous to the bi-excimer interactions observed in aromatic molecular excimers in crystals and solutions ( 98), and the kinetic behaviour is described by (8.1) and (8.2). A similar bi-excimer collisional deactivation of Xe,* has been observed in Xe gas at high P and high excitation intensities (Wayne Johnson and Gerard0 1973). Many of the techniques and concepts used in the study of aromatic molecular excimers can be applied to atomic excimers in the high-pressure gas, liquid, solution and solid phases. (a) The rate parameters and thermodynamic parameters can be determined from luminescence quantum yield and lifetime measurements ( 94). The aromatic excimer reaction kinetics of 94 may be applied to atomic excimers by substituting kDMM [1M]2, the rate of the termolecular excimer formation process (XXIII), for ~ D M [ ~ M the ], rate of the bimolecular excimer formation process (11), and by omitting the molecular and excimer radiationless decay processes. (b) Diffusion-controlled processes can be studied in the gaseous, solution and liquid phases ( 94.5). (c) Different modes of atomic excimer formation, analogous to those discussed in 95, can be considered. There is the possibility of delayed atomic excimer fluorescence in fluid solutions or in the liquid phase, and of associated magnetic field effects (95.3). ( d ) Photophysical processes in atomic excimers can be studied (97). Although radiationless processes, other than predissociation, are inhibited in the gas phase, the excimer radiative lifetime and its transient absorption transitions, observed by nanosecond flash photolysis, may assist in state assignment. Vibrational relaxation occurs in the liquid and solid phases, and there is the possibility of internal conversion and inter-system crossing between adjacent atomic and/or excimer states. One interesting observation is the reduction of the scintillation efficiency of liquid helium at temperatures below the A transition (Moss and Hereford 1963), suggesting that excimer formation may be inhibited in helium 11. (e) Excimer excitons may exist in crystals of noble gases. They can be studied either with doped crystals or with pure crystals under intense laser excitation (98). (f)Triplet atomic excimers (99) are common in noble gases of high atomic number, because of the intrinsic spin-orbit coupling, but further studies are required to distinguish them clearly from singlet excimers.

956

J B Birks

The most immediate problem is the identification and assignment of the various singlet and triplet excimer emissions in the noble gases. Monoenergetic photon or electron excitation provides a useful tool to unravel the complexities. Kinetic studies of the atomic and excimer luminescence provide a means of studying the excimer formation and radiative decay. Paradoxically, the physics of simple atomic excimers is complex, while the physics of complex aromatic molecular excimers is relatively simple. I n this instance the physicist's preference for simple systems has proved illusory : atomic excimers are more complicated than aromatic molecular excimers. Kasha's rules explain the paradox: the high density of vibrational states of a complex molecule simplifies its luminescence behaviour. No such simplification occurs in an atomic system.

11. Other excimers


Excimer formation is not restricted to the aromatic hydrocarbons and their derivatives (table Ala) and monatomic gases. Excimer fluorescence is also observed from several heteroaromatic compounds in which a nitrogen (N) and/or oxygen (0)atom is incorporated in a five- or six-membered ring (table Alb). Unlike the unsubstituted aromatic hydrocarbons, many of these molecules have a finite electric dipole moment, and the influence of this on the ID" formation and conformation has been discussed by Lami and Laustriat (1968). The substituted oxazoles and oxadiazoles, PPO, BPO, a-NPO, cl-NPD and PBD (table Alb), are commonly used as scintillator solutes, often at concentrations at which the ID* fluorescence is significant (Birks 1964a). Neither aromaticity nor heteroaromaticity is necessary for excimer formation by organic molecules. Some saturated organic amines exhibit excimer fluorescence in solution and in the vapour phase (Halpern and Maratos 1972). A detailed study of excimer formation by one of these compounds, 1-azabicyclo [2.2.2] octane (ABCO), in the vapour phase has been made by Halpern (1974). The fluorescence spectrum comprises a structured 1M* emission band (250-310 nm) and a structureless ID" emission band (Dm= 370 nm). Halpern (1974) analysed observations of the 1M" ) i ~ ( tas ) a function of pressure P and ID" fluorescence response functions i ~ ( tand in terms of the reaction kinetic scheme of $4.2, with the additional ID* quenching process ID" + 1M+3 1M (or otherwise) (XXIV) of rate parameter kQD. The following rate parameters were obtained for ABCO vapour at 26 "C with hex = 265 nm: RM = 2.75 x 106 s-1, RD = 2.27 x 106 s-1, RDM = 3.1 x 1010 M-1 S-1, K M D = 1.3 x 105 S-1, RQD = 8.4 x l o 9 M-1 S-1. Assuming an effective collision diameter of 6 A, the value of KDM corresponds to p -0.2, which is smaller than for ABCO in n-hexane solution. I n the vapour p < 1, since ID" formation requires a specific orientation of the two molecules, while in solution the cage effect can increase the value of p . The other rate parameters for ABCO in n-hexane (RM = 1.8 x lo7 s-1, RD = 5.6 x 106 s-1, RQD = 4.6 x l o 7 M-1 s-1) also differ from those for ABCO vapour. The addition of n-hexane vapour as a carrier gas (pressure Pc = 0-150 Torr) to ABCO vapour (P= 1-6Torr) enhances the excimer fluorescence yield @FD. The enhancement is a maximum at PC 90 Torr, where @FD is increased by a factor of 4 with excitation at hex= 249 nm ( N Sy), and by a factor of 10 with hex= 228 nm

Excimers

957

( N- 8 : ) . Collisions with n-hexane molecules induce (a) vibrational relaxation within the singlet manifold of the excited ABCO molecule (hence the dependence on A,,) and (b) vibrational relaxation and stabilization of the excimer. The fluorescence behaviour of pure ABCO vapour is consistent with a two-body ID* formation process

(IIB) and there appears to be no need to invoke a three-body collision process (XXIII) as in the noble gases ($10). The ABCO excimer contains 42 atoms, and internal redistribution of the excess energy B (- 0.43 eV) among the many vibrational modes is probably responsible for the initial ID* stabilization. The excess energy B can, however, only be dissipated by radiation, energy transfer or collision. The enhancement ~ the addition of n-hexane vapour is due to collisional dissipation of the of @ ' p by excess energy. The collision may occur after the formation of ID" by process (11). Halpern (1974) has, however, obtained evidence that KDM may increase with the nhexane pressure Pc. This suggests the alternative possibility of a three-body ID* formation process, catalysed by a carrier gas molecule IC, (XXV) in which 1C takes up the excess energy in the same manner as the second 1M atom in process (XXIII). Process (XXV) can also occur in atomic and aromatic molecular systems in the gas phase. The addition of a carrier gas (IC) provides an alternative means to process (XXIII) for removing the excess ID* energy and thus stabilizing the excimer.
12. Exciplexes, D A complexes and mixed excimers

1M*+ 1 M S D * ( + B )

1M*+ 1M + 1C+D*

+ 1C

12.1. Exciplexes and DA complexes


Excimers belong to a much larger group of physical entities known as exciplexes. An exciplex is defined as an electronically excited atomic or molecular complex of definite stoichiometry, which is dissociative or dissociated in its electronic ground state (Birks 1969b, 1975a). T h e stoichiometry is usually 1 : 1, but triple 2 : 1 exciplexes have also been observed (Beens and Weller 1968). An excimer is a 1 : 1 homopolar exciplex between identical atoms or molecules. The first aromatic molecular exciplex was discovered by Leonhardt and Weller (1961, 1963), who observed exciplex formation between excited perylene and unexcited dimethylaniline (DMA) molecules in solution, yielding a structureless exciplex fluorescence band analogous to excimer fluorescence. Since then numerous other aromatic molecular exciplexes have been observed and studied. The subject has been reviewed by Birks (1970a, 1975b) and Beens and Weller (1975), the latter from a more theoretical standpoint. Exciplexes, in turn, are particular examples of stoichiometric atomic or molecular complexes. Research on molecular complexes led to and, to a large extent, was stimulated by the Mulliken theory of intermolecular charge transfer (CT) interaction (Mulliken 1950, 1952a,b, 1963, Mulliken and Person 1962, 1969). Molecular complexes are generally formed between electron-donor molecules (D) and electronacceptor molecules (A), with the CT state (D+ A-) playing an important role in the interaction. Molecular complexes can be classified according to their ground-state

958

J B Birks

binding energy AEo (Birks 1970a). Complexes which are stable in the ground state (AEo2 KT) are known as donor-acceptor ( D A ) complexes (Briegleb 1961). Complexes with kT> ARo2 0 are called contact D A complexes. ExcZplexes, by definition, have AEo < 0. The complex between an aromatic hydrocarbon X (= D or A) and another molecule Y ( = A or D) is referred to as an XY complex. I n its ground state an aromatic hydrocarbon normally functions as an electron donor D and it forms DA complexes. Excitation of a molecule reduces its ionization potential and increases its electron affinity, so that an excited aromatic hydrocarbon may act as an electron acceptor A and form an AD exctplex (Birks 1970a) with a suitable donor D. Pyrene, perylene and other aromatic hydrocarbons form AD exciplexes with aniline, DMA and similar donor molecules. Substitution of one or more cyano groups in an aromatic hydrocarbon markedly increases the electron affinity, so that 9-cyanoanthracene (CNA)and 9,lO-dicyanoanthracene (DCNA) are strong electron acceptors. I n consequence, phenanthrene, pyrene and other aromatic hydrocarbons act as electron donors in forming DA exciplexes with CNA or DCNA. The aromatic hydrocarbon excimers represent the intermediate case between DA and AD exciplexes, in which the hydrocarbon is both donor and acceptor. The CT interaction, mainly responsible for DA complex and AD and DA exciplex formation, becomes in the excimer case the charge resonance (CR) interaction ($2.2). It is supplemented by and configurationally mixed with the exciton resonance (ER)interaction ( $2.1), which is absent in complexes formed between dissimilar molecules.

12.2. Mixed excimers


Birks (1970a) proposed that the term mixed excimer be used to describe 1 : 1 exciplexes formed between pairs of molecules MA, ~ M Bof ) similar, but different, species, such as molecules with a common aromatic core, Other authors have used the terms 'mixed excimer' and 'hetero-excimer' as synonyms for 'exciplex', but this obsolete and confusing terminology is not recommended (Birks 1969b). The core of an aromatic molecule primarily determines its n electronic energy levels. Alkyl or similar substitution causes only a small perturbation of the energy levels, so that the components of a mixed aromatic molecular excimer ~(MAMB)* are approximately in resonance. This is not so for other exciplexes. The formation of mixed aromatic molecular excimers in solution by the diffusioncontrolled processes 1MZ ~MB+~(MAMB)" (XXVIA)

MA + ~MZ$(MAMB)+

(XXVIB)

was first observed by Birks and Christophorou (196213) in the following ~ M A / ~ M B systems: pyrene (Py)/l-methyl py ; 1-methyl ~y/4-methylpy ; 1:2-benzanthracene ( ~ ~ ) / 6 - m e t hBA; y l and 5-methyl ~ ~ / 6 - m e t hBA. y l Selinger (1964) observed the mixed excimer fluorescence of py14-methyl py. Neznaiko et al (1966) observed the mixed excimer fluorescence of the following systems: anthracene (~)/g,lo-methylA ; and ~/9,10-di-n-propylA (DPA). Obyknovennaya and Cherkasov (1967) determined the binding energy B of the foIIowing mixed excimer systems : methyl A ( B= 0.20 ev) ; DPA/9-n-propyl A ( B= 0.22 ev>; DPA/l-a-oxyethyl A ( B= 0.22 ev); ~ ~ ~ / 9 - a - o x y e t h Ay( lB = 0.20 ev) ;

Excimers

959

and ~ ~ ~ / 9 - a c e t A o( xB y= 0.19 eV). The values of B are similar to those of the excimers of 9-n-propyl A ( B= 0.23 eV) (Obyknovannaya and Cherkasov 1967), 9-methyl A (B=0.23 ev), and 9,lO-dimethyl A (B=0.19 eV) (Barnes and Birks 1966), but they are less than that of the excimer of DPA (B=0.46 eV) (Bazilevskaya and Cherkasov 196513). Anthracene and its derivatives also form mixed photodimers, and photolysis of these in rigid solution yields mixed sandwich dimers. The fluorescence of the mixed sandwich dimer of ~/9,10-dichloroA has been studied by Chandross and Ferguson (1966a) and Fielding and Jarnagin (1967). Ferguson et a1 (1973) have observed the fluorescence lifetime T D and spectrum of the following mixed sandwich dimers in rigid solution at 77 K: A/g-phenyl A ( T D = 115 ns); ~/9-methylA ( T D = 130 ns); A ( T D = 105 ns); ~/9,10-dichloro A ( T D = 100 ns); and ~/g,lO-dimethyl A ( ~ ~ = ns). 8 0 The values of TD are about half those of the pure sandwich dimers of A (TD= 200 ns), 9-phenyl A ( T D = 260 ns), 9-methyl A ( T D = 220 ns), and 9-cyano A (TD= 170 ns) (Ferguson et a1 1973). The steric factors, which influence photodimer and excimer formation by the meso-substituted anthracenes (96), will have a similar effect on the mixed excimers, photodimers and sandwich dimers of these compounds. Intramolecular mixed excimer fluorescence, due to the interaction of a phenyl group with an o-tolyl or m-tolyl group within the same molecule, has been observed by Hirayama (1965). Thus several of the phenomena associated with 'pure' aromatic molecular excimers have also been observed in mixed aromatic molecular excimer systems.

cyano

12.3. Atomic exciplexesand complexes


Noble gas atoms may form 1 : 1 exciplexes (AEocO) or complexes (AEo>O) with several other types of atoms, including (i) atomic hydrogen (H), (ii) other noble gas atoms (He, Ne, Ar, Kr, Xe), (iii) alkali metals (Li, Na, K, etc), (iv) mercury (Hg), and (v) oxygen (0). Calculations of the electronic states of HeH show the ground state to be dissociated and the excited states to be associated, analogous to He2 (figure 5 ) (Michels and Harris 1963, Bender and Davidson 1966). The emission spectrum of HeH consists of structureless bands in the far ultraviolet due to transitions to the dissociated state, and structured bands in the visible and near infrared due to transitions between the associated excited states. Similar exciplexes of NeH, ArH, KrH and XeH are to be expected on theoretical grounds, but to date only ArH has been identified unambiguously from its emission spectrum (Ginter 1970b). The noble gas exciplexes, HeNe, HeAr, HeKr, HeXe, NeAr, NeKr, NeXe, ArKr, ArXe and KrXe are predicted to have dissociated ground states and associated excited states. Emissions attributed to (HeNe)", (ArXe)" and (KrXe)* exciplexes have been reported from gaseous mixtures (Rosen 1970), and (HeNe)" has been observed in absorption (Tanaka and Yoshino 1972). The (KrXe)" emission band at 158nm, intermediate between the Kr,* and Xe,* emission bands at 148 and 172 nm respectively, has been observed in liquid and solid Xe/Kr mixtures (Cheshnovsky et a1 1973b). The (ArKr)" emission band at 135 nm, intermediate between the Ar,* and Kr,* emission bands at 127 and 148 nm respectively, has been observed in gaseous, liquid and solid Kr/Ar mixtures (Cheshnovsky et a1 1973a). From these and other emission data on noble gas mixtures (Gedanken et a1 1973b), the authors conclude that NeAr, NeKr, NeXe and ArXe do not form stable exciplexes.

960

J B Birks

Calculations of the electronic states of HeLi indicate that only three of its excited states (4II and the two lowest 2C+ states) are associated, the remainder being repulsive (Schneiderman and Michels 1965). It is expected that only a few of the excited states of the noble gas/alkali metal exciplexes will be bound (Ginter 1970b), and until recently spectroscopic data on these exciplexes were lacking (Rosen 1970). Hedges et al (1972) have studied the emission spectra of the Cs/noble gas exciplexes, and Drummond and Gallagher (1974) those of the Rb/noble gas exciplexes. The metal vapour is excited by an atomic resonance line in the presence of 400 Torr of He, Ne, Ar, Kr or Xe. The exciplex emission is a structureless band extending up to 100 nm from the resonance line, and its temperature dependence has been analysed to determine the ground- and excited-state potentials of the exciplex. The addition of a high-pressure noble gas to mercury vapour similarly introduces a structureless band in the vicinity of the atomic lines in both the absorption and fluorescence spectra (Oldenberg 1929, Kuhn and Oldenberg 1932, Kuhn 1937). The bands are due to weakly bound HeHg, NeHg, ArHg, KrHg and XeHg complexes. Theoretical calculations of these complexes have been made by Heller (1941), and the spectroscopic analysis of the bands has been discussed by Herzberg (1950). The high-frequency discharge of xenon containing a trace (0.05%) of oxygen shows an unusual band system at 4900-5577A which is attributed to the XeO exciplex (Kenty et aZl946). Similar spectroscopic evidence has been obtained for the exciplexes of H e 0 (Weniger and Herman 1948), ArO (Cooper and Lichtenstein 1958) and KrO (Cooper et al 1961). Tisone (1974) has observed the lifetime (50 ps) and quenching rate constants for the green emission of XeO. Golde and Thrush (1974) have observed vacuum uv chemiluminescence from (ArO)", (KrO)" and (ArCI)" c 1 4 . exciplexes formed by the quenching of noble gas atoms by N20, 0 3 , Cl2 and c

13. Conclusion
This review concludes where it began, with aromatic molecules and atomic gases and their common property of forming excimers, exciplexes and complexes. Aromatic molecules are complex in structure but have simple molecular and excimer fluorescence spectra, studies of which have yielded a wealth of data about the photophysics of aromatic excimers. Atoms are simple in structure but have complex emission spectra, which include continua due not only to excimers but also to dimer ions and radiative two-body collisions. With the notable exception of helium (figure 5), the spectroscopic assignments of the noble gas continua are incomplete, and further photophysical studies are required to complete the task. It is a sad commentary on the fragmentation of science that the extensive bibliography includes the names of only six people who have been interested in both atomic and aromatic excimers. Perhaps this article will help to remove the artificial interdisciplinary barrier between other workers in the two fields. They have much to learn from each other.

Acknowledgments
Finally I wish to thank my colleagues and former students, named in the bibliography, who have shared my interest in excimers. Without their efforts the contents of this article, and possibly its authorship, would have been different.

Exciws
Appendix
Table Al. Conditions of observation of excimer (ID*) fluorescence and photodimer
formation by aromatic compounds

961

(M2)

Excitation. opt. =optical. el. =intense electron beam. State. soln.=fluid solutihn; liq.=pure liquid; cryst.=qingle crystal; microcryst. =microcrystals ; def.=crystal with high defect concentration; intra. =intramolecular; sand. =sandwich dimer.

(a) (1)

Compound Structure Condensed aromatic hydrocarbons and derivatives Benzene (E)

Conditions

References

Toluene

. j :
5 4

lD* (opt.) soln., liq. Iv62, Bi65 Ly67, B166b, H169a b Cu72d, br72 intra. Hi65 (el.) liq. Ca67, Ch68

l==CHs

lD* (opt.) s o h , liq. Iv62 Bi65


intra.

o-xylene m-xylene p-xylene Mesitylene 1,2,3,5-tetramethyl B Ethyl B 1 4-diethyl B 1'4-dimethyl 2-ethyl B 1:3-dimethyl: 4-ethyl B n-propyl B Isopropyl B Isobutyl B n-pentyl B (2) Naphthalene N)

1,2=CH8 1,3=CH8 1 4=CHS (3,s =CHI


1 2 3 5 = CH, 1l'dA

liq. liq.

lis.
liq.

lD* el.) liq. ID* el.) li 'D*


ID* el. ID* el. liq.

[zff) sa.,
liq. liq.

4s.

Lu65, B16ib, Gr68, B169 Hi65 Ca67, Ch68 Lu67 Ch68 Lu67 Ch68 Ch68 Bi65 68b Ch68 Ch68 Lu67 Ca67 Ch68 Ch68' Ch68 Ch68 Ch68 Lu67 Ch68 Ch68 Ch68

' D e (opt.) soin.


def.

1-methyl N

intra. 1=CH,

D662b, Al65, Ma65, Se66, Sm66 Jo65, Be71, Ka72 Ch70a,b St63a A165 Se66 'Sm6$, cu7la St63a Ka72 Ch68 St63a 64 A165 be6'6, Sm6b St63a, Bi64a A165 Sm66 Bi649, Ch66d Ch68 A165 Sm66 A165' Sm66 A165' Sm66 A165: Bi66b, Se66 Sm66 Bi64;, 66b Ka72 Ch68 Sm66 A165 Sm66 Al65' Sm66 Bi64; A165 Se66, Sm66 Ka72 A165 A165, Sm66 A165 A165 Ch66d Ch68 St63a, 64 St63a St63a St63a

ID*(opt.) soln.
liq. def. (el.) li ID* (opt.) A n . li ID* (opt.) s & . ly. ID* kin. lD* opt. soh. lD* (opt.) soln. lD* (opt.) soh. lia

2-methyl N 1,2-dimethyl N
1 3-dimethyl N 1'4-dimethyl N 1'5-dimethyl N

2=CH8 1,2=CHa

{ZkLj

1:6-dimethyl N

d 2
1,7-dimethyl N 1 8-dimethyl N 2:3-dimethyl N 2,6-dimethyl N 2 7-dimethyl N 2'3 5-trimethyl N l%hyl N 2-ethyl N 1-fluor0 N 2-flUOrO N 17-CH 1:8=CH: 2,3=CH; 2,6=CH; 2 7=CHa 2'3 5=CHs 1)=)C,H6 2=CaH,
1= F

lD*t;!

1D* (opt:] soln. lD* (opt.) soln. lis. lD* (opt.) soh. def.

k.

2=F

%D* (opt.) 8%.

lis.

962
Compound Acenaphthene

J B Birk
Structure

H C---CH,

7'

Conditions lD* (opt.) s o h .

References St64

(3.1)

Anthracene

(A)

All (opt. below def. ID* : microcryst. def. sand. soln.


$2

M,

: S A .

~171
j065

Ca59 Ch59

Pe64

M: : soln.
ID* : soln. sand. 9-n-propyl A 9-methoxy A 9-acetoxy A 9-A carboxylic acid 1-chloro A 9-ChlOrO A 9-bromo A 9-cyano A 9=CaH7 9=OCH8 9 = COOCHS 9 = COOH i=c1 9=C1 9=Br 9-CN

;: ; :

M, : soh.

ID* : soln. Me : soln.


MI : s o h .
ID* :soln. ID* : soh.

Ma :soh.
lD* : s o h . ID* :microcryst. M .soln. lD'. Isand. M, : soh. ID* : sand.

MI : soh.
def. lD* : soh. cryst.

9-methyl, 10-methoxy A 9,lO-dimethyl A 9 10-di-n-propyl 9'10-dichloro A 9:lO-diphenyl A


A

9==CHa,10=OCH, 9,10=CH8 9,10=C3H, 9,lO-Cl 9,10=

MI : soln.

sand.

ID* : soln. ID* : soh. ID* : soln. ID* :microcryst. S o h : no fluor. self-quenching

5t63b Ch66c Ch66c, Fi67, Fe73 Ch66c Ch66c, Fi67, Fe73 Ca59, La60, Cab5 Cr66, Co69a,b, Ka70 '2071 Ch66b St63b, Cr66, Co69a,b, 71 Chbbb, Fe73 La60 Ca65, Chd Ch65 Ba65c, Ch65, Ba66 Ba65b, Ch65 5t63b Me61

Ch66a b c Mo66' ' Ca59 La60 Ca59' La60 Ca59' La60 Ca65' A173; Bi63& Ba66, A173a Ch66a,c, Fi67, Fe73 Ca59, La60, Cab5 0b67 Ca59, LabO, Cab5 0b67 Ca59, La60, Cab5 Ba65a Ca59, La60, Cab5 Ba65a

(3.2)

Phenanthrene (ph)

S o h : no fluor. self-quenching

Bi68c

(4.1)

Pyrene (py)

43
1

ID* : soh.

cryst.

liq.

Excimers
Compound 1-methyl py 3-methyl PY 4-methyl py 3-ChlOrO py 3-bromo py 3-cyano Py 3-py Na sulphonate 3 5 8 10-Py tetra sa'shphonate (4.2) Tetracene Structure l=CHI 3=CH8 4=CH8 3=C1 3=Br 3=CN 3=S08Na 3,5,8,10= SO,Na Conditions lD* : soh. ID* : soh. ID* : soh. microcryst. lD* :soh. microcryst. ID* : soln. microcryst. ID* :soh. microcryst. 'De - : soln. microcryst. ID* :soln. microcryst.

963
References Bi62b 63c Bi62b' 63c Do629 Bi62b,'63c Dd62a D662a Da62a D362a Do62a Da62a DB62a D662a Do62a Do62a D662a Bi63b

M, : soln.

(4.3)

1%benzanthracene (BA)

7cd?
3'
4' 5
10
4

ID* : soln. microcryst.

Bi62a,b,,63d,e, St64 Fo63, B164c, Pe64


.

?'-methyl BA 3'-methyl BA 4'-methyl BA 3-methyl BA Cmethyl BA 5-methyl BA 6-methyl BA 7-methyl BA 8-methyl BA 6=CH8 7=CHs 8 = CH.

3';6-dimeth$l BA 5-ethyl BA 5-n-propyl BA 5-butyl BA 5-amYl BA 6-iso-propyl BA Cholanthrene

lD* : soh. lD* : soln. ID* : soh. lD* : soln. lD* : soln. ID* : soln. microcryst. 'D* : soln. microcryst. ID* : soln. 1D* :s o b . microcryst. lD* : soln. lD* : soln. lD* : soln. ID* :so+. ID* : microcryst. lD* : soln. lD* : soln. 'De : soln. ID* : soln. ID* : soln. ID* : soln.

64c 64c

64c

20-methylcholanthrene (4.4) Chrysene

'D* : microcryst.

Bi59, St62 Bi64b

S o h : no ID*

(4.5)

3:4-benzophenanthrene

(B

Ph)

1-methyl B Ph

w
I

I-CH,

ID* : microcryst.

Bi59

964
(4.6) Compound Triphenylene

J B Birks
Structure Conditions lD* : microcryst. References Pe64

(5.1)

Perylene

(5.2)

1 :2 benzopyrene

83
cc033

ID* : cryst.
S O h

liq.

Sa56 Sh60, St62: Ta63 Ch72 In72 Bi64b) La68b

ID* : s o h .
liq.

Bi62b, 63c, 64b Bi64a

(5.3)

3 :4-benzopyrene

ID* : soln.

cryst.

B!62b, 63c B159, St62

(5.4) Pentacene

MI : soh.

Bi63b

(5.5)

1 : 2 :3 :4-dibenzanthracene

ID* : soh. microcryst.

Bi62b, 64b Pe64

ID* :microcryst.

Pe64

(5.7)

1 :2-benzochrysene

lD* : microcryst.

Bi59

Excimers
(5.8) Compound 5 :6-benzochrysene Structure Conditions
I D * : microcyst.

965
References Bi59

(6.1)

Anthanthrene

ID* : soln.

microcryst.

bi64b No56

(6.2)

1 : 12-benzoperylene

I D * : soh.

microcryst.

Bi62b, 64b St62

(6.3)

1:2 :3 :4-dibenzopyrene

ID* : soln

Bi62b, 63c

(6.4)

1 :2 :3 :4-dibenzotetracene

ID* : microcryst.

Pe64

( 6 , s ) Tetrabenzonaphthalene

ID* : microcryst.

Pe64

(7.1)

Coronene

I D * : microcryst.

St62

966
Compound (10.1) Ovalene

J B Birks
Structure Conditions ID* : microcryst. References St62

(a)

Other aromatic compounds Fluorene

ID* : soln.

Ho70

Dibenzofuran

ID* : soln.

Ho70, Pi71, No72

Quinoline

lD* : soln.

B173

Isoquinoline

"

lD* : soln.

B173

2,5-diphenyloxazole (PPO)

'De : soln.

Be61 Bi62b, La68, 68a

2-(4-hiphenylyl), 5-phenyl OXazOk (BPO)

La66, 68,

2-(l-naphthyl), 5-phenyl oxazole (a-NPO)

lD* : soln.

La66,68a

2-(l-naphthyl), 5-phenyl 1,3,4-oxadiazole ( ~ - N P D )

ID* : soln.

La66. 68a

2- henyl 544-bi henyl)

1,5),4-ox~diazole &BD)

Excimers
Referenc A165 73a 73b Ba65a 65b 65c 66 Be61 71 Bi59 62a 62b 63a 63b 63c 63d 63e 631 64a 64b 64c 64d 64e 65 66a 66b 68a 68b 68c 68d 68e 69
Cu72b D662a 62b 62c Fe58 73 Fi67 F654 55 63 63 Gr68 Hi65 69a 69b !1069 70 In72 Is67 Iv62 Jo65 Ka57
77

967
Cundall and Robinson (1972a) Doller (1962) Doller and FBrster (1962~) Daller and Farster (1962b) Ferguson (1958) Ferguson et a1 (1973) Fielding and Jamagin (1967) Farster and Kasper 1954) Forster and Kasper {1955) Fonter et al (1963) FBrster and Seidel (1965) Greenleaf et al(1968) Gregory and Helman (1972) Hirayama (1965) Hirayama and Lipsky 1969a) Hirayama and Lipsky [1969b) Horrocks (1969) Horrocks and Brown (1970) Inoue et a1 (1972a) IshiiandMatsui (1967) Ivanova et aZ(l962) Jones and Nicol(1965) Kasper (1957 Kawada and t a b e s (1970) Kawakubo (1972) Lalande and Calas (1960) LamietaZ(1966) Lami and Laustriat (1968) J Langelaarand G J Hoytink (1968, y h t e communication) umb and Weyl (1967) Mataga et a1 (1965) Melhuish (1961)

zo
lL

La60 66 68a 68b Lu67 Ma65 Me61 Mo66 No56 72 Ob67 Pa63 Pe64 Pi71 Sa56 Se66 Sh6O Sm66 St62 63a 63b 64 67 Ta63 Wi72

Blij Cas9 65 67 Chj9 65 66a . . . 66b 66c 66d 68 70a 70b 72 Co69a 69b 71 Cr66 Cu72a
~

71

Stevens and Dickinson (1963b) Stevens and Ban (1964) Stevens et a 2 (1967) Tanaka (1963) Williams and Thomas ( 1 972)

64

968
References

J B Birks

ALADEKOMO J B 1973 J . Luminesc. 6 83-95 ALADEKOMO J B and BIRKS J B 1965 Proc. R, Soc. A 284 551-65 ALWATTAR AH, LUMB M D and BIRKS J B 1973 Organic Molecular Photophysics vol 1, ed J B Birks (London, New York: Interscience) pp403-56 APPLEQUIST D E , FRIEDRICH E C and ROGERS M T 1959 J . Am. Chem. Soc. 81 457-8 AVAKIAN P and ABRAMSON E 1965 J. Chem. Phys. 43 821-3 AVAKIAN P, GROFF R P, KELLOGG R E, MERRIFIELD R E and SUNA A 1971 Organic Scintillators and Liquid Scintillation Counting ed D L Horrocks and C-T Peng (New York, London: Academic) pp499-5 10 AVAKIAN P and MERRIFIELD R E 1968 Molec. Cryst. 5 37-77 AZUMIT, ARMSTRONG A T and MCGLYNN S P 1964J. Chem. Phys. 41 3839-52 AZUMI T and A ~ U M H I 1966a Bull. Chem. Soc.Japan 39 1829-36 -1966b Bull. Chem. Soc. Japan 39 2317-20 -1967 Bull. Chem. Soc. Japan 40 279-84 AZUMI T and MCGLYNN S P 1963a J. Chem. Phys. 38 27734 -196313J . C h m . P h p . 39 1186-94 -1964J. Chem. Phys. 41 3131-8 -1965 J . Chem. Phys. 42 1675-80 BADGER B and BROCKLEHURST B 1968 Nature, Lond. 219 263 -1969a Trans. Faraday Soc. 65 2582-7 -1969b Trans. Faraday Soc. 65 2588-94 BADGER B, BROCKLEHURST B and RUSSELL R D 1967 Chem. Phys. Lett. 1 1 2 2 4 BARNES R L and BIRKS J B 1966 Proc. R. Soc. A 291 570-82 BASILE L J 1964 Trans. Faraday Soc. 60 1702-14 BASOV N G, BALASKOV E M, BOGDANKEVITCH 0 V, DANILYCKEV V A, KASHNIKOV G N, LANTZOV N P and KHODKEVICH D D 1970J. Luminesc. 1-2 834 BAZILEVSKAYA N S and CHERKASOV A S 1965a Opt. Spectrosc. 18 30-2 -1965b Opt. Spectrosc. 18 77-9 BAZILEVSKAYA N S, LIMAREVA L A , CHERKASOV A S and SHIROKOV V I 1965 Opt. Spectrosc. 18 202-3 BEARDSLEE R A and OFFEN H W 1971 J. Chem. Phys. 55 3516-9 BEENS H and WELLER A 1968 Chem. Phys. Lett. 2 140-2 -1975 Organic Molecular Photophysics vol 2, ed J B Birks (London, New York: Interscience) ppl59-215 BENDER C F and DAVIDSON E R 1966 J . Phys. Chem. 70 2675-85 BERLMAN I B 1961 J. Chem. Phys. 34 1083-4 BIRKS J B 1951 Proc. Phys. Soc. A 64 874-7 -1963 J . Phys. Chem. 67 2199-200 -1964a Theory and Practice of Scintillation Counting (Oxford : Pergamon) -1964b Acta Phys. Polon. 26 367-78 -1964c J . Phys. Chem. 68 439 -1967 Chem. Phys. Lett. 1 304-6 -1968a Chem. Phys. Lett. 1 625-6 -1968b Acta Phys. Polon. 34 603-17 -1969a Molecular Luminescence ed E C Lim (New York: Benjamin) pp219-37 -1969b Molecular Luminescence ed E C Lim (New York: Benjamin) p907 -1970a Photophysics of Aromatic Molecules (London, New York: Interscience) -1970b Prog. Reaction Kinetics 5 181-272 -1970c Chem. Phys. Lett. 4 603-6 -1973 Organic Molecular Photophysics vol 1, ed J B Birks (London, New York: Interscience) ppl-55 -1975a Proc. Int. Conf.on Exciplexes, University of Western Ontario 1974 in the press -1975b Organic Molecular Photophysics vol 2, ed J B Birks (London, New York: Interscience) pp409-613 BIRKS J B and ALADEKOMO J B 1963 Photochem. Photobiol. 2 41 5-8 -1964 Spectrochim. Acta 20 15-21 BIRKS JB, ALWATTAR A J H and LUMB M D 1971 Chem. Phys. Lett. 11 89-92

Excimer s

969

BIRKS J B, APPLEYARD J H and POPER 1963a Photochem. Photobiol. 2 493-5 J B, BRAGA C L and LUMB M D 1964a BY. J. Appl. Phys. 15 339404 BIRKS -1965 Proc. R. Soc. A 283 83-99 BIRKS J B and CAMERON A J W 1959 Proc. R. Soc. A 249 297-317 J B and CHRISTOPHOROU L G 1962a Nature, Lond. 194 442-4 BIRKS -1962b Nature, Lond. 196 33-5 -1963a Spectrochim. Acta 19 401-10 -1963b PYOC. R. SOC. A 274 552-64 -1963c Nature, Lond. 197 1064-5 -1964 Proc. R. Soc. A 277 571-82 BIRKS J B, CHRISTOPHOROU L G and HUEBNER R H 1968a Nature, Lond. 217 809-12 J B and CONTE J C 1968 Proc. R. Soc. A 303 85-95 BIRKS J B, CONTE J C and WALKER G 1968b J. Phys. B: Atom. Molec. Phys. 1 9 3 4 4 5 BIRKS BIRKS J B and DYSON D J 1963 Proc. R. Soc. A 275 135-48 BIRKS J B, DYSON D J and KINGT A 196413 Proc. R. Soc. A 277 270-8 J B, DYSON D J and MUNRO I H 1963b Proc. R. Soc. A 275 575-88 BIRKS BIRKS J B and GEORGHIOU S 1968 J. Phys. B: Atom. Molec. Phys. 1 958-65 J B and KAZZAZ A A 1967 Chem. Phys. Lett. 1 307-8 BIRKS -1968 P ~ o cR. . SOC. A 304 291-301 BIRKS J B, KAZZAZ AA and KINGT A 1966a Proc. R. Soc. A 291 556-69 J B and KINGT A 1966 Proc. R. Soc. A 291 244-56 BIRKS BIRKS J B, LUMB M D and MUNRO I H 1964c Acta Phys. Polon. 26 379-84 -1964d Proc. R. Soc. A 280 289-97 BIRKS J B and MOORE G F 1967 The Triplet State ed A B Zahlan (London: Cambridge UP) pp407-14 BIRKS J B, MOORE G F and MUNRO I H 1966b Spectrochim. Acta 22 323-31 J B and MUNRO I H 1967 Prog. Reaction Kinetics 4 239-303 BIRKS J B and SEIFERT H G 1965 Phys. Lett. 18 127-8 BIRKS BIRKS J B, SRINIVASAN B N and MCGLYNN S P 1968c J. Molec. Spectrosc. 27 266-84 BLAUNSTEIN R P and GANTK S 1973 Photochem. Photobiol. 18 347-9 BOUCIQUB R and MORTIER P 1970J. Phys. D : Appl. Phys. 3 1905-11 G 1961 Elektronen-Donator-Acceptor-Komplexe (Heidelberg : Springer) BRIEGLEB BRIGMAN G H , BRIENT S J and MATSEN F A 1961 J. Chem. Phys. 34 958-60 BROCKLEHURST B and RUSSELL R D 1967 Nature, Lond. 213 65 -1969 Trans. Faraday Soc. 65 2159-67 R, HAENSEL R, HAHNU, NIELSEN U and ZIMMERER G 1974 Chem. Phys. Lett. BRODMANN 29 250-2 BROWNE J C 1965 J. Chem. Phys. 42 2826-9 BUCKINGHAM R A 1958 Trans. Faraday Soc. 54 453-9 A 1952 Proc. R. Soc. A 213 327-49 BUCKINGHAM R A and DALGARNO CALAS R and LALANDE R 1959 Bull. Soc. Chim. France 763-5, 770-2 J G and MOULINES F 1965 Bull. Soc. Chim. France 119-20 CALAS R, LALANDE R, GANGERE P 1960 Bull. Soc. Chim. France 148-9 CALAS R, LALANDE R and MAULET CARTER J G, CHRISTOPHOROU L G and ABU-ZEID M E M 1967 J. Chem. Phys. 47 3879-84 CASTRO G and HOCHSTRASSER R M 1966 J. Chem. Phys. 45 4352-3 CHANDRA A K and LIME C 1968 J. Chem. Phys. 48 2589-95 -1969 Molecular Luminescence ed E C Lim (New York: Benjamin) pp249-66 CHANDROSS E A 1965 J. Chem. Phys. 43 4175-6 CHANDROSS E A and DEMPSTER C J 1970a J. Am. Chem. Soc. 92 7 0 3 4 , 704-6 -1970b J. Am. Chem. Soc. 92 3586-93 J 1966a J. Chem. Phys. 45 397-8 CHANDROSS E A and FERGUSON -1966b J. Chem. Phys. 45 3554-64, 3564-7 CHANDROSS EA, FERGUSON J and MCRAE E G 1966 J. Chem. Phys. 45 3546-53 J W and VISCO R E 1965 J. Am. Chem. Soc. 87 3259-60 CHANDROSS EA, LONGWORTH CHAPMAN 0 L and LEE K 1969 J. Org. Chem. 34 4166-8 S C 1970 J. Phys. C: Solid St. Phys. 3 1791-6 CHAUDHURI M K and GANGULY CHERKASOV A S and BAZILEVSKAYA N S 1965 Bull. Acad. Sci. USSR, Phys. Ser. 29 1288-99 CHERKASOV A S and VEMBER T M 1959 Opt. Spectrosc. 6 207-10, 319-24 0, GEDANKEN A, RAZ B and JORTNER J 1973a Chem. Phys. Lett. 22 23-5 CHESHNOVSKY

970

J B Birks

Chem. Phys. 59 5554-61 C J and SIMMONS H E 1965 3.C h m . Phys. 42 1127-8 CHESNUT D B, FRITCHIE CHRISTOPHOROU L G, ABU-ZEID M E M and CARTER J G 1968 J. Chem. Phys. 49 3775-82 L G and CARTER J G 1966 Nature, Lond. 209 678-80 CHRISTOPHOROU CHUN Y C, KAWAOKA K and KEARNS D R 1971 J. Chem. Phys. 55 3059-67 CHUN Y C and KEARNS D R 1972 Molec. Cryst. Liquid Cryst. 16 61-74 COHEN M D 1969 iMolec. Cryst. Liquid Cryst. 9 287-95 M D, LUDMER Z, THOMAS J M and WILLIAMS J 0 1969 Chem. Commun. 1172-3 COHEN -1971 Proc. R. Soc. A 324 459-68 COLLI L 1954 Phys. Rev. 95 892-4 E L 1961 J . Molec. Spectrosc. 7 223-30 COOPER C D , COBB G E and TOLNAS COOPER C D and LICHTENSTEIN M D 1958 Phys. Rev. 109 2026-8 J K 1968 J. Chem. Phys. 48 5097-102 COOPER R and THOMAS CRAIG D P and SARTI-FANTONI P 1966 Chem. Commun. 742-3 CUNDALL R B and OGILVIE SMCD 1975 Organic Molecular Photophysics vol 2, ed J B Birks (London, New York: Interscience) pp33-93 CUNDALL R B and PEREIRA L C 1972a Chem. Phys. Lett. 15 383-6 -1972b J . Chem. Soc. Faraday Trans. II68 1152-63 R B and ROBINSON D A 1972aJ. Chem. Soc. Faraday Trans. 1168 113344,1145-51 CUNDALL -197213 Chem. Phys. Lett. 13 257-9 CUNDALL R B and VOSSA J R 1969 Chem. Commun. 116 CZARNECKI S 1961 Bull. Acad. Polon. Sci., Se. Math. Astr. Phys. 9 561-3 E 1962 2. Phys. Chem. (NF) 34 151-62 DBLLER E and F~RSTER TH1962a 2. Phys. Chem. ( N F )31 274-7 DBLLER -1962b 2. Phys. Chem. ( N F ) 34 132-50 DRUMMOND D L and GALLAGHER A 1974 J. Chem. Phys. 69 3426-35 D D and NAKANO H H 1973 Chem. Phys. Lett. 23 112-4 ECKSTROM D J, HILLR M , LORENTS M, SHULMAN R G and YAMANE T 1966 Proc. Natn. Acad. Sci. U S A 55 EISINGER J, GUERON 1015-20 FAULKNER L and BARD A J 1968J . Am. Chem. Soc. 90 6284-90 -1969 J. Am. Chem. Soc. 91 6495-7 FERGUSON J 1958 J . Chem. Phys. 28 765-8 FERGUSON J and MAUA W H 1974 Molec. Phys. 27 377-87 FERGUSON J, MAW A W H and MORRIS J M 1973 Aust. J. Chem. 26 91-102 P C 1967 J. Chem. Phys. 47 247-52 FIELDING P E and JARNAGIN FINK E H and COMES F J 1975 Chem. Phys. Lett. 30 267-72 FINKELNBURG W 1938 Kontinuierliche Spektren (Berlin: Springer) FISCHER D, NAUNDORF G and KL~PFFER W 1973 2. Nuturf, 28a 973-9 FBRSTER TH1962 Pure Appl. Chem. 4 121-34 1963 Pure Appl. Chem. 7 73-8 -1969 Angew. Chem. (fnt.)8 333-43 FBRSTER THand KASPER K 1954 2. Phys. Chem. ( N F ) 1 275-7 -1955 2. Elektrochem. 59 976-80 FBRSTER TH,LEIBER CO, SEIDEL H P and WELLER A 1963 2.Phys. Chem. ( N F ) 39 265-9 F~RSTER THand SEIDEL H P 1965 Z. Phys. Chem. ( N F ) 45 58-71 T R , COZZENS R F and MCDONALD J R 1972J. Chem. Phys. 57 534-41 Fox RB, PRICE FREEMAN N G , MCEWAN H J , CLARIDGE R F C and PHILIPS L F 1971 Chem. Phys. Lett. 10 530-2 FROINES J R and HAGERMAN P J 1969 Chem. Phys. Lett. 4 135-8 GEDANKEN A,JORTNER J, RAz B and Szbm A 1972 J. Chem. Phys. 57 3456 A,RAz B and JORTNER J 1973aJ. Chem. Phys. 59 1630-3 GEDANKEN -197313 J. Chem. Phys. 59 5471-83 GEORGE G A and MORRIS G C 1970 MoZec. Cryst. Liquid Cryst. 11 61-83 0L J 1972 Doctoral Thesis University of Amsterdam GIJZEMAN GIJZEMAN 0 L J, LANGELAAR J and VAN VOORST J D W 1970 Chem. Phys. Lett. 5 269-72 GIJZEMAN 0L J, VAN LEEUWEN WH, LANGELAAR J and VAN VOORST J D W 1971 Chem. Phys. Lett. 11 528-31,532-4
I _

-1972b J . Chem. Phys. 57 4628-32


-1973bJ.

CHESHNOVSKY 0, RAz B and JORTNER J 1972a Chem. Phys. Lett. 15 475-9

Excimers

97 1

GINTER M L 1970a International Tables of Selected Constants, 17. Spectroscopic Data Relative to Diatomic Molecules ed B Rosen (Oxford: Pergamon) pp193-201 -1970b International Tables of Selected Constants, 17. Spectroscopic Data Relative to Diatomic Molecules ed B Rosen (Oxford: Pergamon) pp440-2 GOLDE M F and THRUSH BA 1974 Chem. Phys. Lett. 29 486-90 GOLDSCHMIDT C R and OTTOLENGHI M 1970 J. Phys. Chem. 74 2041-2 GOLDSCHMIDT C R, TOMKIEWICZ Y and BERLMAN I B 1968 Chem. Phys. Lett. 2 520-2, 536-8 GREENE F D 1960 Bull. Soc. Chim. France 1356-9 GREENLEAF J R, LUMB M D and BIRKS J B 1968 J. Phys. B: Atom. Molec. Phys. 1 11 57-9 GREGORY T A and HELMAN W P 1972 J. Chem. Phys. 56 377-85 GUBERMAN S L and GODDARD W A 1972 Chem. Phys. Lett. 14 460-5 GUPTA B K and MATSEN F A 1969 J. Chem. Phys. 50 3797-803 HALPERN A M 1974 J. Am. Chem. Soc. 96 4392-8 HALPERN A M and MARATOS E 1972 J . Am. Chem. Soc. 94 8273-4 HAUSER M and HEIDTG 1970 2. Phys. Chem. ( N F ) 69 201-15 HEDGES R, DRUMMOND D and GALLAGHER A 1972 Phys. Rev. A 6 1519 HEINZELMANN W and LABHART H 1969 Chem. Phys. Lett. 4 20-4 HELLER R 1941 J, Chem. Phys. 9 154-63 HERZBERG G 1950 Molecular Spectra and Molecular Structure, 1. Spectra of Diatomic Molecules (Princeton, NJ : Van Nostrand) HILLIER I H , GLASS L and RICES A 1966 J. Chena. Phys. 45 3015-21 HIRAYAMA F 1965 J. Chem. Phys. 42 3163-71 HIRAYAMA F and LIPSKY S 1969aJ. Chem. Phys. 51 1939-51 -1969b Molecular Luminescence ed E C Lim (New York: Benjamin) pp237-47 HOCHSTRASSER R M 1962 J . Chem. Phys. 36 1099-100 HOCHSTRASSER R M and MALLIARIS A J 1965 J. Chem. Phys. 42 2243-4 HOLZMAN P and JARNAGIN R C 1969 J. Chem. Phys. 51 2251-3 HOPFIELD J J 1930a Phys. Rev. 35 1133-4 -1930b Astrophys. J. 72 133-45 HORROCKS A R 1970 Can.J. Chem. 48 1000-2 HORROCKS D L 1969 J. Chem. Phys. 50 4962-6 HORROCKS D L and BROWN W G 1970 Chem. Phys. Lett. 5 117-9 HOWARTH 0 W and FRAENKEL G K 1966 J. Am. Chem. Soc. 88 4514-5 -1970 J. Chem. Phys. 52 6258-67 HOYTINK G J 1968 Discuss. Faraday Soc. 45 14-22 HUFFMAN R E , LARRABEE J C and TANAKA Y 1965 Appl. Opt. 4 1581 HURST G S, BORTNER T E and STRICKLER T D 1969 Phys. Rev. 178 4-10 INOUE A, YOSHIHARA K, KASUYA T and NAGAKURA S 1972a Bull. Chem. Soc. Japan 45 720-5 INOUE A, YOSHIHARA K and NAGAKURA S 1972b Bull. Chem. Soc. Japan 45 1973-6 ISHIIY and MATSUI A 1967 J . Phys. Soc. Japan 22 926 IVANOVA T V , MOKEEVA C A and SVE~HNIKOV BYa 1962 Opt. Spectrosc. 12 325-8 JONELEIT D 1969 2. Naturf. 24a 1809-20 JONES P F and NICOLM 1965 J . Chem. Phys. 43 3759-60 JORTNER J, MEYER L, RICES A and WILSON E G 1965 J . C h m . Phys. 42 4250 KASHA M 1950 Discuss. Faraday Soc. 9 14-9 -1952J. Chem. Phys. 20 71-4 KASPER K 1957 2. Phys. Chem. (NF) 12 52-67 KAWADA A and LABES M M 1970 Molec. Cryst. Liquid Cryst. 11 133-44 KAWAKUBO T 1972 Molec. Cryst. Liquid Cryst. 16 333-53 KAWAOKA K and KEARNS D R 1964 J. Chem. Phys. 41 2095-7 -1966 J. Chem. Phys. 45 147-53 KENTY C 1967J. Chem. Phys. 47 2545-51 KENTY C, AICHER J 0, NOELEB, PORITZKY A and PAOLINO V 1946 Phys. Rev. 69 36-7 KLOPFFER W 1973 Organic Molecular Photophysics vol 1, ed J B Birks (London, New York: Interscience) pp357-402 KLBPFFER W and BAUSER H 1970 Chem. Phys. Lett. 6 279-81 KLOPFFER W, BAUSER H , DOLEZALEK F and NAUNDORF D 1972 Molec. Cryst. Liquid Cryst. 16 229-45 KOEHLER H A , FERDERBER L J , REDHEAD D L and EBERT P J 1972 Appl. Phys. Lett. 21 198-200

972

J B Birks

KONDRATJEW V and LEIPUNSKY A 1929 Trans. Faraday Soc. 25 736-7 KONIJNENBERG E 1963 Doctoral Thesis University of Amsterdam KOYANAGI M 1968 J . Molec. Spectrosc. 25 273-90 KUBOTA S, TAKAHASHI T and DOKE T 1968 Phys. Rev. 165 225-30 KUHN H 1937 Proc. R. Soc. A 158 212-29,23041 KUHN H and OLDENBERG 0 1932 Phys. Rev. 41 72 LABHART H and WYRSCH D 1971 Chem. Phys. Lett. 12 378-81 LALANDE R and CALAS R 1959 Bull. Soc. Chim. France 766-70 1960 Bull. Soc. Chim. France 144-7 LAMI H, GRESSET J and LAUSTRIAT G 1966 Proc. Int. Symp. on Luminescence (Munich: Karl Thiemig) pp190-201 LAMI H and LAUSTRIAT G 1968 J . Chem. Phys. 48 1832-40 LANGELAAR J 1969 Doctoral Thesis University of Amsterdam LANGELAAR J, JANSEN G, RETTSCHNICK R P H and HOYTINK G J 1971a Chem. Phys. Lett. 12 86-91 LANGELAAR J, RETTSCHNICK R P H and HOYTINK G J 1971b J . Chem. Phys. 54 1-7 LANGELAAR J, RETTSCHNICK R P H, LAMBOOY A M P and HOYTINK G J 1968 Chem. Phys. Lett. 1 609-1 2 LAWSON C W, HIRAYAMA F and LIPSKY S 1969J. Chem. Phys. 51 1590-6 LEONHARDT H and WELLER A 1961 2. Phys. Chem. ( N F ) 29 277-80 -1963 Bey. Bunsenges. Phys. Chem. 67 791 LEWIS I C and STRINGER L S 1965 J . Chem. Phys. 43 2712-27 LIME C and CHAKRABARTI S K 1967 Molec. Phys. 13 293-6 LUMB M D and WEYL D A 1967 J . Molec. Spectrosc. 23 365-71 MCCLURE D S 1949 J . Chem. Phys. 17 905-13 MASUHARA H and MATAGA N 1970 Chem. Phys. Lett. 7 417-9 MATAGA N, TOMURA M and NISHIMURA H 1965 Molec. Phys. 9 367-75 MEDINGER T and WILKINSON F 1966 Trans. Faraday Soc. 62 1785-92 MELHUISH W H 1961 J . Phys. Chem. 65 229-35 MICHELS H H and HARRIS F E 1963 J. Chem. Phys. 39 1464-9 MOORE G F 1966 Nature, Lond. 212 1452-3 MOORE G F and MUNRO I H 1967 Spectrochim. Acta 23A 1291-8 Moss F E and HEREFORD F L 1963 Phys. Rev. Lett. 11 63 MUEL B 1962 C.R. Acad. Sci., Paris 255 3149-51 MULLIKEN R S 1932 Rev. Mod. Phys. 4 1-86 -1950 J . Am. Chem. Soc. 72 600-8 -1952aJ. Am. Chem. Soc. 74 811-24 -1952b J . Phys. Chem. 56 801-22 -1963 J . Chim. Phys. 61 20-38 -1964a Phys. Rev. 136 962-5 -1964bJ. Am. Chem. Soc. 86 3183-91 -1966J. Am. Chem. Soc. 88 1849-61 -1969 J. Am. Chem. Soc. 91 4615-21 MULLIKEN R S and PERSON W B 1962 Ann. Rev. Phys. Chem. 13 107-26 -1969 Molecular Complexes (New York: Interscience) J N and POPLE J A 1956 Proc. Phys. Soc. A 69 245-52 MURRELL MURRELL J N and TANAKA J 1964 Molec. Phys. 7 363-80 NAKASHIMA T T and OFFEN H W 1968 J . Chem. Phys. 48 4817-21 NEZNAIKO N F, OBYKNOVENNAYA I E and CHERKASOV A S 1966 Opt. Spectrosc. 21 23-5, 285-6 NICHOLS L L and VALIW 1968 J . Chem. Phys. 49 814-7 NORTHROP D C and SIMPSON 0 1956 Proc. Roy. Soc. A 234 136-49 NORTHROP J A and GURSKY J C 1958 Nucl. Instrum. 3 207 NORTHROP J A and NOBLES R 1956a Nucleonics 14 (No. 4) 36 -1956b I R E Trans. Nucl. Sci. NS-3 (No. 4) 59 NOTTP R and SELINGER B K 1972 J . Luminesc. 5 138-42 NOYES R M 1961 Prog. Reaction Kinetics 1 131-60 OBYKOVENNAYA I E and CHERKASOV A S 1967 Opt. Spectrosc. 22 172-3 OKAT , %MA RAOK V S , REDPATH J L and FIRESTONE R F 1974 J. Chem. Phys. 61 4740-6 OLDENBERG 0 1929 2. Phys. 55 1-15

Excims

973

PACKARD RE, RIEFF and SURKO C M 1970 Phys. Rev. Lett. 25 1435-9 PARKER C A 1963a Nature, Lond. 200 231-2 -196313 Spectrochim. Acta 19 989-94 -1964a Ads. Photochem. 2 305-83 -1964b Trans. Faraday Soc. 60 1998-2008 -1967 The Triplet State ed A B Zahlan (London: Cambridge UP) pp353-90 PARKER C A and HATCHARD C G 1962a Proc. Chem. Soc. 147 -196213 Proc. R. Soc. A 269 574-84 -1963 Trans. Faraday Soc. 59 284-95 PARKER C A and JOYCE T A 1967 Chem. Commun. 744-5 PARKER C A and SHORT G D 1967 Trans. Faraday Soc. 63 2618-22 PERKAMPUS H H and POHL L 1964 2.Phys. Chem. ( N F ) 40 162-88 PETER L and VAUBEL G 1973a Chem. Phys. Lett. 18 531-4 197313 Chem. Phys. Lett. 21 158-60 PHILLIPS D H and SCHUG J C 1969 J. Chem. Phys. 50 3297-306 PHILLIPS L F 1973 Chem. Phys. Lett. 21 28-9 PINION J P, MINNF L and FILIPESCU N 1971 J. Luminesc. 3 245-52 PLATT J R 1949 J. Chem. Phys. 17 484-95 POLAK R and PALDUS J 1966 Theor. Chim. Acta 4 37-43 POSHUSTA R D and MATSEN F A 1963 Phys. Rev. 132 307-9 PRINCE J F and ROBERTSON W W 1966 J. Chem. Phys. 45 2577 -1967 J. Chem. Phys. 46 3309 RAYLEIGH LORD 1927 Proc. R. Soc. A 116 702-19 RICHARDS J T and THOMAS J K 1970 Chem. Phys. Lett. 5 527-8 ROMPE R 1936a 2.Phys. 101 214-33 -1936b Phys. 2.37 807-8 ROSEN B (ed) 1970 International Tables of Selected Constants, 17. Spectroscopic Data Relative to Diatomic Molecules (Oxford: Pergamon) SANGSTER R C and IRVINE J W 1956 J. Chem. Phys. 24 670-714 SCHMIDT G M J 1967 Reactivity of the Photoexcited Organic Molecule (London: Interscience) pp227-84 SCHNEIDERMAN S B and MICHELS H H 1965 J. Chem. Phys. 42 3706-19 SCOTT D R , GREENAWALT E M , BROWNE J C and MATSEN F A 1966 J. Chem. Phys. 44 2981-4 SELINGER B K 1964 Nature, Lond. 203 1062-3 1966 Aust. J. Chem. 19 825-34 SHPOL'SKII E V and PERSONOV R I 1960 Opt. Spectrosc. 8 172-6 SLIFKIN M A 1963 Nature, Lond. 200 766-7 SLIFKIN M A and AL-CHALABI A 0 1973 Chem. Phys. Lett. 20 211 SMITH F J, ARMSTRONG A T and MCGLY" S P 1966 J. Chem. Phys. 44 442-8 SMITH-SAVILLE R J 1968 PhD Thesis University of Manchester STERNLICHT H, NIEMAN G C and ROBINSON G W 1963J. Chem. Phys. 38 1326-35 STEVENS B 1962 Spectrochim. Acta 18 439-48 -1969 Chem. Phys. Lett. 3 233-6 -1971 Adv. Photochem. 8 161-226 STEVENS B and BANM I 1964 Trans. Faraday Soc. 60 1515-23 -1968 Molec. Cryst. 4 173-81 STEVENS B and DICKINSON T 1963a J. Chem. Soc. 5492-6 -1963b Spectrochim. Acta 19 1865-70 STEVENS B and HUTTON E 1960 Nature, Lond. 186 1045-6 STEVENS B and MCCARTIN P J 1960 Molec. Phys. 5 425-33 STEVENS B, THOMAZ M F and JONES J 1967J. Chem. Phys. 46 405-6 STEVENS B and WALKER M S 1963 Proc. Chem. Soc. 181 -1964 Proc. R. Soc. A 281 420 STEWART T E, HURST G S, BORTNER T E, PARKS J E, MARTIN F W and WEIDNER H L 1970 J. Opt. Soc. Am. 60 1290-7 STRICKLER S J and BERG R A 1962 J. Chem. Phys. 37 814-22 STRICKLER T D and ARAKAWA E T 1964J. Chem. Phys. 41 1783-9 L 1972 Phys. Lett. 39A 349-50 STUPAVSBY M, DRAKE G W F and KRAUSE SVESHNIKOV B 1935 Acta Physicochim. URSS 3 257

974

J B Birks

SVESHNIKOV B 1937 Acta Playsicochim. U R S S 7 758 SWENBERG C E and GEACINTOV N E 1973 Organic Molecular Photophysics vol 1, ed J B Birks (London, New York : Interscience) pp489-564 TAKAHASHI T, KUBOTA S and DOKE T 1966 Phys. Lett. 23 321-2 TANAKA C, TANAKA J, HUTTON E and STEVENS B 1963 Nature, Lond. 198 1192 TANAKA J 1963 Bull. Chem. Soc. Japan 36 1237-49 TANAKA Y and YOSHINO K 1972 J. Chem. Phys. 57 2964 THONNARD N and HURST G S 1972 Phys. Rev. A 5 1110-21 TISONE G C 1974J. Chem. Phys. 60 3716-7 TOMKIEWICZ Y and LOEWENTHAL E 1969 Molec. Cryst. Liquid Cryst. 6 211-25 TOMLINSON W J, CHANDROSS E A , FORK R L , PRYDE C A and LAMOLA A A 1972 Appl. Opt. 11533-48 TOMURA M and TAKAHASHI Y 1971 J. Phys. Soc. Japan 31 797-801 VALA M T , HAEBIG J and RICES A 1965 J. Chenz. Phys. 43 886-97 M T, HILLIER I H, RICES A and JORTNER J 1966 J. Chem. Phys. 44 23-35 VALA 0 E, CHRISTOPHOROU L G and CARTER J G 1969 Chenz. Phys. Lett. 4 224-5 WAGNER J B 1973 J. Chenz. Phys. 59 1738-41 WAYNE JOHNSOX A and GERARDO WEIHOFEN W H 1974 J. Chenz. Phys. 60 445-53 L 1948 C.R. Acad. Sci., Paris 227 1022-4 WENIGER C and HERMAN T C, CHANG J and HERCULES D M 1970 J. Am. Chem. Soc. 92 763-8 WERNER L L 1970 J. Phys. Chem. 74 2404-6 WESTM L and NICHOLS WILKINSON P G 1965 J. Quant. Spectrosc. Radiat. Transfer 5 503 WILKINSON P G and BYRAM E T 1965 Appl. Opt. 4 581 WILKINSON P G and TANAKA Y 1955 J. Opt. Soc. Am. 60 1290-7 F E and HEBB M H 1951 Phys. Rev. 84 1181-3 WILLIAMS WILLIAMS J 0 and THOMAS J NI 1972 Molec. Cryst. Liquid Cvyst. 16 371-5 WINANS J G and STUECKELBERG E C G 1928 Proc. Natn. Acad. Sci. U S A 14 867-71 D and LABHART H 1971a Chenz. Phys. Lett. 8 217-9 WYRSCH -1971b Chem. Phys. Lett. 12 373-7 S, BOVEY F A and LUMRY R 1963 Nature, Lond. 200 242-4 YANARI

Das könnte Ihnen auch gefallen