Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Organic Scintillators and Scintillation Counting
Organic Scintillators and Scintillation Counting
Organic Scintillators and Scintillation Counting
Ebook1,530 pages

Organic Scintillators and Scintillation Counting

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Organic Scintillation and Liquid Scintillation Counting covers the proceeding of The International Conference on Organic Scintillators and Liquid Scintillation Counting, which was held on July 7-10, 1970 at the University of California, San Francisco. This conference was held to discuss ideas concerned with the theory and physics of organic scintillators and the use of liquid scintillation for radioactivity measurement and other analytical applications. This text discusses liquid scintillator solvents, the vacuum ultraviolet excited luminescence of organic systems, and the application of scintillation counters to the assay of bioluminescence. Also covered are topics such as scintillation decay and absolute efficiencies in organic liquid scintillators, dose rate saturation in plastic scintillators, and the mass measurements in a liquid scintillation spectrometer. The book is recommended for physicists who would like to know more about the advancements in the field of organic and liquid scintillation and its applications.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780323141765
Organic Scintillators and Scintillation Counting

Related to Organic Scintillators and Scintillation Counting

Physics For You

View More

Reviews for Organic Scintillators and Scintillation Counting

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Organic Scintillators and Scintillation Counting - Donald Horrocks

    PAPERS

    LIQUID SCINTILLATOR SOLVENTS*

    J.B. Birks**,      Chemistry Division, Argonne National Laboratory, Argonne, Illinois 60439

    Abstract

    The scintillation efficiency of a liquid scintillator is determined by the molecular properties of its constituents. The energy expended in the solvent by an ionizing particle initiates a complex sequence of processes including ion formation and recombination, excitation of higher excited states, internal conversion, excimer formation and dissociation, and energy migration, leading to energy transfer to the fluorescent solute molecules. This paper discusses current knowledge of these solvent processes and of their influence on the efficiency of a liquid scintillator. The origin of the delayed scintillation component is also considered.

    Unitary Scintillators

    The simplest organic scintillator consists of a pure fluorescent aromatic compound X, which may be a crystal, a liquid or a polymer. The impingement on the scintillator of an ionizing particle of energy E(eV), range r, and specific energy loss (dE/dr), generates Nx excited molecules (X*) in their first excited π-electronic singlet state S1x. We shall express the yields of all products of the ionizing particle in terms of their G-values (number per 100 eV energy expended), in accordance with radiation physics and chemistry practice, so that the S1x excitation yield is

    (1)

    The S1x-excited molecules have a fluorescence quantum yield of ϕx, and they emit Lx(=ϕxNx) fluorescence photons of mean energy hvx, so that the scintillation yield

    (2)

    Pure aromatic crystals, such as anthracene, trans-stilbene and p-terphenyl, which have high values of ϕx∼1, are commonly used as scintillator standards. Pure aromatic liquid and polymers, on the other hand, have relatively low values of ϕx and they are rarely used as scintillators.

    The scintillation yield G(hvx) depends on the specific energy loss (dE/dr) of the ionizing particle initiating the scintillation.

    (3)

    where G0(hvX) is the scintillation yield for a minimum ionizing particle (e.g. a fast electron), and Ax is the ionization quenching coefficient [1]. ϕx is independent of (dE/dr), and we deduce from (2) and (3) that

    (4)

    where G0(S1x) is the S1x excitation yield for a minimum ionizing particle. In a crystal Ax depends on the direction of the ionizing particle relative to the crystallographic axes. This effect, which is due to anisotropic exciton migration, results in an anisotropic scintillation response to α-particles, protons and other strongly ionizing particles [2, 3]. In a liquid or polymer is independent of particle direction and the scintillation response is isotropic. The parameter Ax is of similar magnitude for aromatic crystals, liquid and polymers, and it corresponds to an α/β ratio of about 0.1 [3].

    Binary Solution Scintillators

    A binary organic solution scintillator consists of an aromatic solvent X containing a mole fraction Cy of an organic solute Y of high fluorescence quantum efficiency qy. There are three main types of binary solution scintillator:

    (i) crystal solutions, in which X is an aromatic molecular crystal, such as naphthalene or anthracene;

    (ii) liquid solutions, in which X is an aromatic liquid, such as benzene or toluene; and

    (iii) plastic solutions, in which X is an aromatic polymer, such as polystyrene or polyvinyl toluene.

    The scintillation process in these three types of solution scintillator has several common features, and we may derive a general expression for the solute scintillation yield [3].

    If the solute Y is optically excited into its first excited π-electronic singlet state S1y, the quantum yield ϕy of the solute fluorescence is

    (5)

    where Ky is the Stern Volmer coefficient of concentration quenching which is due to excimer formation [4]. If the solvent X is optically excited into S1x, the quantum yield ϕxy of the solute fluorescence, excited by radiationless energy transfer from S1x to S1y, is

    (6)

    where ϕxy is the Stern-Volmer coefficient of radiationless solvent-solute energy transfer. If the solution is excited by an ionizing particle with an S1x excitation yield of G(S1x), leading via radiationless solvent-solute energy transfer to the emission of fluorescence photons of Y of mean energy hvy, the solute scintillation yield is

    (7)

    From equations (4) – (7) we obtain the general binary solution scintillator relation

    (8)

    Partial differentiation of (8) with respect to cy shows that G(hvy) is a maximum at an optimum mole fraction of

    (9)

    In a thick scintillator solution radiative solvent-solute energy transfer [3, 4], due to absorption of the solvent fluorescence by the solute, introduces an additional term into (6) and (8). In thick liquid and plastic Solutions, where the solvent fluorescence quantum efficiency qx is small, the radiative transfer component is only significant at low values of cy. It is negligible at the higher values of cy (∼(cy)m) used in practical scintillator solutions, where the radiationless transfer component is dominant, and (8) is applicable. In a thick crystal solution of high qx radiative solvent-solvent migration, radiative solvent-solute transfer, and solvent exciton traps all influence the behavior. These effects have been discussed elsewhere [5].

    The prompt scintillation process in a binary organic scintillator consists of the sequence of events, described by the three consecutive terms in (8). We shall proceed to discuss these in terms of more fundamental processes.

    Solvent Ionization and Excitation

    The ionizing particle produces ionization and excitation of solvent molecules. At cy≤(cy)m direct ionization and/or excitation of solute molecules can be neglected. The solvent ionization generates positive molecular ions (X+) and electrons (e-). Recombination of X+ and e- yields solvent molecules in an excited π-electronic singlet state SI, an excited π-electronic triplet state TI, or an excited π-electronic state σI. SI undergoes internal conversion to the lowest excited π-electronic singlet state S1x with a quantum efficiency of C(SI). TI undergoes internal conversion to the lowest excited π-electronic triplet state T1x with a quantum efficiency of C(TI). σI dissipates its energy radiationlessly as heat and/or radiation products.

    The excitation by the primary ionizing particle and by secondary electrons (δ- rays) yields solvent molecules in an excited π-electronic singlet state SE, an excited π-electronic triplet state TE, or an excited σ-electronic state σE. SE undergoes internal conversion to S1x, with a quantum efficiency of C(SE). TE undergoes internal conversion to T1x with a quantum efficiency of C(TE). σE dissipates its energy radiationlessly as heat and/or radiation products.

    The total S1x excitation yield is

    (10)

    and the total T1x excitation yield is

    (11)

    The prompt solute scintillation yield G(hvy) is proportional to G(S1x). In a unitary scintillator bimolecular interaction between pairs of T1x-excited molecules yields an S1x-excited molecule and an unexcited molecule (S0x),

    (12)

    and the delayed S1x fluorescence corresponds to the delayed scintillation component, which decays non-exponentially over a few µs [3, 6]. The origin of the delayed scintillation component in a binary liquid solution scintillator will be discussed later.

    Cooper and Thomas [7] have determined the following G-values for liquid benzene excited by high-energy electrons, using nanosecond flash photolysis:

    The total S1x yield G0(S1x) = (G0I(S1x) + G0E(S1x)) = 1.6 agrees satisfactorily with the value of G0(S1x) = 1.55 ± (±0.05) obtained by Lipsky and co-workers [8] from measurements of the absolute scintillation yield of solutions of p-terphenyl in benzene excited by β-particles.

    The spectrum of molecular states SE excited by a high-energy electron is similar to the optical absorption spectrum. The dominant π-electronic state excited in benzene is the third excited singlet state S3x(=¹E+1u), which has a fully allowed optical transition from the ground state S0x(=lAlg−). The most probable energy of SE is therefore equated to that of the Franck-Condon maximum in the optical absorption spectrum (= 54, 300 cm−1 in benzene). Lipsky and co-workers [9] have determined the quantum efficiency C(SE) of S3x-S1x internal conversion in liquid benzene toluene and p-xylene, from observations of the fluorescence excitation spectra and their values are listed in Table 1.

    TABLE 1

    Properties of Liquid Scintillator Solvents Excited by Fast Electrons

    (a) 3gm/litre PPC [10] (b) 5 gm/litre PPO [11]

    For benzene and toluene C(SE)<1; for p-xylene C(SE) = 1, and the latter value is also assumed for the higher alkyl benzenes (m-xylene, 1, 2, 4 - trimethylbenzene, iso-durene) and for 1-methyl naphthalene.

    For benzene C(SE) = 0.45, G0E(S1X) = 0.4, and hence G0(SE) = 0.9. For the other compounds G0(SE) is assumed to be inversely proportional to SE obtained from the Franck-Condon maximum in the optical absorption spectrum [4]. The values of SE, G0(SE), C(SE) and G0E(S1x) thus obtained are listed in Table 1.

    The excited singlet state SI produced by ion recombination is of lower energy than SE, and it will be assumed that it internally converts to S1x, with unit quantum efficiency in all the liquids, so that C(SI) = 1.0 and G0(SI)= G0I(S1x) = 1.2 for benzene. For the other compounds G0I (S1x) is assumed to be inversely proportional to Ix, the ionization potential [4]. The values of Ix, G0I(S1x) and G0(S1x) (= G0I(S1x) + G0E(S1x)) are listed in Table 1.

    For comparison with experiment the predicted S1x yields are expressed relative to toluene = 100, in terms of the parameter

    (13)

    The final two columns of Table 1 list the experimental values of the relative scintillation yield

    (14)

    for solutions containing (a) 3 gm/litre of PP0[10] and (b) 5 gm/litre of PP0 [11]. These PP0 concentrations approximate to (cy)m, so that the product of the second and third terms in (8) approximate to qy, which is practically independent of the solvent. It is therefore to be expected that g0(hvy) will approximate to g0(S1y). The solution scintillation yields increase in the predicted order, and the quantitative values agree satisfactorily. Somewhat higher scintillation yields are predicted for solutions based on iso-durene and on 1-methylnaphthalene.

    The differences in the scintillation yield of solutions based on benzene, toluene and xylene arise mainly from the differences in the S3x-S1x internal conversion quantum efficiency C(SE). The scintillation yields of plastic solutions in polyvinyl benzene (=polystyrene), polyvinyl toluene and polyvinylxylene increase in the same order [3], showing that C(SE) is a property of the aromatic molecular species. In benzene the S3x-S1x internal conversion and vibrational relaxation are in competition with an efficient internal quenching process to S0x, via a vibrationally distorted isomer. This process has been discussed in detail elsewhere [4]. In the vapor phase C(SE) = 0 for benzene; in solution C(SE) depends on the solvent, attaining its maximum value of C(SE) = 0.45 in pure benzene [9]. In the liquid phase C(SE) increases towards unity with alkyl substitution.

    Optical transitions from S0x to excited triplet states are spin-forbidden, so that the observed value [7] of G0(TE) = 0 is to be expected. If the four possible excited states, one singlet and three triplets, formed by ion recombination are populated with equal probability (multiplicity weighting) it is to be expected that

    Comparison with the observed value of G0I(T1x) = 1.7 yields a value of C(TI) = 0.47 for the quantum efficiency of TI - T1x internal conversion in benzene, similar to that of C(SE) = 0.45.

    Solvent Excitation Migration and Solvent-Solute Energy Transfer

    The exciton migration process in anthracene crystals is well established theoretically and experimentally [4, 5, 12]. The exciton migration (hopping) frequency kxx is related to the intermolecular interaction energy ΔW by the uncertainty principle relation

    (15)

    where h is Planck’s constant. In crystal anthracene exciton interaction energies of ΔWi = 220 cm−1 and ΔWe = 330 cm−1 are observed spectroscopically for interactions between translationally inequivalent molecules (i) and between trans lationally equivalent molecules (e), respectively [1. The transfer occurs collisionally from an excited anthracene molecule to an adjacent tetracene molecule.

    In an aromatic liquid S1x excitation migration, assisted by molecular diffusion, occurs with a migration frequency of kxx∼4 × 10¹¹s−1 [4, 15]. In an aromatic polymer similar S1x excitation migration occurs between the aromatic segmers, but the process is less rapid than in a crystal or liquid because of the wider and more random spacing of the aromatic units and the absence of diffusion. Excimer sites, at which two adjacent segmers are suitably spaced to form an excimer, act as excitation traps, and inhibit the migration [4].

    In each solution system the solvent excitation migration (rate kxx) culminates in solvent fluorescence (rate kFx), solvent internal quenching (rate kIx) or energy transfer to a solute molecule (rate kxycy). Hence the quantum yield ϕxy of solute fluorescence, excited by S1x-S1y radiationless energy transfer is

    (16)

    where

    (17)

    and kx(= 1/tx) is the reciprocal of the fluorescence lifetime of X in the absence of Y.

    The S1x-S1y energy transfer coefficient σxy is determined experimentally by optical excitation of the solvent into S1x, observation of ϕxy as a function of cy, and analysis in terms of (k3xycy, the rate of solvent-solute energy transfer from S3x. Horrocks [11] has obtained experimental evidence for the latter process in binary liquid solutions containing a very high solute mole function (∼ 10²(cy)m), such that k3xycy> kIc.

    With α-particle excitation of a binary liquid solution the behavior is consistent with (8) but the transfer coefficient σ1xy is less than σxy, obtained with β-particle or optical excitation. This nas been attributed to dynamic quenching of the S1x excitation by radiation products (e.g. free radicals) from the α-particle [20]. The dynamic ionization quenching (rate kQx = Bx(dE/dr)) is additional to the static ionization quenching described by the parameter Ax. The inclusion of kQx in the denominator of (16) reduces σxy to

    (17a)

    Two alternative models of the excitation migration and transfer processes in organic liquid solution scintillators have been proposed by Voltz et al. [21] and by Birks and Conte [15], respectively. The models differ both in the proposed solvent-solute transfer process and in the proposed solvent excitation migration process, and we shall discuss these two processes separately. Birks and Conte [15] propose that the solvent-solute transfer occurs collisionally between adjacent molecules with p = 1, as in anthracene/tetracene crystal solutions. Table 2 lists the experimental values of the rate parameters (in M−1s−1) of solvent-solute energy transfer (k′xy) from benzene, toluene, p-xylene and mesitylene to various fluorescent solutes. Also listed are the rate parameters (in M−1s−1) of the impurity quenching of the solvent fluorescence (k′Xq) by three solutes, oxygen, biacetyl and carbon tetrabromide, each of which quench collisionally with p = 1 [4, 15]. The close similarity between the values of k′xy and k′xq for the different fluorescent solutes and collisional quenchers provides strong evidence that the solvent-solute transfer occurs collisionally with p = 1.

    TABLE 2

    Rate parameters of solvent-solute energy transfer (k’xy) to various fluorescent solutes, and of impurity quenching of solvent fluorescence (k’xq) by various collisional quenchers. (Units 10⁹M−1s−1)

    See Ref [4] for reference to the individual observers.

    Voltz et al. [21] propose that the solvent-solute transfer occurs non-collisionally by dipole-dipole interaction over the Forster critical transfer distance R0. In the Forster theory [22] R0 is defined as the separation distance of individual donor and acceptor molecules at which the rate of donor-acceptor transfer is equal to the rate of de-excitation of the donor molecule by other means.

    Voltz et al. [21] equate the latter to kx~4 × 10⁷s−1, the reciprocal of the bulk solvent fluorescence lifetime tx~25 ns, which approximates to that of an individual donor (solvent) molecule in dilute solution. They thus obtain values of R0 = 16.5 - 20 Å for transfer from the alkyl benzenes to various solutes. This calculation is valid for energy transfer in a dilute solution of donor and acceptor molecules in an inert solvent. It is considered to be incorrect when the donor forms the bulk solvent. In the latter case the lifetime and fluorescence quantum efficiency of an individual excited donor molecule are only ∼10−4 those of the bulk solvent, because of the efficient solvent excitation migration. Unless X* and Y are adjacent, the rate kxx∼4 × 10¹¹s−1 of excitation migration from X* to a neighboring solvent molecule X exceeds the rate of energy transfer from X* to Y. Substitution of kxx as the de-excitation rate of the excited donor molecule in the Förster relation reduces RO to the collisional separation distance. This is consistent with the Birks-Conte model and with the data of Table 2. Similar considerations explain why the solvent-solute transfer in anthracene/tetracene crystals occurs collisionally and not over the Förster transfer distance RO∼35 Å, evaluated for isolated anthracene and tetracene molecules [4].

    Voltz et al. [21] proposed that the excitation migration in an aromatic liquid is similar to that in a crystal. If X* and X are excited and unexcited solvent molecules, and A, B, C …, are different neighboring solvent molecules excitation migration is considered to occur by the processes

    (18)

    + D). A solvent excitation migration frequency kxx∼4 × 10¹¹s−1 is equivalent, from (15), to a molecular interaction energy ΔW∼13 cm−1. Voltz et al. propose that ΔW originates from Coulombic (octupole-octupole) interaction, which they estimate to be of the required order of magnitude. Electron exchange interaction may also contribute to ΔW. Davydov splitting of the triplet energy levels in benzene [23], naphthalene [24] and anthracene [25] crystals has been observed and it is attributed to an electron-exchange interaction ΔW∼10-20 cm−1.

    The Voltz model in its original form neglects solvent excimer formation, although it has been subsequently modified to include excimers as excitation migration traps [38]. The observation of excimer fluorescence in the alkyl benzenes [26] and the determination of the excimer binding energy B∼2000 cm−1 [27] show the existence of strong intermolecular interactions between adjacent excited molecules (X*) and unexcited molecules (X) leading to excimer (D*) formation,

    (19)

    The excimer interaction, which originates from configurational mixing of charge-resonance and exciton-resonance states [28], greatly exceeds the magnitude of ΔW originating from octupole-octupole and/or electron-exchange interaction, and it should therefore be included in any realistic model of excitation migration in the alkyl benzenes. In the alkyl benzenes at room temperature X* and D* in (19) are in dynamic equilibraium i.e., the rates of D* formation and dissociation greatly exceed the rates of X* decay (kx) and D* decay (kD). Under these conditions, the X* and D* fluorescences decay exponentially with a common lifetime

    (20)

    where x and d are the fractions of excited species in the X* and D* states, respectively [15].

    The Birks-Conte model [15] proposes that the solvent excitation migration is due to successive excimer formation and dissociation processes, i.e.

    (21)

    An analysis in terms of this model [, has been proposed [4, 40].

    A solvent excitation migration process related to (21) has been recently observed in pyrene crystals [29]. The pyrene crystal is dimeric in structure, in that the molecules are arranged in parallel pairs (AA′), (BB′), (CC′) etc. The crystal absorption spectrum is molecular (X + hv →X*), but the crystal fluorescence originates from the excimer D* which is formed rapidly from the interaction of X* and its parallel neighbor [30]. Klöpffer [29] has observed solvent-solute energy transfer in crystal solutions of coronene in pyrene. The dependence of the solute fluorescence quantum yield ϕxy on Cy is consistent with (6). The pyrene exciton migration frequency kxx∼10¹¹s−1 at room temperature, but it decreases rapidly with decrease in temperature, becoming negligible at 77°K.

    The possible processes involved in the exciton migration in pyrene crystals are shown diagrammatically in Figure 1. The downward transitions, which are rapid, correspond to excimer formation by a pair of adjacent parallel molecules. The upward transitions, which require thermal activation, correspond to dissociation of the excimer pair. The upper horizontal transitions correspond to molecular exciton migration between non-parallel molecules belonging to different neighboring pairs. The lower horizontal transitions correspond to possible excimer exciton migration between non-parallel neighboring pairs. The molecular exciton and excimer exciton migration processes can arise due to Coulombic and/or electron-exchange interactions between adjacent translationally inequivalent molecules and pairs, respectively. The molecular exciton migration process requires a thermal activation energy corresponding to the pyrene excimer binding energy B∼3400 cm−1 [30], while the excimer exciton migration process should be insensitive to temperature. The strong temperature dependence of kxx which is observed [29] indicates that this exciton migration occurs primarily by the former process. The overall sequence may be written in the form

    Fig. 1 Possible processes involved in exciton migration in a pyrene crystal. X* = excited molecule, X = unexcited molecule, D* = excimer, (AA′), (BB′), (CC′) = molecular dimer pairs.

    (22)

    The molecules (AA′), (BB′) represent pairs of adjacent molecules which are suitably oriented for excimer formation. The molecules (A’B), (B’C) represent pairs of adjacent molecules which are not suitably oriented for excimer formation, but which can interact to give intermolecular excitation migration. Steps (i) and (iii) are equivalent to the Birks-Conte model (21). Steps (ii) and (iv) are equivalent to the Voltz model (18).

    The combined model described by (22) is also applicable to solvent excitation migration in liquids. In a pyrene crystal the different pairs (AA′) and (A’B) are defined by the crystal structure. In an aromatic liquid the pairs are continually changing due to molecular rotation, collisions and diffusion. It is hoped to assess the relative importance of steps (i) and (iii) and of steps (ii) and (iv), in the overall solvent excitation migration process in aromatic liquids by studies of the temperature dependence of kxx. Such studies are currently in progress in the author’s laboratory at the University of Manchester.

    Delayed Scintillation Component

    King and Voltz [6] described the origin of the delayed scintillation component in unitary scintillators in terms of the migration of triplet (T1x) excitons from the ionization track and their subsequent bimolecular interaction.

    (12)

    leading to delayed S1x fluorescence. The delayed scintillation decay of anthracene and trans-stilbene crystals agrees with that predicted by this model. Optical studies in which the triplet excitons are directly excited by ruby laser irradiation of an anthracene crystal [31] confirm that the delayed fluorescence originates from T1x-T1x interaction.

    A similar delayed scintillation emission occurs in binary liquid solution scintillators. Voltz and Laustriat [32] have proposed that this originates from the diffusion of triplet-excited solvent molecules (T1x) from the ionization track, leading to bimolecular T1x interaction (12), followed by solvent-solute energy transfer:

    (23)

    leading to delayed S1y fluorescence. An analysis of the scintillation decay of oxygen-saturated and air-saturated toluene solutions containing 8 gl−1 PBD in terms of the Voltz-Laustriat (VL) model [33] yields triplet lifetimes of 110 ns and 400 ns, respectively, which extrapolate to a triplet lifetime of tT= 1.3 µs in the absence of oxygen. Voltz et al. [33] identify tT with the solvent triplet (T1x) lifetime (tT)x. This identification is considered to be incorrect. It is proposed that tT corresponds to the solute triplet (T1y) lifetime (tT)y.

    The VL model neglects the triplet-triplet energy transfer process

    (24)

    This electron-exchange interaction process occurs as an efficient diffusion-controlled process in fluid solutions [(Cy)m, as in 8 gl−1 PBD in toluene solutions, the quantum efficiencies of both transfer processes are of the order of unity. The delayed scintillation component in binary liquid solutions is attributed to T1x-T1y energy transfer (24), followed by diffusion-controlled bimolecular interaction of triplet-excited solute molecules

    (25)

    leading to delayed S1y fluorescence. The delayed scintillation decay is of the same form as that predicted by the VL model, but tT is equated to the solute triplet (T1y) lifetime (tT)y, and not to the solvent triplet lifetime (tT)x. Spurny [34] has also concluded that the delayed scintillation emission of binary benzene and toluene solution scintillators arises from T1y-T1y interaction (25).

    Studies [35, 36] of the triplet lifetimes of benzene and its alkyl derivatives in fluid solutions at room temperature indicate that (tT)x∼10-20 ns, which is comparable with the singlet lifetime tx. Concentration quenching of T1x, associated with triplet excimer formation [4], is one factor which reduces the triplet lifetime of the liquid alkyl benzenes at room temperature. In o-xylene (tT)x is reduced from ∼900 ns in dilute solution to 13 ns in the pure liquid [37].

    A kinetic analysis of the sequence of processes leading to the delayed S1y emission gives the following expression for the delayed scintillation yield of a binary liquid solution:

    (26)

    where

    kTx = rate of S1x - T1x intersystem crossing

    kx = rate of S1x decay = 1/tx

    qTx = kTx/kx

    kzT = rate parameter of T1x-T1y energy transfer (24).

    kT = rate of T1x decay = 1/(tT)x

    αzT = kzT/kT

    kyzz = rate parameter of T1y-T1y interaction yielding S1y and S0y(25).

    kzz = rate parameter of all T1y-T1y interactions leading to quenching of T1y.

    The first term in (26) is the T1x yield which has two components: G(T1x) produced by ion recombination and excitation, and a second component produced by S1x-T1x intersystem crossing. The second term in (26) is the quantum efficiency of T1x-T1y energy transfer. The third term is the quantum yield of S1y from T1y-T1y interaction, and the final term is the S1y fluorescence quantum yield.

    1/σxy, 1/σzT) the prompt and delayed scintillation yields are, respectively,

    (27)

    (28)

    Fuchs et al. [39] have determined G0(hvy)/GO(hvy)del = 0.85/0.15 for a deoxygenated solution of 5 gm/litre α-NPO in benzene excited by fast electrons. Substituting this ratio and G0(S1x) = 1.6, G0(T1x). = 1.7 [7] for benzene in (27) and (28) we obtain

    (29)

    for α-NP0.

    There are nine possible intermediate excited dimer states, one singlet, three triplets and five quintets, produced in a triplet-triplet bimolecular encounter. In a fluid solution it is to be expected that these will be produced with equal probability (multiplicity weighting). Only the singlet dimer state can dissociate into S1y and S0y, so that the rate parameter of process (25) is

    (30)

    where kdiff is the diffusion-controlled T1y-T1y encounter rate parameter and p1 ≤1. The triplet and quintet dimer states may quench (i.e. yield S0y and S0y), with probabilities of p3 (≤ 1) and p5 (≤ 1), respectively, they may dissociate into T1y and T1y, or they may yield T1y and S0y, thereby increasing the unimolecular decay rate. Hence

    (31)

    (32)

    The parameter w has a maximum value of 1 when p3 = p5 = 0, and a minimum value of (1 + 8/p1)−1. The α-NPO value of w = 1/3 is consistent with p3 = 2p1/3, p5 = 0. Further experiments are required to determine kyzz and kzz for α-NPO and other solutes.

    I am grateful to Drs. D. L. Horrocks, W. Klöpffer, F. Wilkinson and R. Voltz for useful discussions, and to the Argonne National Laboratory for their hospitality.

    References

    1. Birks, J.B. Proc. Phys. Soc. 1951; A64:874.

    2. Heckmann, P.H. Z. Phys. 1959; 157:139.

    3. Birks, J.B.The Theory and Practice of Scintillation Counting. Oxford: Pergamon Press, 1964.

    4. Birks, J.B.Photophysics of Aromatic Molecules. London and New York: Wiley-Interscience, 1970.

    5. J. B. Birks, J. Phys. B. (Atomic and Molecular Physics) in press.

    6. King, T.A., Voltz, R. Proc. Roy. Soc. 1966; A289:424.

    7. Cooper, R., Thomas, J.K. J. Chem. Phys. 1968; 48:5097. [J. K. Thomas, Private communication (1969)].

    8. Skarstad, P., Ma, R., Lipsky, S. Mol. Cryst. 1968; 4:3.

    9. Lawson, C.W., Hirayama, F., Lipsky, S. J. Chem. Phys. 1969; 51:1590.

    10. Hayes, F.N., Rogers, B.S., Sanders, P.C. Nucleonics. 1955; 13(No 1):46.

    11. Horrocks, D.L. J. Chem. Phys. 1970; 52:1566.

    12. Wolf, H.C. Advances in Atomic and Molecular Physics. 1967; 3:119.

    13. Benz, K.W., Wolf, H.C. Z. Naturforsch. 1964; 19a:177.

    14. Korsunskii, V.M., Faidysh, A.N. Opt. Spectr Supp. 1966; 1:62.

    15. Birks, J.B., Conte, J.C. Proc. Roy. Soc. 1968; A303:85.

    16. Wright, G.T. Proc. Phys. Soc. 1953; A66:777.

    17. Lipsky, S., Burton, M. J. Chem. Phys. 1959; 31:1221.

    18. Swank, R.K., Buck, W.L. Phys. Rev. 1953; 91:927.

    19. Birks, J.B., Kuchela, K.N. Disc. Faraday Soc. 1959; 27:57.

    20. Berlman, I.B. J. Chem. Phys. 1961; 34:598.

    21. Voltz, R., Laustriat, G., Coche, A. CR. Acad. Sci. Paris. 1963; 257:1473. Voltz, R., Klein, J., Heisel, F., Lami, H., Laustriat, G., Coche, A. J. Chem. Phys. 1966; 63:1259.

    22. Förster, Th. Naturwiss. 1946; 33:166. [Ann. Physik 2, 55(1948); Z. Naturforsch 4a, 321(1949); Disc. Faraday Soc 27, 7(1959)].

    23. Burland, D.M., Castro, G. J. Chem. Phys. 1969; 50:4107.

    24. Hanson, D.M., Robinson, G.W. J. Chem. Phys. 1965; 43:4174. Castro, G., Robinson, G.W. J. Chem. Phys. 1969; 50:1159.

    25. Clarke, R.H., Hochstrasser, R.M. J. Chem. Phys. 1967; 46:4532. Avakian, P., Ern, V., Merrifield, R.E., Suna, A. Phys. Rev. 1968; 165:974.

    26. Ivanova, T.V., Mokeeva, G.A., Sveshnikov, B.Ya. Opt. Spectr. 1962; 12:325.

    27. Birks, J.B., Braga, C.L., Lumb, M.D. Proc. Roy Soc. 1965; A283:83.

    28. Vala, M.T., Hillier, I.H., Rice, S.A., Jortner, J. J. Chem. Phys. 1967; 44:23.

    29. W. Klöpffer, Private communication (1970). Paper this Conference.

    30. Birks, J.B., Kazzaz, A.A. Proc. Roy. Soc. 1968; A304:291.

    31. Avakian, P., Merrifield, R.E. Mol. Cryst. 1968; 5:37.

    32. Voltz, R., Laustriat, G. J. Physique. 1968; 29:159.

    33. Voltz, R., Dupont, H., Laustriat, G. J. Physique. 1968; 29:297.

    34. Spurny, F. Collection Czechoslov. Chem. Comm. 1970; 35:565.

    35. Wilkinson, F., Dubois, J.T. J. Chem. Phys. 1963; 39:377.

    36. Cundall, R.B., Griffiths, P.A. Trans. Faraday Soc. 1965; 61:1968.

    37. Cundall, R.B., Voss, A.J.R. Chem. Comm. 1969; 116.

    38. Voltz, R. Radiation Res. Rev. 1968; 1:301

    39. R. Voltz, Private communication (1970). C. Fuchs, F. Heisel, R. Voltz and A. Coche. Paper this Conference.

    40. J. B. Birks, M. I. T. Conference on Current Status of Liquid Scintillation Counting (1969). In press.


    *Work performed under the auspices of the U. S. Atomic Energy Commission.

    **On leave of absence from Atomic And Molecular Physics Group, The Schuster Laboratory, University of Manchester, England.

    ON AN EMPIRICAL CORRELATION BETWEEN NUCLEAR GEOMETRY AND CERTAIN SPECTROSCOPIC PARAMETERS OF AROMATIC COMPOUNDS*

    Isadore B. Berlman,      Radiological Physics Division, Argonne National Laboratory, Argonne, Illinois 60439

    Abstract

    From empirical spectroscopic data it is deduced that as the nuclear conformation of a chromophore becomes more planar and linear, the fluorescence and absorption spectra become more structured, narrower, and long-wavelength shifted, the Stokes loss becomes smaller, and the maximum value of the molar extinction coefficient, εmax, becomes larger. Moreover, compounds planar in both the ground state and first excited state are usually more susceptable to excimer formation and/or concentration quenching than nonplanar compounds. The fluorescence and absorption characteristics of several compounds are presented as examples.

    Introduction

    The fluorescence and absorption characteristics of a large number of aromatic compounds in dilute solution have been measured and many of these data have already been published.¹ An attempt is now made to correlate some of these findings with the topology of the chromophore in the ground state as well as in the excited state. The characteristics that will be considered are due to π-π* transitions and since π electrons are generally delocalized over the whole conjugated system, these characteristics are expected to be influenced by the geometry of the system. Our conclusions, derived from empirical results and supported by many examples, are presented in tabular form. In fact, it is our contention that when such data are applied judiciously they can be used as a spectroscopic straightedge to provide qualitative evidence concerning the planarity of an aromatic chromophore in its ground as well as in its first excited singlet state.

    Procedures

    Standard procedures¹ are employed in making the measurements. Briefly, the fluorescence and absorption spectra of a compound in a dilute, nitrogen-bubbled cyclohexane solution are recorded. The quantum yield Q is determined from the ratio of the area of the fluorescence spectrum (photon intensity versus wave number) of the sample to that of 9, 10-diphenylanthracene after proper adjustments are made as described in Ref. 1. The natural fluorescence lifetime τ0 is computed from the data from both the fluorescence and absorption curves, according to the method of Strickler and Berg, ² rather than that of Forster described in Ref. 1. The fluorescence decay time τ is measured in a separate experiment. Stokes loss (in cm−1) is determined from the energy difference between the 0-0 transition and the first moment of the fluorescence spectrum. Because the widths of the spectra of various compounds can vary by more than a factor of 2, a simple method is used to determine the relative breadth of each curve, namely measuring the full width (in cm−1) at reciprocal e (FWRE) of the peak value. FWRE (Fl) is a measure of the breadth of the fluorescence spectrum, and FWRE (Abs) that of the absorption spectrum. The measurement of a structured spectrum is inexact because it depends on an estimation of the wave number at which the curve has a value of 37% of the maximum. A better defined quantity, the statistical width W, is given by the equation

    ) and v−(2) is the second moment, where

    The quantity W was not computed for the absorption spectra because of the frequent overlap of adjacent bands from other transitions. Values of the various parameters are assembled in Table I.

    Table I

    Absorption and Fluorescence Characteristics of Some Aromatic Compounds.

    Emphasis is placed on τ the fluorescence decay time and Q the fluorescence quantum yield because their ratio Q/τ is proportional to the strength of the fluorescence transition. When the strength of the fluorescence transition is equal to the absorption transition and when there are no second-order processes, τ/Q = τ0, where τ0 is the natural fluorescence lifetime and is computed by an integration over τ values of the absorption spectrum. The ratio τ/Q is an experimentally derived quantity and it is particularly helpful in identifying the lowest excited singlet state. This is especially true when the long wavelength absorption bands are obscured by bands from a more intense transition of slightly higher energy, and it is difficult to evaluate τ0 properly. Moreover, the fluorescence transition (proportional to Q/τ) is usually not equal to the absorption transition (proportional to τ0−1) when a change in conformation takes place on excitation.

    Whereas the discussion above was concerned with measurements on dilute solutions, mention is now made of the effect of the solute concentration on the intensity and shape of the fluorescence spectrum, the fluorescence decay time, etc. In the case of planar compounds, i.e., chromophores plus substituents, as the concentration is increased, the intensity and decay time of the monomer emission are both reduced by a process called self-quenching and/or a process called excimer formation (in which a new long-wavelength structureless emission band is generated). The effects of both processes are proportional to the solute concentration. Since the radiative lifetime of excimer emission is often longer than that of monomer emission, the fluorescence decay curve is usually composed of two components.

    Proposal and Examples

    Certain correlations between conformation and spectroscopic data are well known:³ planar compounds that have the same nuclear equilibrium configuration in their ground state and first excited singlet state (do not change their geometry on excitation) usually have structured fluorescence and absorption spectra; nonplanar compounds generally have diffuse absorption spectra. Moreover, when the nuclear geometry of a chromophore in the ground state is made more planar, e.g., by bridging, the absorption spectrum becomes more structured and red-shifted, and the values of ε become larger. If the same chromophore is at the same time made more planar in the excited state, the fluorescence spectrum also becomes more structured and red-shifted. On the other hand, when the nuclear topology is made more nonplanar, e.g., by steric hindrance, the spectra become diffuse and blue-shifted. As a rule of thumb, in a series of similar molecules, the larger the number of π electrons in a chromophore, the greater the shift to the red of the fluorescence and absorption spectra. Thus the characteristics of a planar chromophore can be interpreted as being due to the fact that in this configuration the number of functional π electrons is a maximum.

    Somewhat less well known are the following: the width W of each spectrum and the Stokes loss become large when a change in conformation accompanies the electronic transition; the strength of the fluorescence transition, as measured by Q/τ, becomes larger with increased planarity; and the susceptibility to concentration quenching and excimer formation increases with the planarity of the whole molecule (chromophore and substituents). Thus the spectroscopic data can be related to the conformation of the main chromophore and the concentration studies to the conformation of the whole molecule.

    Our data lead us to believe that the amount of structure in a spectrum depends on the planarity of the π system rather than its rigidity. Therefore, emphasis is placed on the correlation between spectroscopic data and planarity, not rigidity. Yet rigidity is important in maintaining a nuclear configuration on excitation or at least in restricting the change in configuration on excitation. In the present discussion a chromophore is considered rigid if there is no free or hindered rotation about a single bond by a component π-system, such as a phenyl or vinyl group, and nonrigid if there is rotation by such groups.

    The emphasis of this study is on ring systems and not on the number and type of their substituents because the latter may introduce phenomena other than those of conformal changes. Substituents such as hydroxy and amine groups in phenol and aniline, respectively, can change the charge distribution in the ring, and compounds containing these substituents have fluorescence and absorption spectra that are structureless. Compounds with substituents will be employed as examples but mainly to illustrate phenomena related to conformation. A phenyl group is one of the most common types of substituent on scintillation materials. Because the ring system is generally prevented by steric hindrance from being coplanar with the main chromophore, a decision has to be made whether to include it as part of the functional chromophore in this discussion of the relationship between conformation and spectroscopic data. Our conclusion was that if the preferred angle between the plane of the ring and the plane of the chromophore is greater than about 60° so that the interaction between the two systems is small (the main effect being that the spectra are red-shifted about 100 Å) the ring system(s) is not considered as part of the main chromophore; if the interplanar angle is less so that the interaction is greater, the ring system is considered as part of the active chromophore.

    Spectroscopically, fluorescence compounds can be divided into five classes - those whose chromophores are (1) planar in the ground state and first excited state, (2) non-planar in both states with the geometry preserved, in the main, on excitation, (3) nonplanar in the ground state and more planar in the excited state, (4) nonplanar in the ground state but more nonplanar in the excited state, and (5) planar in the ground state and nonplanar in the excited state. Whereas rigid and nonrigid chromophores can be in any of these classes, the majority of the rigid chromophores are found in the first two classes. The relative manner in which the absorption and fluorescence characteristics vary with respect to the conformation of a chromophore in its ground and first-excited singlet state is tabulated in Table II.

    Table II

    The Effect of Conformation on Certain Absorption and Fluorescence Characteristics.

    Examples of the above classes are given below and some of their spectroscopic data are found in Table I. Anthracene¹ is an example of a rigid compound that is planar⁴ in the ground state. Because the spectra are structured and narrow, and Stokes loss is small, it is inferred that this compound is also planar in the excited state (class 1). In solution, the fluorescence intensity is very sensitive to self-quenching.⁵ When phenyl substituents are added to certain positions, e.g., in 9, 10-diphenyl anthracene, the concentration effects are completely eliminated.⁶ Jones⁷ has estimated that in the ground state the plane of each phenyl group makes an angle of about 57° with the plane of the anthracene chromophore. Supporting evidence for this large angle is the small interaction between the two π systems and the resulting small red shift of the spectra, ¹ even though the substituents are positioned along the direction of the transition moment. The values of τ/Q for 9-phenylanthracene and 9, 10-diphenyl anthracene are about 30% larger than τ0 obtained by integrating over the absorption curve, indicating that the strength of the fluorescence transition has become weaker than that of the absorption transition. For this to happen, either the phenyl rings assume a more perpendicular orientation with respect to the basic chromophore when the molecule becomes excited, or the anthracene chromophore becomes slightly warped. The former seems more likely from an energy point of view.

    Examples of nonrigid chromophores that remain planar in both states are rare. Most ring-chain systems such as trans-stilbene may be planar in the ground state, but apparently become nonplanar in the excited state (class 5).⁸, ⁹

    The compound 3, 4-benzophenanthrene (Fig. 1) is an example of a rigid compound that is nonplanar in the ground state.¹⁰ Because its spectra are slightly structured and narrow, Stokes loss is small, and it is relatively insensitive to concentration effects, it is presumed that this compound is also nonplanar in the excited state (class 2 or class 4). When this chromophore is made more planar by bridging, as in benzo[ghi]fluoranthene (Fig. 2) (class 1), both spectra become more structured and are red-shifted about 600 Å. Stokes loss and W remain small. Although dibenzophenanthrene is even more nonplanar in the ground state than 3, 4-benzophenanthrene, yet its bridged analogue, benzo(ghi)perylene (Fig. 3) has structured spectra and appears to be planar (class 1).

    Fig. 1 Spectra of 3, 4-benzophenanthrene.

    Fig. 2 Spectra of benzo(ghi)fluoranthene.

    Fig. 3 Spectra of benzo(ghi)perylene.

    Octamethyl biphenyl¹¹ is a nonrigid compound that belongs to class 2 because it is nonplanar in both states. That the angle between the planes of the rings is so large that there is very little interaction between them is supported by the evidence that the fluorescence and absorption characteristics correspond to those of a single ring with alkyl substituents (cf. 1, 3, 5-trimethylbenzene¹): Its absorption spectrum is slightly structured and narrow, and Stokes loss is small. Because each half of the molecule absorbs independently, the intensity of the absorption spectrum is about twice that of a single substituted ring. There is some interaction between the rings because the spectra are red-shifted with respect to alkyl-substituted benzene derivatives.

    The compound 1, 1-diphenylethylene is nonrigid and is made nonplanar in both states by steric hindrance. The angle between the plane of a phenyl ring and that of the ethylene bond is estimated¹² to be about 44° in the ground state. Because the spectra¹ of this compound are diffuse, Stokes loss and FWRE (F1) are large, and the quantum yield is less than 0.01, it is conjectured that a large change in conformation takes place on excitation (class 4).

    Compounds planar in the excited state usually have a larger quantum yield than related, nonplanar compounds. Thus, 1-methyl-3, 2′-methylene-2-phenyl indole (class 1) which has the characteristics¹³ of being planar in both states has a quantum yield of 0.91, and 1-methyl-3, 2′-trimethylene-2-phenyl indole (class 3) which has the characteristics¹³ of being nonplanar in the ground state and only slightly more planar in the excited state has a quantum yield of 0.58 (see Table I). (The latter of the above two compounds is also less sensitive to concentration quenching.)¹⁴ Moreover, cis-stilbene, whose rings in the ground state are about 30° out of plane, in a propeller configuration¹⁵, ¹⁶ has almost zero fluorescence quantum yield at room temperature. Yet, phenanthrene (class 1), which can be considered a bridged and more planar analogue of cis-stilbene, has structured spectra and a quantum yield of about 0.13. Triphenylene and o-terphenyl are similarly related compounds. The angle between the planes of adjacent rings of o-terphenyl in solution has been estimated¹⁷ to be about 43°. Its absorption spectrum is diffuse and blue-shifted, ¹⁸ and has a very small fluorescence quantum of less than 0.01.¹⁹ Triphenylene, which can be considered a bridged and relatively more planar analogue of o-terphenyl, has very structured spectra and a quantum yield of about 0.08.¹

    Biphenyl is a nonrigid compound that belongs to class 3. Its absorption spectrum is diffuse and broad, the fluorescence is structured and narrow, and Stokes loss is large. All of the fluorescence characteristics of biphenyl, i.e., the shape of its spectra, quantum yield, decay time, etc., are invariant over a concentration range from 0.1 to over 80 g/1. This immunity to concentration is characteristic of a molecule that is nonplanar in one or both of its states. Suzuki²⁰ has estimated that in solution, the angle between the planes of the rings (in the ground state) is about 23°. Because the fluorescence spectrum is structured, it is inferred that this compound is relatively planar in the first excited state: the added resonance interaction between the rings in the excited state gives some double-bond character to the coannular bond, and forces the rings into a more planar configuration.²¹

    When the rings of biphenyl are bridged as in fluorene, ¹ dibenzofuran, ¹³ and carbazole, ¹ the spectra are structured and narrow, and Stokes loss is small - characteristic of compounds in class 1. Fluorene and dibenzofuran²² are sensitive to concentration, forming excimers, but the solubility of carbazole is too small to test for concentration effects. In 9, 10-dihydrophenanthrene the rings are about 23° with respect to each other²³ (in the ground state) because the bridge joining them is larger than those of the above compounds. The absorption curve is structureless.¹¹ Because of the slight structure in the fluorescence spectrum it is assumed that the rings assume a more planar geometry in the excited state (class 3). Stokes loss and W, although small, are nevertheless larger than those of the above-bridged biphenyl compounds (Table I). There is no evidence of concentration effects over a range from 0.1 to 20 g/1.

    The compound 3-phenyldibenzofuran (Fig. 4) and benzo-(1, 2b:4, 5b′)bisbenzofuran (Fig. 5) can be considered as oxygen bridged p-terphenyl derivatives. These bridges make the compound more planar. Therefore, the spectra are structured and narrow, and Stokes loss is small.

    Fig. 4 Spectra of 3-phenyldibenzofuran.

    Fig. 5 Spectra of benzo(1, 2-b, 4-b′)bisbenzofuran.

    Ring-chain systems, such as trans-stilbene, 1, 6-diphenylhexatriene, ¹ and 1, 8-diphenyloctatetraene¹³ have characteristics of members of class 5. Their absorption spectra are well structured and the absorption transitions are very strong (εmax is large), indicating that at least in the ground state they are planar (all trans form). Yet, each of these has peculiar fluorescence characteristics. That trans-stilbene on excitation rotates with a finite probability into the cis form is well known.¹⁵ The fluorescence decay time of the other two compounds is more than 10 times larger and the quantum yield is smaller than would be expected if the fluorescence transition were equal to that of the absorption. Independent studies of polarization by W. R. Anderson¹⁹ have shown that the fluorescence characteristics of the latter two compounds are not produced by a hidden transition. Therefore, it is assumed herein and in Ref. 24 that these large decay times are produced by molecules assuming various isomeric forms on excitation. In these forms, the fluorescence transition is much weaker than the absorption transition of the all trans form. The very broad spectra and large Stokes loss support this contention of a change in configuration on excitation. Forster²⁵ has also suggested that the large value of Stokes loss for 1, 8-diphenyloctatetraene could be explained by a large change of the nuclear configuration on

    Enjoying the preview?
    Page 1 of 1