Вы находитесь на странице: 1из 37

Pulsed optoacoustic spectroscopy of condensed riiatter

C. K. N. Patel
Bell Laboratories, Murray Hill, New Jersey 07974

A. c. Tarn
IBM Research Laboratory, San Jose, California 95193
The authors discuss the theory and experiments dealing with the pulsed optoacoustic effect (i.e., generation of
a transient acoustic wave by absorption of an optical pulse) in condensed matter. Their primary interest lies in
the measurement of small absorption coefficients (&10 ' cm '). At present an experimental capability of
measuring absorption coefficients as small as 10 ' cm ' has been demonstrated, and further improvement is
foreseen. The pulsed optoacoustic absorption measurement technique has been applied to the following linear
spectroscopic studies: t 1) precise measurements of the optical absorption spectra of H, O and D, O; (b) accurate
determination of absorption strengths and profiles of high harmonics (n = 6, 7, and 8) of vibrational modes in
transparent organic liquids (e.g. , benzene); (c) quantitative absorption spectra of thin(-1— 10 p m) liquid films;
and (d) quantitative absorption spectra of solids and finely powdered crystals. The usefulness of the pulsed
optoacoustic technique to nonlinear spectroscopy has been demonstrated in the following studies: (a)
quantitative two-photon absorption spectroscopy of the weak two-photon ('B, „~'3, ) transition in benzene;
and (b) optoacoustic Raman-gain spectra for a variety of liquids where an ability to measure Rarnan gains as
small as 10 ' cm ' has been demonstrated. In addition to reviewing the above studies the authors discuss
future possible applications and compare the pulsed optoacoustic spectroscopy technique with other
optoacoustic absorption measurement techniques.

CONTENTS l. j NTRGDUCT ION


Introduction 517
Theory 518 The field of optoacoustic (OA) spectroscopy of con-
A. Optoacoustic signal generation 518 densed matter has undergone tremendous growth in the
1. Phenomenological description of OA signal
gener at ion 519
past several years. Hordvik (1977}, Robin (1976),
2. Rigorous description of OA signal generatio 520 Rosencwaig (1978), and Somoano (1978) have reviewed
B. Acoustic pulse arrival time 522 various aspects of the field, and Pao (1977}has edited
C. Calibration of OA spectroscopy 522 a book on the subject. The "conventiona'1" technique
D. Sensitivity 522 of optoacoustic spectroscopy (also called photoacoustic
Apparatus 523 spectroscopy) utilizes a gas-phase microphone, which
A. Transducer 523 senses the heating and cooling of a gas layer in thermal
B. Transducer preamplifier 524
C. OA cell contact with the sample of condensed matter being ir-
525
D. Arrangement for films and powders 525 radiated by a chopped light beam. Such a technique was
IV. Experimental Procedure 526 first used by A. G. Bell and others and has been re-
V. Experimental Results 529 vived recently by Harshbarger and Robin (19'73) and
A. Linear spectroscopy 530 Rosencwaig (1973). This recent work is an extension
1. Water 530 of the earlier gas-phase optoacoustic spectroscopy
2. Transparent organic liquids 532 work (e. g. , Kreuzer, 1971). Analytical studies of the
3. Thin liquid films 534 signal dependence on the chopping frequency, absorp-
4. Powders and solids 535
B. Nonlinear spectroscopy 537
tion coefficient, and thermal properties have been
1. Two-photon absorption 538 carried out, for example, by Rosencwaig and Gersho
2. Raman-gain spectroscopy 540 (1976), Aamodt and co-workers (1977, 1978), and
VI. Other OA Spectroscopy Studies 542 McDonald and Wetsel (1978). However, the gas-phase
A. OA spectroscopy and luminescence spectrosco py 542 microphone technique for condensed samples relies on
B. Pulsed OA spectroscopy in strongly absorbing inefficient thermal diffusion into a gas, and can be
liquids viewed actually as a photothermal (Brilmyer ef al. ,
C. Other phenomena at high intensities
D. cw-source-based OA spectroscopy with 1977) technique and not a photoacoustic one, since the
piezoelectric transducer acoustic signal generated in the sample plays only a
E. Application of OA spectroscopy to trace minor role. Thus this technique has low sensitivity
detection in liquids 545 (Hordvik and Schlossberg, 1977; Farrow et a/. , 1978;
VII. Discussions 545 McQueen, 19'19), and is often only useful for absorp-
A. Summary of present status 545 tion typically exceeding 1%. Some of the disadvantages
B. Examples of further experiments 546
of the inefficient coupling are partially overcome by
C. Other OA methods 546
VIII. Acknowledgments 547 the availability of high-efficiency gas-phase micro-
Appe ndix: Recent Work in the OA Study of Condensed phones. Several authors (Hordvik and Schlossberg,
Matter 547 1977; Farrow et a/. , 1978) have recognized the in-
References 549 efficiency of the "conventional'* photoacoustic technique,

Reviews of Modern Physics, Vol. 53, No. 3, July 1981 Copyright Amer'ican Physical Society 517
C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

and they show that improvements in sensitivity are 5. Future applications of the technique.
possible with the use of piezoelectric transducers in
Recent literature on OA studies of condensed matter
contact with solid samples (Hordvik and Schlossberg,
1977) or liquid samples (Farrow ef a/. , 1978), but is given in the reference list. We have made an effort
to make the list representative of the current interest
these authors still use light beams chopped at low
in OA studies of condensed matter; however, rapid
audio frequencies.
expansion in the field of optoacoustic spectroscopy
We have recently demonstrated a highly sensitive
makes the omission from our list of some newcon-
pulsed OA spectroscopy technique (Patel and Tam,
tribut ions unavoidable.
1979a), involving the use of pulsed lasers, a piezo-
electric transducer in direct contact with the liquid I I. THEOR Y
sample, and gated detection. This technique is truly
optoacoustic in character (Hordvik, et a/. , 19'77; A. OA signal generation
Farrow ef a/. , 1978) because the original acoustic Numerous authors have treated the problem of acous-
pulse generated in the condensed sample is directly tic generation in a liquid (or in a solid) arising from
detected by a pie zoelec tr ic transducer. We note that the absorption of a light beam whose intensity varies
many authors (Bonch-Bruevich ef a/. , 1975; Burt, with time. Most authors were interested in the case
1979; Callis, 1976; Emmony et «.
, 1976; Farrow of surface heating by a pulsed laser (Askaryan e/ a/. ,
et a/. , 1978; Von Gutfeld and Budd, 1979; Hordvik and 1963; Bell and Landt, 1967; Bushanam and Barnes,
Schlossberg, 1977; Kohanzadeh eg a/. , 1975; Lahmann 1975; Emmony et a/. , 1976; Gordienko eg a/. , 1978;
and Ludewig, 1977; Oda et al. , 1978; Razumova and Gournay, 1966; Von Gutfeld and Budd, 1979; Sigrist
Starobogatov, 1977; Rentzepis and Pao, 1966; Sigrist and Kneubuhl, 1978; Teslenko, 197'7; White, 1963),
and Kneubiihl, 1978; Sladky et a/. , 1977; White, 1963) i.e. , a case where most of the pulsed energy is ab-
have used piezoelectric elements in contact with liquids sorbed in a thin surface layer. Other authors
or solids for optoacoustic detection, and the advantage (Kohanzadeh et a/. , 1975; Rosencwaig, 1978) have
of good acoustic impedance matching is generally treated the case of the steady-state solution for a
recognized. However, our work appears to be the first sinusoidally modulated light beam, so that all time-
to demonstrate the extremely high sensitivity of the varying quantities have the exp(us/) form. However,
technique for detecting weak absorptions in liquids or we are primarily interested in the case when the time-
solids. We further note that Bonch-Bruevich eg a/. dependent signal is due to a pulsed excitation, with low
(1977}have used an optoacoustic detection scheme optical absorption (i.e. , a sample that is highly trans-
similar to ours, but their achieved sensitivity is at parent to the light). This case has been considered
least a factor of 10' lower than ours. by a few authors (Kasoev and Lyamshev, 19'l7;
We he. ve applied our optoacoustic technique to several Naugol'nykh, 1977; Tam and Patel, 1979b} with vary-
problems in linear and nonlinear optical spectroscopy. ing degrees of mathematical rigor. It is interesting to
For linear spectroscopy, we have carried out (1) the note that in each of the three diff@rent cases mentioned
first unambiguous determinations of visible absorption above (surface absorption of a pulsed la, ser, steady-
spectra. of liquid H, O and D, O (Patel and Tam, 1979d; state solution for a sinusoidally modulated light beam,
Tam and Patel, 1979b), (2) a systematic study of the or weak bulk absorption of a pulsed laser) the solution
absorption profiles of high vibrational overtones in is characterized by the following intuitively appealing
benzene (Patel and Tam, 1979b; Patel et al. , 1979d), form:
and (3) studies of the absorption spectra of micron-
thick liquid films (Patel and Tam. 1980) and powder =const ' a,
pv,
films (Tam and Patel, 1979c). We show that, at pres- Cp
ent, absorption coefficients -10 'to 10 ~ cm ' can be where po is the acoustic pressure amplitude developed,
measured. In the area of nonlinear spectroscopy, we E, is the laser energy in a pulse or in a cycle, P is the
have reported (1) the first two-photon absorption thermal expansion coefficient, Cp is the specific heat
spectroscopy and cross-section measurements in at constant pressure, v, is acoustic velocity, and n
liquids by optoacoustic detection (Tam and Patel, is the optical absorption coefficient of the medium.
1979a), and (2) the first Raman-gain spectroscopy of In the following we shall consider in more detail the
liquids (Patel and Tam, 1979c) using the pulsed opto- case of weak bulk absorption of a pulsed laser beam.
acoustic spectroscopy technique. We consider a thin cylinder of a liquid, illustrated
The high sensitivity possible from this pulsed opto- in Fig. 1 (or solid if we consider only longitudinal
acoustic technique points to numerous further applica- waves), of length / and optical absorption coefficient n,
tions. In view of the wide usefulness of the technique, irradiated by a laser pulse of duration 7'~ =2r and pulse
the preserit article wi11. cover the following topics: energy Eo. We are primarily interested in the case of
weak absorption, i.e. , nl «1, so that the absorbed
1. The process of pulsed acoustic generation; energy Eonl is distributed uniformly along the length l. ,

2. The construction of the piezoelectric transducer, Note that the case of large o. (i.e. , surface absorption)
OA cell, and preamplifier; has been treated by several authors (Askaryan et a/. ,
3. Expe rimental results us ing pulsed optoacoustic 1963; Bell and Landt, 1967; Bonch-Bruevich et a/. ,
spe c troscopy; 1975; Bushanam and Barnes, 1975; Emmony et a/. ,
4. Experimental results using other optoacoustic 1976; Gournay, 1966; Sigrist and Kneubuhl, 1978;
methods; and and Teslenko, 1977), and results in sPherical acoustic

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 5't 9

WIDTH = 2T
—ILLUMINATED LIQUID
COLUMN (5)
For el«1, a situation which we want to explore,
&'b. = &On
LASER
. PULSE

Since we are assuming that nonradiative relaxation


predominates in the medium, the thermal energy E,h
is given by
th ab~ 0
OBSERVATION Knowing the specific heat at constant pressure C~, we
POINT
can evaluate the rise in temperature AT of the il-
FIG. 1. Theoretical calculation of the pulsed OA effect. There luminated volume (neglecting thermal conduction} from
is full cylindrical symmetry in the weak absorption limit for a
long thin liquid column irradiated. E,h =CpVATP,
where U is the illuminated volume and p is the density
waves originating from the point of incidence on the of the condensed material. With
surface. In our case, the weak absorption results in a
V =&R2l
cylindhical wave produced by the thin cylindrical
volume. Here we examine the acoustic pressure we obtain
P(r, t) at an observation point as a function of the per-
E~cv
pendicular distance r arid time t. There is no de- QT = (lo)
&R Cpp
pendence on other spatial variables besides r, since
the acoustic source is assumed to be a long thin cylin- Now if we as sume adiabatic, isobar ic expans ion, we
der. can calculate the new volume of the illuminated region.
With ~ being the increase in the radius of the il-
1. Phenomenological description of OA signal generation luminated volume,
The generation of an optoacoustic signal relies on the &R2E —&(R +~}
l =PVb. T,
conversion of the absorbed optical radiation into heat where P is the volumetric expansion coefficient. By
through nonradiative relaxation processes. Rigorous
manipulation of Eq. (10), we obtain for bA «R
derivation of the acoustic pulse generation following
the absorption of a pulsed beam of optical radiation is ,'RP~T . — (12)
given in the next section. Here we shall go through a Inserting b. T from Eq. (10),
phenomenological description which contains most of the
physics underlying optoacoustic signal generation. The
2 AC, p
key parameter is the rapid conversion of the absorbed
energy into heat by nonradiative relaxation processes. This expansion creates a pressure wave which travels
Figure 1 shows schematically a laser beam of radius R radially outwards from the illuminated cylinder at
traversing a cell containing a liquid and a piezoelectric sound velocity. The change in pressure P created at a
transducer placed within. The laser illumination is point is related to the frequency of the sound wave f,
provided by a flash-lamp-pumped dye laser producing and the displacement Ax through
pulses of duration T~ and a repetition rate of f. For the
purposes of the present discussion, we assume that the p = 2 nf. v. exp . (14)
laser pulse length is much longer than the nonradiative But bx is proportional to ~ evaluated in Eq. (12).
relaxation time 7NR, and the time v, for an acoustic
pulse to travel across the laser-illuminated region in
Thus we have Ax = const x ~, and
the liquid, i. e. , =const~ pv,' E o o. .
P R (15)
C~
For a fixed geometry of laser illumination and laser
2R properties, f, and R can be lumped into the constant.
V~ Thus we have
where v, is the acoustic velocity in the medium. Fur- '
ther', we assume that the laser pulse is much longer p =const Eon.
P
than the response time of the piezoelectric transducer,
1 e j
The condition described in Eq. (3) is just satisfied for
R =1 mm, v, =2x10' cm sec ', and 7~=1 —2 psec for
Tpzt
p ~

pzt (4) a flash-lamp-pumped dye laser.


Under these situations, achievable when using a flash- When the illuminated cylinder is "thick" compared
lamp-pumped dye laser, with an input pulse energy of to the acoustic pulse propagation distance during the
Eo per pulse, the energy absorbed by the medium hav- excitation pulse, i.e.,
ing an absorption coefficient of n and length E is given

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


520 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

we can show (Nelson and Patel, 1980) that the transient where v, is the velocity of sound, and S(t') is the time
pressure change caused by absorption of the pulse derivative of the cross-sectional area S(t') of the
optical radiation is given by cylindrical source at retarded time t', i.e. ,
Pv. ' S(t') =~[a+m(t')]2
P =const C Each (18)
= &ft'+2&It m(t'). (26)
when use is made of sufficiently fast transducers.
Here R is the unperturbed radius of the cylindrical
Thus some care has to be exercised in using the func-
source and ~(t') is the expansion at time t' due to the
tional dependence described in Eq. (16) or Eq. (18)
laser pulse. We shall see that ~(t') is always many
for comparing optoacoustic signals from different orders of magnitude smaller than R for cases of in-
experiments.
Since the electrical signal output from the piezo-
terest, so that we are justified in retaining only the
lowest-order term in Eq. (26).
electric transducer is given by Now we derive P by using Eq. (24). Since a limit of
V, =const&&P (19) integration in Eq. (25) is t dependent, it is convenient
(Landau and Lifshitz, 1959) toha've a change of varia-
we obtain
ble t" = [t —t' (r/v, )—
] in Eq. (4) before differentiation,
and change back to &' after differentiation. Thus we
(20) first write
where K' is a constant that includes the geometrical
parameters as well as the response properties of the
piezoelectric transducer. Now we see that the norma- @(r, t) =— (2'7)
lized optoacoustic signal S is given by (for r~» r, )
a, nd hence we obtain
V)3
Eo
fbi Pvg
Cp
™ (21)
(28)

Thus a measurement of the normalized optoacoustic Using Eq. (26), we have


signal for a given substance gives direct information
P
pv. '-("t" ) 2~It ~(t )dt (29)
regarding the absorption coefficient n. Notice that 2 [ 2(t t)2 2]
while the phenomenological description obtained in
Eq. (21) contains most of the important parameters Equation (29) is the basic "causality expression" re-
necessary for the practical use of optoacoustic lating the pressure at the observation point and the
spectroscopy for quantitative measurements of small time derivative of the velocity of expansion of the
absorption coefficients in condensed matter, it does source at retarded times.
not provide any information regarding the shape of To proceed further, we need some simple forms for
the optoacoustic output pulse. Detailed derivation ~(t'). We approximate the time dependence of the
(next section) of the acoustic pulse generation shows pulsed laser irradiation by a Gaussian shape centered
that the pressure pulse contains a positive going pulse at t'=0, with width of 27 and total energy Eo:
followed by a negative going pulse with a time separa-
tion between the two peaks which is approximately equal
to the laser pulse duration.
dE(t')
dt'
E„(, —t g,T)
(30)

where E(t') is the integrated laser energy up to time t'


and E(~) =ED. We further make an adiabatic approxi-
2. R igorous description of OA signal generation
mation (i.e. , neglect heat diffusion from the cylindrical
The problem of sound generation bya cylindrical source volume during the time of the laser pulse) and
line source has been treated in general terms by simply equate the volume expansion of the source at t'
Landau and Lifshitz (1959). Following their notation, with that due to the temperature rise at t' neglecting
we use P(r, t) to denote the potential of the small- any heat diffusion:
amplitude acoustic wave. The velocity V and pressure
P (in excess of the unperturbed value) are derivable
m[(R +~(t')]' —vR' =PnE(t')/(pC ), (31)
from the potential @ by where P is the thermal expansion coefficient, and C~
is the specific heat at constant pressure (pressure
(23)
fluctuation is always very small compared to equili-
and brium pressure in our case). We have tacitly as-
=- pay/st, sumed here that the laser pulse is long compared to the
p (24)
acoustic transit time across the cylindrical optical
where p is the density. Landau and Lifshitz (1959) beam in the liquid, so that local pressure equilibrium
derived in their equation (73.15) for cylindrical sym- is maintained. %e note that heat diffusion is always
metry negligible in our case, since diffusion length ~Q t-f
&-(&/t ) S(t')dt' in a time interval 7 is given by
@(r, t) =- [v 2(t —ti)2 r 2]14 t (25)
.„=(47D )'+,
2m &d, (32)

Rev. Mod. Phys. , Vol. 53, No. 3, July. 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 521

where D is the thermal diffusivity. Typically, we have the pressure amplitude P{r, t) decreases with distance .
v -10
~ sec and D-10 ~ cm'sec '
' for most liquids. as r ' '; this is in accordance with the conservation of
Thus ~d;;;-10 cm and is much smaller than typical energy in cylindrical symmetry, as noted by Landau
values for R, which are within the range of 10 ' to and Lifshitz (1959) [Eq. ('70. 8)j. Thirdly, we can esti-
1 cm. mate from Eq. (38) maximum pressure amplitudes
Neglecting the ~~(t'} term in Eq. {31), we have that can be expected for our typical experimental con-
2~R m(t') = pnz(t')/{pc, ) .
ditions of P-10~ K n-10 ' cm ' (or less),
',
(33) E0-10 ' J, C~-1 Jgm 'K ', v-10 sec, v, -10' 6
Combining Eqs. (30) and (33), we have cmsec ', r-1
cm. With these values substituted in
Eq. (38}, we estimate
2~R~ (t') = ' e
pC I'0-10 dyncm '
Substituting Eq. (34) into Eq. (29), we have -10 ' atm. (39)
Thus pressure fluctuations are always much less than
P(r, t) =- the equilibrium pressure (1 atm in our experiments so
far). Furthermore, we can estimate the maximum
t -(r/vg) tie-(t'/&) dtl expansion
p-1 gmcm
~'.
from Eq. (33) if we put R -10 2 cm and
2(t t )2 2]» We thus have

We may call Eq. (35} the pulsed optoacoustic equation, ~ -10 9 cm.
relating the transient acoustic pressure P at the ob-
I

This indicates that we are always dealing with small-


servation point to the retarded time development of the amplitude waves.
laser pulse. It has the form indicated in Eq. (1), with In the above derivation, we have assumed that the
the "const" containing v„r,
and an r-dependent re-
tarded time integral. Approximate expressions simi-
irradiated region is a thin cylinder, i.e. , R If «r.
this is not true, we may subdivide the irradiated
lar to Eq. (35) have also been given by other authors, cylinder of radius R into many thin cylinders, each
e.g. , Bunkin and Komissarov (1973), Lyamshev and being an acoustic source. The pressure at the ob-
Naugol'nykh (1976), Naugol'nykh (197'7), and Westervelt servation point 0 will be the algebraic sum of pres-
and Larson (1973). sures produced by these line sources. If R «v,
v, then
The pulsed optoacoustic signal generation described the double acoustic pulses from all the line sources
in Eq. (35) can be integrated numerically, and the are in phase, and we should still have a compression
predicted time dependence is yet to be compared with and rarefaction pulse at the observation point sepa-
observations in future experiments. Simplif ication of rated approximately by T. However, if R» v, v, then
Eq. (35) is possible for sufficiently large r, so that the the double pulses from various line sources are out
integrand in Eq. (35) is most important when of phase and cancellation occurs, with the results that
t —t'-r/v, . Hence we can make the following approxi- the compression and rarefaction pulses at the observa-
mation: tion point will now be separated by R/v„ the "transit
v', (t —t')' —r ' =2r[v, (t —t') —r], (36)
time" across the source. In any case, the acoustic
amplitude is still of the general form given by Eq. (1).
and Eq. (35) becomes Naugol'nykh (1977) has derived approximate solutions
for the acoustic pressure pulse generated .in a liquid
subsequent to the absorption of an optical pulse energy
and the release of the absorbed energy through non-
(37) radiative processes. His expressions are as follows.
a. For a thin illuminated cylinder, i.e. , R «v, T~.
From Eq. (34) we see that the source term ~(t') has
the largest amplitude at t'=+ 0.7T, and vanishes rapidly
at large t'I . Thus we would also expect P (r, t) to be of
I

largest magnitude at t, = rv/, + .07.7Approximate


evaluation of the integral in Eq. (36) shows that b. For a thick illuminated cylinder, i.e. , R» v, r~.
(38} (42)

Several interesting observations a, re possible with the The distinction, between the two cases depends upon the
pressure amplitudes predicted in Eq. (38). First, we relative dimensions of the radius of the illuminated
note that the compression pulse at time t is followed region [see Fig. 1 and also Eqs. (16) and (18)J and the
by a rarefaction pulse (of nearly equal magnitude) at spatial scale of the acoustical pulse. We have as-
time t, with the time interval between the compres- sumed here that the time duration of the acoustical
sion and rarefaction being approximately the laser pulse is the same as that of the optical excitation pulse,
pulse width. This prediction has been noted by several a case which is applicable when the nonradiative decay
authors (Bunkin and Komissarov, 1973; Lyamshev of the absorbed optical energy is much faster than the
and Naugol'nykh, 1976; Naugol'nykh, 1977). Secondly, duration of the optical pulse. The case (a) for R«v, r~

Rev. IVlod. Phys. , Vol. 53, No. 3, July 1981


522 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

(thin cyiinder geometry) is applicable to the long-exci- defined as the normalized optoacoustic signal S(v), is
tation-pulse, focused radiation experiments such as the obtained as a function of the laser frequency v. By
nonlinear optoacoustic experiments described later in Eq. (1),
this paper where QK 0.01 cm and 'U Tp for microsecond
S(v) =Ko. (v}, (44)
excitation is =0.2 cm. For the linear optoacoustic ab-.
sorption studies mhere the incident Ias'er radiation is with th'e constant of proportionality K dependent on
not focused we are just out of the thin cylinder ca.se (a), sensitivities of the detectors, on the properties of the
R«z, v~ since A=2 mm. With the use of dye laser sample (e.g. , P, C~, v, ) and on geometry and laser pulse
pulses of shorter duration, such as those obtained from shape.
Q-switched doubled Nd: YAG laser-pumped dye lasers, In principle, the proportionRlity constant K can be
we would always be in the thick cylinder regime (b), estimated theoretically by evaluating Eq. (38) together
i. e. , /» v, z~. It is then necessary to use the appro- with knowing the efficiencies of the detectors. In prac-
priate scaling relations involving P, v„p,
and C~ for tice, K must be measured experimentally for a given
transferring the absolute calibration from one liquid to liquid by comparing the normalized OA signals 8 and 8'
another as we see in what follows. Lyamshev and for the pure liquid and the liquid doped with a dye. -
Naugol'nykh (1976) have also derived similar expres- Several conditions must be satisfied for the dye to be
sions for spherical volumes but these do not correspond use ful:
to any of the experimental results reported here and
thus they mill not be discussed here. (1) It should be strongly absorbent in the wavelength
Gorodetskii et al. , (1978) have also carried out de- of interest so that only a very dilute doping is needed
tailed analysis of sound generation by absorption of to cause significant increase in optical absorption (di-
optical pulses in liquids. Their general results do not lute doping implies that the thermoelastic properties
differ from those of Naugol'nykh (1977) or those de- of the solution and the pure liquids are the same);
rived above. Further, they have measured the acoustic (2) No significant fluorescence of the dye should be
pulse amplitude (using a piezoelectric transducer) pro- produced~
duced in water through the absorption of a 15 nsec (3) No photochemical reaction of the doped solution
Nd: YAG laser (1.06 pm) pulse in water, which has an should occur;
absorption coefficient of -0.14 cm ' at ~ =1.06 p. m a — (4) The dye should be stable in the OA cell (for ex-
number which is very large compared to what opto- ample, it should not be absorbed by the cell mall either
acoustic spectroscopy is ideally suited for measuring. through chemical reaction or adsorption).
They also have verified (over a rather limited range With these conditions satisfied. we can apply Eq. (44)
of 7 to 17 C) the variation of the optoacoustic signal to the cases of the pure liquid and the doped liquid:
with P of water whose temperature dependence is
known. Further, for even larger absorption coefficients (46)
at 1.06 pm, obtained by the addition of copper sulfate Rnd
to water when the optical absorption occurs over small
distance, they have studied the angular distribution of S =K(DO+A~) q

the acoustic signal amplitude. where no is the absorption coefficient of the pure liquid
at the wavelength investigated, and e, is the additional
B. Acoustic pulse arrival time absorption at the same wavelength due to the dye. We
A further point not clearly seen from Eqs. (16) and can use a dual-beam spectrophotometer, and two
(18), but apparent from Eq. (38), is the time delay be- cuvettes containing the pure liquid and the doped liquid,
tween the laser pulse and the arrival of the pressure to determine e, independently. Thus we can solve K
pulse at the piezoelectric transducer placed at a dis- RS
tance r from the laser-illuminated cylinder as shown
in Fig. 1. The time delay Td is given by K =(S'- S)/a, . (47)
Ha, vingobtained A for a given experimentRI situation
(at one wavelength), we can proceed and obtain ct. (v)
from the measured S(v) using Eq. (44).
This time delay is a crucial parameter in implementa, —
tion of the optoacoustic techniques for absorption spec- D. Sensitivity
tra measurements, in that the delay allows the use of The sensitivity of the optoacoustic technique de-
proper time gating of the output signal, using a boxcar scribed above mas determined by measuring the weak
averager to discriminate against undesirable transi- sixth harmonic of vibrational absorption (Patel and
ents. Tam, 1979a; Tam, Patel, and Kerl, 1979) in C, H~
(see Sec. V. A. 2 for spectroscopy) with increasing di-
C. Calibration of OA spectroscopy lution in CCl„which has very small absorption in
Equations (16), (18), and (19) are, of course, the the wavelength region of interest. Without going
bases of pulsed optoacoustic spectroscopy: t/„„.is into detail, we have shown (Tam, Patel, and Kerl,
measured by a piezoelectric transducer in direct con- 1979) that absorption coefficients as small as
tact with the sample, and Eo is measured by an optical -10 ' —10~ cm ' can be measured (with a signal-to-
detector (generally a, pyroelectric detector which has noise ratio of 1, input pulse of energy of —1 mJ/pulse,
wavelength-independent sensitivity}; the ratio V, /Eo, , pulse repetition frequency of 10 sec ', and integration

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

time of 1 sec). Bebchuk et a/. (1978) were one of the first to investi-
The factors affecting and limiting the sensitivity of gate the .interference effects of electrostriction in
the measurement technique include (a) optical absorp- optoacoustic spectroscopy using pulsed lasers and
tion signals from windows, (b) light scattering in the immersed (or attached) transducers. However, their
bulk of the liquid, and (c) electrostriction. The signals observations that an optical absorption coef f ic ient of
arising from the first two can be minimized by (1} -3 cm ' gave optoacoustic signals only a factor of ten
choice of low-loss windows, (2) reduction of scattering larger than those arising from electrostriction surely
impurities in the liquid, (3) appropriate time gating of is in error, since our experimental spectroscopic data
the acoustic signals observed from the piezoelectric show that electrostriction plays no role in the opto-
trarisducer. The time gating is very important, acoustic signal for absorption coefficients as small as
since scattered light travels at the velocity of light in 10 'cm'.
the medium, v, =c/n, where n is the refractive index
of the material, while the acoustic signal generated in III. APPARATUS
the bulk will travel at the acoustic velocity v, in the
In this section we present. a detailed description of
medium. Thus the signals arising from the scattered
our experimental apparatus for pulsed OA spectroscopy
light being absorbed by the transducer will occur es-
of condensed matter. Most of our discussions are con-
sentially promptly after the laser pulse, while the bulk
acoustic signal is delayed by 7„defined in Eq. (43). cerned with liquids. The case of solids (mainly
For the case of acoustic signals from optical absorp- powders) is discussed in Sec. III.D.
tion by the windows, the acoustic pulse arrives after
that originating from the bulk. The acoustic pulse A. Transducer
generated due to electrostriction processes in the In recent years, several authors have realized that
laser-irradiated region is, however, of similar form the photoacoustic spectroscopy scheme {Harshbarger
to that arising from the bulk absorption, and hence and Robin, 1973; Roqencwaig, 1973; Rosencwaig and
there is no reasonable way in which the two can be Gersho, 1976; Rosencwaig, 1977; Bosencwaig, 1978;
distinguished. The electrostriction signal pulse is Rosencwaig e/ g/. , 1979) first used by Harshbarger and
estimated (Brueck et a/. 1980) to be of comparable size Robin (1973), and by Rosencwaig (1973), utilizing gas-
to that arising from optical absorption for ~ = 10 6 cm '. phase microphones, is quite insensitive for QA studies
Brueck et ai. studies, however, were biased to show o f condensed matter. Thus the technique o f piezoelec-
the superiority of the photorefractive technique over the tric detection of the OA signal has been used by I ah-
pulsed optoacoustic technique. This can be clearly seen mann e/ ~/ (1977), .Bonch-Bruevich e/ a/. (1977), Sladky
from their use of short laser pulse duration of -70 nsec e/ a/. (1977), Farrow et a/. (19'/8), Burt (1979), and
(giving rise to high peak powers and the attendant high others. 'The reason is that acoustic impedance matching
optical frequency electric fields which determine the is very good for piezoelectric detectors directly coupled
electrostriction effects), and tight focusing of the inci- to solids, and acoustic transmission exceeding 50% from
dent laser beam (200 p. m diameter), even for linear ab- from the solid to the detector is generally possible. For
sorption studies. The latter gives rise to high optical liquids, the transmission generally exceeds 10%%uo. This
intensities and correspondingly high electric fields to is in sharp contrast to the transmission across an
accentuate the effects of electrostriction. It can be interface between condensed matter and gas, where
shown that the absorption coefficient at which the opto- the transmission is typically -10 '. (Some of the
acoustic signal and the acoustic signal due to the elec- coupling inefficiency is made up by the very high sensi-
trostrictive effect become equal is inversely propor- tivities of commercially available gas microphones. )
tional to the diameter of the laser spot size. Thus, for Various piezoelectric transducers have been used.
typical linear optoacoustic spectroscopy studies where A geometry that seems to be most popular is a piezo-
the laser spot size is rarely smaller than -3 mm, the electric cylinder that actually encloses the liquids
acoustic signal arising from the electrostrictive effect (Lahmann e/ a/ , 197'7; Slad. ky et a/. , 19'/7; Farrow,
would equal that arising from optoacoustic effect for 19'18; Oda et a/. , 1978; Burt, 1979). However, we
optical absorption coefficient ~, = Gx10 ' cm ' even found that it is easier and more reliable to use a home-
keeping the laser pulse length fixed. Further, if we made transducer with a flat face that protrudes into
extrapolate to the case of long laser pulse lengths of the liquid being studied. The construction and the ad-
1 p. sec used in most of the present studies, the optical vantages of such a transducer are discussed below.
absorption coefficient cy, becomes -4x10 ' cm ', a value Figure 2 shows the construction of our piezoelectric
which is much smaller tha. n that estimated by Brueck transducer (Tarp and Patel, 19'79b). The heart of the
Finally, there is no (or only very weak) depen- transducer is a lead zirconate-lead titanate (PZT}
dence of the electrostriction pulse on laser wavelength. cylinder that. is poled axially and silvered on both ends.
Hence, even the 4x10 cm level does not represent a The cylinder, of diameter 4 mm and height 4 mm, is
limitation on the smallest absorption coefficient that purchased from Transducer Products of Connecticut
can be measured using the optoacoustic spectroscopy if (material LTZ-2). This material is commonly used in
the material being studied has a wavelength dependent hydrophones because of its high sensitivity, high long-
absorption. Subtraction of background signal (constant) term stability, and resistivity at elevated temperatures.
as large as a factor of 10 to 100 times that correspond- It is nearly identical to the more commonly known
ing to a well-defined peak can easily be subtracted elec- material PET-5A produced by Vernitron of Ohio. The
tronically. height of the cylinder is chosen to provide a rise time

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Putsed optoacoustic spectroscopy
t

POLISHED the PZT material are nearly the. same, so that acous-
P2T CYL INOER SURFACES tic loss due to the presence of the front diaphragm is

THIN
:rX rrllllJI TEFLON
small.
We note that most commercial hydrophones have a
LAYER
RING front seal made of plastics (e.g. , neoprene or epoxies),
OF
GREASE and thereby lack the above advantages. Furthermore,
—LEAD we note that if we had used the more commonly prac-
ticed method (Farrow et al, , 1978; Hordvik and
COPPER Schlossberg, 1977; Oda et a/. , 1978) of directly ex-
posing the PZT cylinder to the liquid instead of coupling
through a stainless steel diaphragm, we would have
STAt NLESS SPRING
STEEL lost the advantages accrued from points (a)-(d) above.
GAS IN 6 We have chosen the PZT material with the lowest
~ I

i
W
mechanical Q for our transducer, to minimize mechani-
cal ringing of the transducer. Our PZ T mater ial has a
Q that is an order of magnitude smaller than some
ICRODGT" other PZT materials. However, a significant degree
CONNECTOR of ringing still occurs. This does not constitute a major
1/ problem for our OA studies so far, but if the true time
TG development of the OA signal is to be studied, then the
PREAMPL I F I E R
degree of transducer ringing needs to be further re-
FIG. 2. Cross section of the homemade piezoelectric trans- duced. In these cases, the use of suitable piezoelectric
ducer utilizing a lead-zirconate-titanate cylinder (LTZ-2 or
P ZT-5A).
films (Wickramasinghe et al. , 1978)„like ZnO or poly-
vinylidene fluoride „may be necessary.
B. Transducer preamplif ier
-1 @sec, which is fast enough to detect the OA ultra-
sonic signal generated in our experiments. Since the OA signal has an amplitude usually ~10 ' V,
The PZT cylinder is firmly mounted in a stainless suitable low-noise preamplification is necessary to in-
steel casing, as illustrated in Fig. 2. The front crease the signal amplitude sufficiently for boxcar
diaphragm of the casing (about 1 mm in thickness) is detection. This can best be done by using a battery-
polished on both sides. The PZT cylinder is spring- powered preamplifier located near the transducer.
loaded against the diaphragm, with a thin layer of sili- Since substantial overall amplification must be used,
cone grease applied between them to ensure good acous- the amount of electrical pickup must be minimized.
tic coupling. The backing plate of the PZT cylinder is As mentioned earlier, the purpose of mounting the PZT
a lead disc soldered to a copper disc. The use of lead cylinder inside the stainless steel casing is to mini-
(which has high ultrasonic attenuation) leads to reduced mize pickup. Various preamplifiers have been used by
acoustic reflection back into the PZT transducer, thus us, and we find eommercia1. preamplifiers like Ithaco
reducing "ringing" effects. A thin layer of grease is Model 143F to be satisfactory. However, lower noise
also applied between the PZT cylinder and the backing is obtained with a homemade preamplifier with about
plate. The OA signal is fed through the phosphor bronze 40 dB gain, designed by G. I . Miller and implemented
spring to the microdot connector, as shown in Fig. 2. by H. J. Kerl. This preamplifier is used in most of
For our purposes, this transducer has a performance our studies„and the circuit for it is given in Fig. 3.
comparable to or better thanthat of commerical hydro- Here Ql is a low-noise junction field-effect transistor
phones (which are available from numerous companies,
like Celesc, o of California or KSP Industries of Vir- R4 +12 V
ginia). However, besides good sensitivities L- 3x 10 5 J~ 100 ++ 40ys
V/(Nm 2)], our transducer has several important Cz ~40pa 1.8 K
R3 (metat)
C7
OUTPUT (1)
advantages over comme rical hydrophones: C6
I
C'
R5

0.01' 4.7 K
100
(a) The flat polished front face of the transducer NPU T
R2

reflects most of the stray light scattered onto it; fur- C12
I

thermore, the large thermal mass and conductivity R 2N6550

of the diaphragm minimize the effect of light absorp- 22M


(mica) Cs
J.
~&40'.
C7
Ir
OUTPUT(2)

tion; 0.01@,
(b) Nucleation of microbubbles is unlikely on the Re 200
6 3.3K
polished stainless steel face of the transducer;
0. tp.
(c) There is less likelihood that the liquid sample
being studied will be contaminated by the stainless R7 —12 V
steel transducer than by other materials (like silver, C4 &~4OI 100
c)) +sop.
T+
aluminum, or epoxies);
(d) The enclosing of the PZT cylinder in a metal FIG. 3. Suitable preamplifier circuit (gain -40 dB) for the
casing minimizes electrical pickup; transducer (modified from a circuit designed by G. L. Miller
(e) The acoustic impedances of stainless steel and
. of Bell Laboratories).

Rev. Mod. Phys. , Vol. 53, No. 3, juty 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 525

(PET) (Teledyne-Crystalonics type 2N6550} having a


series noise spectral density of -1 nV Hz '+, corres-—
ponding to a series noise resistance R, of -50 G. The
system parallel noise resistance R~ is dominated by
the equivalent loss resistance of the piezoelectric
transducer employed, and is -3.5 kO at 1 MHz.
The total input capacitance C~, comprising the trans-
ducer together with the stray capacitances and the pre-
amplifier input capacitance, is -140 pF. This value,
together with Q~ and R„yields a noise corner time
constant
V, =C, (R,fl, } ~. (48}
T, in our case is evaluated to be 0.06 @sec. The pre-
amplifier low-frequency rolloff is set by R3C3, while
the high-frequency limit is set by A1 together with its
(trimmed) stability capacitor C4. Two outputs are pro-
vided, a wideband one via R8 and an optional 2 p sec
clipped version set by C7A9.
The OA signal is further amplified by an amplifier
(Princeton Applied Research Model 113) before being
fed to a boxcar integrator.
C. OA cell
A sketch and a photograph of our OA cell for liquids
is shown in Figs. 4 and 5. The cell body is bored out
of a cube of stainless steel, 2.5 cm on each side. The
bore is 1.3 cm in diameter, and the surface of the
bore is electropolished. High polish results in good
ref lectivity for the scattered light, so that spurious FIG. 5. Photograph of the stainless steel OA cell. with a flange
signals due to absorption of stray light at the cell containing a quartz window removed. The polished transducer
walls can be minimized. As can be seen from Fig. 4, front diaphragm protruding into the cell can be clearly seen.
the front surface of the transducer extends into the
bore of the OA cell. A pipe attached to the side of the at the cell wall, the heat generated is distributed over
cell allows direct distillation of the liquid into the cell. a diffusion depth &~E«as given by Eq. (32). This is as-
The cell is equipped with two fused quartz windows, suming that &d'ff'( 10 ' cm for our cases) is much
which are seated in stainless steel flanges as shown; greater than the optical absorption length (-10 ' cm
teflon 0-rings are used between the quartz window for metals). The heat distributed over the depth &„.
and the end surface of the cell. To minimize possible
contamination of the liquid sample being studied, the
causes a temperature rise ~T given by
const/(
„, «

only materials exposed to the liquid are stainless steel, ~~II diff p A) (49)
quartz, or teflon. Figure 5 shows a photograph of the where p and C~ are the density and specific heat of the
stainless steel optoacoustic cell. metal. The "const" in Eq. (49) contains surface optical-
We have chosen stainless steel as the main con- absorption and reflection factors. We see from Eq.
struction material of the OA cell. Since stray light (49) that construction material for the OA cell should
absorbed at cell walls constitutes a possible source of have small surface optical absorption, large thermal
la. rge spurious signals, the choice of material needs to diffusion length, high density, and specific heat. In
be examined carefully. When stray light is absorbed this respect, silver or aluminum are a better choice
than stainless steel; the factor 1/(Xd, ff pCA) is smaller
EXPANS ION
CHROMEL- ALUMEL in silver or aluminum than in stainless steel by
q~THERMOCOUPLE
TUBE
(l. roughly a factor of 2. However, the high degree of
FLON 0- RING
chemical reactivity of silver or aluminum makes the
QUARTZ
P~)
FLANGE
choice. of stainless steel a reasonable compromise.
L
'WINDOW

D. Arrangement for films or powders


r l LASER

~
~ TEFLON
—y//WATER
BEAM

DISTILLED
We have recently developed (Patel and Tam, 1980;
Tam and Patel, 1979c) a modified form of the OA cell
~ STAINLESS STEEL
PZT CYLINDER
//g"' ~ for spectral analysis of thin liquid films or for films of
powders. This modified arrangement is shown in Fig.
TRANSDUCER
1 cm 6. The film sample to be studied (typically several,
microns in thickness) is clamped between fused quartz
TO PREAMPLIFIER
plates, and the OA signal generated in the liquid or
FIG. 4. Schematic drawing of stainless steel optoacoustic cell. solid film is launched as acoustic waves (both trans-

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


526 C. K. N. Patel and A. C. Tarn: Pulsed optoacoustic spectroscopy

THIN LAYER OF GREASE


'I tering onto the transducer occur immediately, while
CLAMP the desired OA signal begins only after an acoustic
TRAN SPRING
propagation time. The time-gating technique is dis-
cussed further in Sec. IV.
The geometry of the substrate (a right-angle bend, of
width -1 cm and length of each arm -10 cm) has a
TIC WAVE
waveguiding effect for the OA waves, and very efficient
RED LIGHT
detection of the OA signal can be achieved. The detec-
TO PREANPLIFI
tion is done with a PZT transducer identical to that used
POWDER BETW
for liquids (Fig. 2). The flat polished front diaphragm
QUARTZ PLAT of the transducer is now firmly spring-loaded against
-~FUSED QUARTZ the quartz substra, te at the position indicated in Fig. 6.
PLATES
A thin layer of silicone grease is used between the
transducer and the substrate to facilitate acoustic
FIG. 6. Suitable OA "cell" for films, powders, or any thin coupling.
layer of highly light-scattering material.
IV. EXPER IIVIENTAL PROCEDURE
The block diagram for the experimental investigation
verse and longitudinal, and possibly also surface acous-
of weak absorption in liquids is shown in Fig. 7. The
tic waves) into the substrate. Quartz has a very small
apparatus is very simple, yet it can presently detect
absorption coeff ic ient e for ultrasonic waves
absorption coefficients as small as 10 ' cm '. This
(n -10 ~ cm ' at a frequency of 1 MHz), and so little represents an improvement in detectability by a few
absorption occurs between the point of generation of the orders of magnitude over that previously achieved with
OA signal (i.e. , irradiated position of the film) and
conventional OA measurements (Harshbarger and Robin,
the point of detection (i.e., the transducer) Alth. ough
1973; Rosencwaig, 1973; Rosencwaig and Gersho, 1976;
quartz is an excellent substrate material for the pres-
Rosencwaig, 19VV; Rosencwaig, 19V8). Further im-
ent purpose of OA studies of films, common micro-
provement in detectability is possible by using, for ex-
scope slides are found to be much less useful, resulting
ample, more powerful and stable La, ser sources, better
in an OA signal an order of magnitude smaller than in
shielding against acoustic noise and electrical pickup,
the quartz case. The reasons for this big difference
and a better cell design (e.g. , multipass geometry,
are unclear. windows farther away from the transducer, antireflec-
The apparatus shown in Fig. 6 is also found to be
tion-coated windows with as smalL optical absorption as
ideally suited (Tam and Patel, 19V9c) for OA spectro-
possible, etc. ).
scopy of powders, which may be cooled to low tem'- The light source used in our studies so fa, r is a
perature by refrigerating the part of the substrate con- commercial scannable flash-lamp-pumped dye laser,
taining the powder. %e first make a suspension of the producing -1 mJ energy and 1 @sec duration pulses,
powder in a transparent viscous carrier liquid (e.g. , at a laser bandwidth (full width at ha. lf maximum
ethylene glycol). Then we place a small drop of the FWHM) of -2 cm '. With the use of various dyes, we
suspension on a quartz substrate, and cover it with a can cover the visible region from -450 nm to -700
quartz cover plate. Intense light scattering is ob- nm. It is usually operated at 10 pulses/sec. The band-
served, and this would generally constitute a major width of 2 cm ' is adequate for most liquids and solids;
indeed, a bandwidth -100 cm ' is frequently adequate
source of spurious signals in an OA study. However,
the disastrous effects of intense light scattering are
for resolving spectral features of condensed matter,
nearly eliminated in our experimental arrangement,
shown in Fig. 6, for the following reasons: (a) the
quartz plates sandwiching the powder film (approxi- LIQUID OA CELL
mately several microns thick) are approximately 1 mm PYROELECTRIC
PULSED DETECTOR
in thickness; this thin sandwich geometry allows most
~DYE LASER
of the scattered Light to escape a, nd a very small
amount of the scattered light is trapped in the sub- TRANSOUCER
strate; (b) as can be seen in Fig. 6, the substrate has SCA NNING
DRI V E AMPLIFIER
a right-angle bend, and hence most of the light
trapped in the substrate cannot reach the transducer.
Further reduction in the amount of scattered light I BOXCAR A I BOXCAR B I
X-Y
reaching the transducer can be achieved by having RECORDER
B
more right-angle bends in the substrate. On the other A/B
RATIOME TER
hand, acoustic waves launched into the substrate can
readily diffract around the bends and reach the trans-
ducer; and (c) the effect of the small amount of light
~ OA SPFCTRUM

that actually reaches the transducer indirectly can be


minimized by a time gating technique. This is possible FIG. 7. Block diagram of the simple experimental arrange-
because of the much faster speed of light than the speed ment needed for linear spectroscopy of liquids with the pul. sed
of sound, and so the spurious signals due to light scat- QA detection method.

Rev. Mod. Phys. , Vol. 53, No. 3, July 198l


C. K. N. Patel and A. G. Tam: Pulsed optoacoustic spectroscopy

and so pulsed arc light sources together with a scan- dependent on the distance d between the laser beam
- nable low-resolution monochromator can be used in- and the transducer surface. By moving the cell with a
stead of the laser. However, in certain studies (e.g. , micrometer translation stage, d can be varied by sever-
investigation of rare-earth oxide powders at low tem-- al mm, and t is observed to change by several gsec.
perature), resolution of 0.1 cm ' or better may be The dependence of the change in delay At on the change
needed, and narrow-band laser sources must be used. in distance Ad is shown in Fig. 9, where the reciprocal
As shown in Fig. '?, the transient OA signal is ampli- of the gradient is found to be (1.32 + 0.04) x 10' cm sec '
fied and measured by a boxcar integrator A. . The for the case of a benzene-filled QA, cell at 20 C.
laser pulse energy is monitored by a pyroelectric This is in excellent agreement with the established
detector, whose output signal is amplified and mea- ultrasonic velocity at room temperature in benzene,
sured by a boxcar integrator &. The pyroelectric being 1.326&10' cm sec '. Thus we are detecting a
detector presently used is a coated lithium niobate ballistically-propagating OA signal f rom the laser-
detector (Laser Precision Model 2050S), which has a irradiated column of liquid. Our boxcar gate (shown
flat spectral response (flat to within +2%) from near in Fig. 8) is set to detect the amplitude of the early
uv to near ir, and is sufficient for our studies so far. part of the OA signal. This is because later parts are
However, when higher accuracies are needed over an found to be more liable to be contaminated by spurious
extended spectral range, better pyroelectric detection absorptions, e.g. , at the cell windows. This is es-
(e.g. , models with multiple reflections) with improved pecially important when very weak absorption features
spectral flatness will be required. The boxcar outputs of the liquid are being studied. The above technique of
A and B are fed to a ratiometer and the ratio A/B time gating is also very useful for the case of powdered
(def ined as the normalized OA signal S) is obtained as samples, where intense light scattering occurs. Fig-
the laser frequency v is scanned. (This normalization ure 10 shows an observed OA signal for fine powders
procedure is for linear absorption; slightly different of Ho,O„ in the experimental arrangement shown in
normalization procedure is needed for nonlinear, ab- Fig. 6. There is a very important difference between
sorption spectroscopy, as discussed later. ) Subsequent the OA signals detected when the laser is tuned on
to our initial studies where we used the ratiometer for resonance or off resonance of an absorption feature of
obtaining A/B, we have replaced the ratiometer by a Ho, O, . For the off-resonance case, signals are ob-
minicomputer which digitizes the signals A and B from served almost immediately after the laser pulse, be-
the boxcars and also keeps track of dye laser frequency cause of some residual amount of scattered light that
through an appropriate readout unit. The output, then, reaches the transducer. This unwanted signal decays,
is the A/B value as. a function of frequency, which is with a time constant of some tens of p. sec. For the
plotted and printed as necessary. The advantage of us- on-resonance case, the same unwanted signal ap-
ing the minicomputer system (a system using a rela- pears; however, after an acoustic propagation time„
tively inexpensive microprocessor can be used instead an extra signal appears due to the OA wave launched
of the minicomputer) is the increased dynamic range in into the substrate by the powder. Hence by gating the
the ratio determination and the digital information, OA signal at an appropriate delay (as shown in Fig. 10),
which makes it easy to join up sections of S(v) vs v the masking effect of the unwanted signal caused by
data obtained by using different dyes for different spec- light scattering can be strongly suppressed. Further
tral regions. We have seen in Eq. (44) that S(v) is improvements in rejection of scattered-light-induced
proportional to o. (v); with the proportionality constant signals has been accomplished by having multiple bends
determined by Eq. (47), an absolute absorption spec-
trum is obtained.
We now describe some experimental observations that
support the theories of pulsed QA signal generation; in
particular, we want to verify the pulsed OA Eq. (1). An
OA signal detected by the piezoelectric transducer is
shown in Fig. 8. The signal begins at a well-defined
delay time t after the occurrence of the peak of the
laser pulse. This delay time t is found to be linearly

BOXCAR
GATE ~
OA
SIGNAL~

LASER
PULSE~ &d (mm)
I 5p. sec FIG. 9. Dependence of the change in delay tixne (Dt) on a
FIG. 8. Oscilloscope picture shows the laser pulse (lower change in the beam-transducer separation (~d). The line
trace), OA signal (middle trace), and boxcar gate for the OA through the data point corresponds to the ultrasonic velocity
detection (upper trace) for a typical liquid. in the liquid.

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tarn: Pulsed optoacoustic spectroscopy

in the quartz substrate. This is indicated in Fig. 11, radiation. Comparison of Figs. 11(B)and 11(D) indi-
where the scope pictures (A) and (B) correspond to a cates that the scattered-light-induced signal is now
substrate with a single bend, and (C), (D), and (E) reduced by at least an additional factor of 5. That the
correspond to a substrate with multiple bends, as actual reduction is more than this factor can be seen
shown in the corresponding drawings on the right. For by comparing Fig. 11(D) with Fig. 11(E), which shows
Fig. 11, the powder sample being examined is Ho, 03,. the signal from the transducer with the laser blocked.
the dye laser is on an absorption line at 18 735 cm ' The lack of any significant differences between Figs.
(line No. 8 in Fig. 24) for pictures (A) and (C), but for 11(D) and 11(E) is a convincing demonstration that the
pictures (B) and (D), the dye laser is detuned away multiple-bend arrangement shown in Fig. 11 completely
from an absorption line, i.e. , the dye laser frequency eliminates the scattered-light problem associated with
is set at 18800 cm ' where there is little or no ab- the optoacoustic study of powder samples.
sorption by the Ho20, powder. In picture (E), the dye The linearity of the observed OA amplitude versus
laser beam is blocked. In Fig. 11(A), the laser pulse the laser pulse energy has been verified for linear
is shown on the lower trace. The delay of -25 @sec absorption in various liquids (e.g. , benzene, acetone,
seen between the laser pulse and the onset of the opto- water, etc. ) at various absorption levels, from absorp-
acoustic signal is understood as the propagation time tions coefficients larger than 10 ' cm ' to those smaller
for the acoustic pulse to travel from, the illuminated than 10 ' cm '. This linearity is in accordance with Eq.
region to the transducer. In Fig. 11(B), the OA signal (1). An example of our linearity study is shown in Fig.
is much reduced; however„ the small signal that we 12, which shows a plot of the amplitude of the OAsignal
see arises primarily from the scattered light reaching measured by the transducer at various laser pulse
the transducer, suggesting incomplete elimination of energies, which are varied by calibrated neutral-
the scattered-light problem. Figure 11(C) shows, density filters. The data of Fig. 12 correspond to
again, the OA signal for the on-resonance case for the, liquid benzene, with the laser tuned to 18 810 cm '
new geometry. Two points should be noted. First, the (peak of the seventh harmonic of the C-H stretch in
start of the signal trace is delayed from the laser pulse benzene). Good linearity is clearly observed.
by an additional 50 psec. Thus the total acoustic delay The dependence of the OA s ignal on the the rmal ex-
is 75 p. sec, which is consistent with the acoustical propa- pansion coefficient P, as given in Eq. (1), has also been
gation time in the quartz substrate in which the acous- verified experimentally. This is found to be most con-
tic pulse is launched from the absorption of laser light
.
in Ho, O, The second point to notice is that the ampli-
tude of the acoustic pulse is identical with that in Fig.
='Illi'I lilll 'IlfiIi'I II
11(A) for a shorter acoustic propagation distance,
indicating that for the distances involved here, there
Iiiiilll&IIIU& ! IIII

is little or no acoustic absorption for the frequencies


involved. Figure 11(D) shows the dramatic improve-
ment that is obtained for the off-resonance laser

(a)

e~ ~gmsi
%III
ASSR

-'liiii'IIlgII.
(b)

5pSeC
FIG. 11. Comparisons of two geometries for optoacoustic
FIG. 10. Each scope picture shows the laser pulse (tower -
spectroscopy of powders. (A) and (C) show optoacoustic
trace), transducer signal (middle trace), and gate position of signals when the dye laser is tuned to the peak of an absorption
boxcar A (upper trace). Time scale = 5 p, sec cm ~. Powder line (no. 7) of Ho&03 and (B) and (0) show corresponding signals
sample is. Ho&03. (a) Laser tuned to the absorption. line at when the dye laser is tuned away from the absorption peak. (E)
18633 cm ~. (b) Laser tuned off an absorption line (laser fre- shows the optoacoustic transducer output when. the laser beam
quency at 18800 cm '). is blocked.

Rev. Mod. Phys. , Vol. 53, 'No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 529

10 after a delay time due to acoustic propagation. We see


from Fig. 13 that the OA signal changes sign at 4 C;
i.e., the signal (at early time) reverses phase as we
follow it from above 4 'C to below 4 'C. (Later signals
CO
I- are contaminated by window absorptions and no clear
phase reversal is seen). We have plotted the ratio of
the OA signal to the known thermal expansion coefficient
P of water in Fig. 14. We find that this ratio is a con-
stant, independent of temperature, in agreement with
Eq. (1), and implying that the other parameters such
as the absorption coefficient e are, at most, very
slowly changing with temperature.
0. 5
V. EXPER IIVIENTAL RESULTS
FIG. 12. Verification of the linearity of the QA signal ampli- In this section we review the results in linear and
tude H for various laser pulse energies (changed by the use of
neutral. density filters). This example is for benzene at a laser nonlinear spectroscopy that we have obtained so far
frequency of 18810 cm ' (peak of the seventh harmonic of the with the pulsed OA detection technique described above.
C— H stretch). The purpose of this section is to exemplify how quan-
titative linear or nonlinear spectroscopy can be per-
veniently dane with liquid H, O, which has a very rapidly formed in cases where new physical information has
changing P with temperature; indeed, P changes sign been obtained. We note that small linear absorptions
at 4 C. Our observed OA signals at various tempera- in liquids have been studied by ather methods, e.g. ,
tures are shown in Fig. 13. At 4. C, a very weak OA by conventional long-path-length absorption spectro-
signal is observed, as seen in Fig. 13(E), correspond- scopy (Stone, 1978), by wavelength modulation
ing to weak absorptions at cell windows and noise due spectroscopy (Moses and Tang, 19'7'1), by laser calori-
to electrical pickup. At lower or higher temperatures, metry (Callis, 1976), by Michelson-type double-arm
OA signals due to water absorption are clearly seen, absorption spectroscopy (Querry et al. , 1978), by

,
2j. , OC

FIG. 13. Transducer output signal for water at various temperatures. In each of the scope pictures, the upper trace is the trans-
ducer output signal for liquid H20 at the indicated temperature (in' C); the lower trace is the laser pulse detected by the pyroelec-
tric detector. The laser frequency is fixed at 16400 cd . Horizontal scale is 5 p, sec cm . Vertical sensitivities are the same
for all the pictures except (A) and (B), where the normal sensitivity is decreased by a factor of 2.5, as indicated. Note the phase
reversal of the early part of the transducer output at 4 C.

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


530 G. K. N. Patel and A. C. Tam: Pulsed optoacoustIc spectroscopy

16 as a factor of 10. Our OA results in water seem to be


the most accurate -up to date.
14— o/'0
A. Linear spectroscopy
0 't. Water

(H/p) Liquid water is a very transparent material in the


10— 0 visible region. Because of the fundamental and practical
/6 importance of water, many researchers have attempted
V)
o to measure its optical absorption over the past century.
z 8— /
However, this has remained a very challenging task,
o and results from different measurements typically dis-
H
agree by factors of -2 or more. Thus we see that all the
0 conventional ways of measuring weak absorption, such
P as the optical absorption of water in the visible region,
o/ suffer from several difficulties: (a) If the extinction co-
z
/
0 efficient is being measured instead of absorption (as is
2
~O done in most investigations), then the amount of scat-
o/
tering loss must be known. This is quite difficult ingen-
eral, because liquid samples frequently contain micro-
I

0 /
A

/ scopic particles that cause strong light scattering. (b)


0/ Absorption and reflection losses at windows are fre-
2~

/
() quently hard to account for accurately. (c) Insufficient
collimation, lateral displacements, reflections at cell
10 20 50. walls, etc. , of the light beam are frequent sources of
T (c) errors. These problems are especially severe when
.long-path-length absorption cells are used.
FIG. 14. Measured optoacoustic signal and the ratio of mea-
sured OA signal to known thermal expansion coefficient for
Our pulsed OA technique specifically eliminates the
water as a function of temperature. above problems. Also, our choice of highly. corrosion-
resistant material for construction of our OA cell
thermal-lensing spectroscopy (Long et al. , 1976), and results in no contamination problems. This is verified
so on. References to these other methods of investi- from the fact that the same OA spectrum is obtained
gation have been made in our recent publication (Patel for a freshly distilled water sample as for a sample
and Tam, 1979a) and these other methods will not be stored in the OA cell for several weeks. Furthermore,
discussed here. Io any case, we believe that the pulsed the high sensitivity obtainable with the pulsed OA tech-
OA technique, in general, produces more accurate nique enables us to obtain the first unambiguously ac-
and/or reliable results compared to the other methods curate optical spectra of liquid H, O and D, O.
of investigation. For example, in the case of the op- Figures 15 and 16 show our observed OA absorption
tical absorption spectrum of liquid H, O, other methods spectra of light and heavy water, respectively, in the
have yielded absorption coefficients (in the green visible region. Detailed tabulation of the absorption
spectral region) with mutual disagreements by as much coefficients is given in Table I (Tam and Patel, 1979b).

4000 4250 4500 4750 5000 5500 6000 6500 7000


10 ~— I I I I I

5— o HULBURT 1945 v HALE 8 QUERRY 1973 d QUERR


o SULLIVAN 1965 ~ KOPELEVICH 1976 — PRESEN
+ IRVINE 8 POLLACK 1968 HAS S 8 DAY I S SON 1 977

HpO

dd dddddddd~ 0
p
P Opd d @pd d d d d d v 0
0 0 o v opo p
0 oo oo

I I I I I
I

24 20 18 22 16
v (1000 cm )
FIG. 15. Absorption spectra of distilled light water by optoacoustic determinations (shown by full line). The dashed parts are in-
terpolations not covered by the tuning range of the laser dyes used. Also shown are some experimental data from other authors
for comparisons. (See Hulburt, 1945; Sullivan, 1963; Irvine and Pollack, 1968; Hale 'and equerry, 1973; Kopelevich, 1976; Hass
and Davisson, 1977; equerry et al. , 1978, for descriptions of their work. )

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

X (A)
4000 4250 4500 4750 5000 5500 6000 6500 7000
10-2 I I I I I

o SULLIVAN 1963 PRESENT WGRK

aoo aoaoaa
ooa a oooo D20
aoo aao
aaa o
Oa OOoo a
aoa aaoaa
aoo

oo o oao o a oo
0 oo oo o.
a a ooaooo

I I I

24 22 20 18
v (1000 cm t)

FIG. 16. Absorption spectra of distilled heavy water by optoacoustic determinations. Notations same as in Fig. 12.

TABLE I. Absorption coefficients of liquid water at 21.5'C by laser optoacoustic spectroscopy.

~ (H, o) ~ (D,o) + (H2o) ~ (D,o)


v (cm-') ~ (nm) (10 cm ) (10 cm ) v (cm ) Z (nm) (10 'cm ') (10-' cm ')

22 400 446. 3 2.38 9.50 18 700 534.7 4.39 2. 40


22 300 448. 3 2.33 8.68 18 600 537.5 4.53 2.48
22 200 450. 4 2.30 8.11 18 500 540.4 4.80 2.49
22 100 452. 4 2. 27 7.69 18 400, 543. 5 5.00 2. 50
22 000 454. 4 2.21 7..27 18 300 546. 3 5.28 2. 50
21 900 456. 5 2. 14 6.96 17 400 574.5 8.05 2. 22
21 800 458.6 2. 13 6.68 17 300 577. 8 8.99 2. 10
21 700 460.7 2. 11 6.46 17 200 581.2 1Q.07 2.Ql
21 600 462. 9 2.06 6. 17 17 100 584.6 11.07 1.91
21 500 465. 0 2.06 6.03 17 000 588.0 12.32 1.89
21 400 467.2 2.00 5.77 16 900 591.5 13.77 1.86
21 300 469.4 2.05 5.60 16 800 595.0 15.89 1.86
21 2'00 471.6 1.84 5.41 16 700 598.6 19.25 1.88
21 100 473.8 1.77 5. 20 16 600 602. 2 23.2 1.92
21 000 473. 1 1.79 4.74 16 500 605.9 24. 8 1.97
20 900 478. 4 1.86 4. 51 16 400 609.6 25.7 2. 00
20 800 480. 7 1.86 4.39 16 300 613.3 26.7 2.03
20 700 483.0 1.84 4.09 16 200 617.1 27.7 2. 09
20 600 485.3 1.81 3.83 16 100 620. 9 28. 8 2. 22
20 500 487.7 1.86 3.70 16 000 624. 8 29.6 2.28
20 400 490.1 1.89 3.51 15 900 '
628.7 30.0 2. 28
20 300 492.5 1.93 3.39 15 800 632.7 30.4 2.2Q
20 200 494. 9 2.00 322 15 700 636.7 31.0 2. 15
20 100 497.4 2. 16 3.19 15 600 640. 8 31.Q 2. 15
20 000 499.9 2.33 3.10 15 500 645.0 31.4 2. 14
19 900 502.4 2.50 3.05 15 400 649.2 32.3 2.06
19 800 605.0 2.70 2.98 15 300 653.4 33.1 2. 12
19 700 507.5 2. 88 2. 89 15 200 657.7 36.1 2. 12
19 600 510.1 3 13
~ 2.72 15 100 662. 1 37.9 2. 15
19 500 512.7 3.48 2.54 15 000 666.5 38.7 2. 26
19 400 515.4 3.65 2.50 14 900 670.9 39.5 2.40
19 300 518.0 3.75 2.55 14 800 675.5 40 2.54
19 200 520. 7 3.80 2.46 14 700 680. 1 42. 6 2. 69
19 100 523.5 '3. 87 2.48 14 600 684. 1 45. 1 2.74
19 000 526. 2 3.97 2.48 14 500 689.2 47. 6 2. 90
18 900 529.0 4.09 2.48 14 400 694.2 52. 6 3.20
18 800 531.8 4.27 2.41

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


532 C. K. N Pate~ and A. C Taam pulsed o p oacoustIc spect roscQpy

Several oth er meassure ments of the Rbsol"ptlon co- 10—


ient o f mate r obta in
efficien
y p t
t k lt in recenty
years o ese d ata disagre e E 6
byfa c t ors as l e accurracy of our OA LA

'mated as +10/p.
a a is estim O
4
ow the large dif-
Figures 15 and 16 Rlsoo clearly show
ferences in the
t absorption s
—4
E
C3
e arger tha. n th a t b yDO and
nees" at 460 cm'i
b in D, O. Th r ibution —2
longer
s and comb'
vibratio or example, the knees
0
o e o— . quencies v~ E
C3
ic are reason
a armonic formula. 0
I

v,
~n (H, O) = &(362 0 —63n)cm '
C

0 I I

$6 800 37000
armonics forrDO
ar ')
e near ir, since g requency
ig
' FIG. 17. M Measured absor
sorption prof il
'
v (cm
m=8 to n, =6 ha-
0 —H stretchin monies of th et i b enzene by laaser OA spec-
etch in liquid
tor of v 2. rs and wh y the red
is easy to underst tro scopy.
-ir absor t'
I d t ll th' ay e used to scopy (Moses and Tang 19 on
' ,
t d l no ogy where wat
dye solvent b ecause of it 1978) A er
fth s ixth harmonic e is
I th d e at- 18 h
ive properties of H 0 y norma1. ized to the o facilitate comp
0
seof DO h 1 dd th e absorption
b b th I

d er of magnitude.
t
n itself by more th PRESENT EXPERIMENT
SWOFFORD et al.
2. Trans
ransparent
a organicc I Iqu Id s MOS
Almost all common puree organic liq ul'd s Rre STON
parent" in th e visible p Ch
I-
'
narefre qu y z 6—
or usuall e. A s in the caseeoof water t
a]or s ource of weak n s ructur Ct

ones of f d tal or combination


ional modes. H'igher
h overtones apidly Co
4—
in intensit
are needed
'
ea ly suitable forr thee investi 'g atiionof
one profiles a is for ben-
l, etone, and ootherer common
c an'
or ganic
Our obserrved 0 s pectra
r d OA e for benzen I 1

I6,800
I . &

16,600
I I

I6, 400
I

16,200
~ (crn-~)
a d e ghth harm
Compar iso rption prof'
l t d db oth ese pre-
l d th ing s e escription of ot h er methods
6), 1 ng t —modulation spec tro-

d. Phys. , Vol.
Rev. Mod. 53, No.
N 3, July 1981
C. K. N. Patea and A. G. Tam: Pulsed optoacoustic spectroscopy

Significant disagreements exist (Patel et al. , 1979) 400— I I I I l 1

between the optoacoustic absorption data and some of


the previous data; however, we believe that our profile,
with its absoLute scale (see Fig. 17) is the most reliable
so far.
The detection of the high overtones in benzene al- 500—
lows us to examine the systematic trend of the series E
of overtones. The frequency v„ofthe nth harmonic
of the C-H stretch is found to be surprisingly well
fitted by the following anharmonic formula: X

v„(CH ) =n(3090. 5 —57. 6n) cm '. (51)


200—
I-
It may seem curious that a high harmonic remains a Cl

well-defined absorption feature, since the large number LU

of possible combinations of vibrationaL modes (of a wide


range of energies because of anharmonicity) may Lead )oo
one to think that a high harmonic absorption profile
would be much more spread out than that actually ob-
served. The reason is that a high overtone is best
described by a local mode (where all the vibrational
excitation is localized in one bond) rather thanby nor- l I

6 7 8
mal modes (where linear combinations of vibrations in 0
various bonds are used). The current research on local
mode theory is given, e.g. , by Henry (1968, 1977) and
FIG. 20. Observed linewidth {full width at half maximum)
for the gath harmonic of the C —H stretch in liquid benzene as a
g„
co-workers (Hayward and Henry, 1976; Hayward et al. , function of ~.
1973).
As mentioned earlier, the absorption intensities of maximum (HWHM) on the blue side to that on the red
higher overtones rapidly drop at higher n. The magni- side, as a function of n where the harmonic absorptions
tude of the peak absorption, and the absorption line- appear as a single component. The liquid-phase
width (full width at half maximum) as functions of n linewidth data are in sharp contrast to the gas-phase
are plotted in Figs. 19 and 20, respectively. No satis- linewidths, which are roughly a factor of 3 narrower
factory theoretical understanding of these functional than in the liquid ease, and slowly diminish with in-
depengences is available at present. Figure 20 shows creasing n for n =5, 6, and 7 [see Patel and Tam (1979b)
that linewidths of the harmonic series broaden linearly for a comparison with the work of Beddy, Bray, and
with n. Further, Fig. 21 shows the assymmetry pa- -
Berry (1978)J. The large difference between the gas
rameter, defined as the ratio of half-width at half and liquid cases must be partially due to the proximity
of benzene molecules in the liquid. This is seen when
liquid benzene is diluted in CC14,. the absorption profile
(Fig. 22) of the sixth harmonic becomes narrower,
more blue-shifted, and more symmetrical as dilution

10
2.0
1
10
E
c )0
O ).6
~o"— IX
lK
C) LLJ
(0 X
CO
to 4
bC

Ld
0
)0

/
//
1.0 t l ! ! l

2 5 4 5 6 7 8
~-5i t I t t i I

1 2 5 4
5 6 7 8 FIG. 21. Measured asymmetry parameter (ratio half-width at
0 half-height on the high-frequency side of the peak to that on the
FIG. 19. Observed variations of peak absorption coefficient
with
n„ low-frequency side of peak) for the overtone spectra of benzene
pg. as a function of n.

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: PuIsed optoacoustic spectroscopy

sorption is equal to the product of the absorption co-


efficient n and the path length I,, weak absorption can
result either from small o. (as in the above cases) or
from small l. For the case of benzene, we have
demonstrated that we can detect a fractional absorption
of ~10 '. If we extrapolate this result to the case of
short path length, we may estimate that we can detect
absorption features due to a th ickness of / —10 ' c m if
the absorption coefficient n is -102 cm ', a moderately
large value. This would clearly open up a lot of pos-
sibilities in thin-film spectroscopy and surface science.
The above estimates of our ability to measure thin-
0 film absorption may be too optimistic, because (a)
the effect of scattered laser light by the film may
create serious problems, and (b) there may be no good
0 way to directly couple the transducer to the film, as
in the bulk liquid cases previously discussed. How-
ever, if the film is mounted firmly onto a transparent
substrate (e.g. , by sandwiching the film between the
substrate and a cover plate), then the pulsed OA effect
in the film can launch both bulk as well as surface
acoustic waves into the substrate. As discussed in
Sec. ID.D, the acoustic waves can be efficiently de-
tected by a piezoelectric transducer directly coupled
to the substrate at a di. stance -10 cm away from the
i6200 laser-irradiated region, thus minimizing light scat-
v {cm-~) tering onto the transducer. Light scattering can fur-
FIG. 22. Effect of dilution (of benzene in CC14) on the profile ther be reduced by a suitable shape of the substrate,
of the sixth harmonic in benzene. S(v) is the absorption coef- i.e. , a shape with bends. These considerations have
ficient in arbitrary linear units. The concentrations of benzene led us to a new type of pulsed OA 'cell, " shown in
by volume in the solution are indicated. Positions of the ab-
,

sorption peaks, obtained by averaging over several scans, are


Fig. 6. We found that this cell is useful both for
given. liquid films and for powder films, clamped between
the quartz plates.
To test the usefulness of this arrangement for liquid
films, we have examined (Patel and Tam, 1980) the
increases, i. .e. , the profile tends toward the gas-phase absorption spectra of film of rare-earth chloride
case as dilution increases. Table II summarizes the aqueous solutions, of concentration about 1 gm cm ~.
measured line center, linewidth, and assymmetry pa- Some results are shown in Fig. 23. From these stud-
rameter for the sixth harmonic of benzene as a function
of dilution in CCl, . Incidentally, we can estimate from
ies, we estimate that we can presently detect films
that cause a fractional absorption of 10 '. This clearly
the signal-to-noise ratio seen in Fig. 22(d) for the 1/p
shows that absorption features due to monolayers on a
dilution case that we can presently detect absorption
transparent substrate can be detected with the pulsed
coefficients at least as small as 10 6 cm '.
optoacoustic spectroscopy technique if the absorption
3. Thin liquid films coefficient of the molecules in the monolayer exceeds
10' cm ', and if the substrate ig of high enough trans-
In the above discussions, we have demonstrated that
parency.
the pulsed OA technique represents a powerful method
for detecting weak absorptions. Since fractional ab-
I I I I I I
0 y SOLUTION
f

(2pm THICK)—
TABLE II. Line center position, linewidth (FWHM), and sym- 2
metry of the sixth overtone of the C-H stretch in benzene as a 1-
function of benzene dilution in carbon tetrachloride. Corre- CL
0- I I I I I ~ r — r —
I I

sponding data for gas-phase results are also given. C)


CA (b) PrCkq SOLUTION (5p, rn THICK)
1-
Dilut ion Position L ines idth
(C6H6 in CCl4) (cln ) (cm )
O (c) Hp(gy SOUjTION (4prA THICK)
1.0 16467+ 3 278+ 9 1.55 I-
0.3 16472 + 3 259+ 9 1.49
0.3 16481+ 3 228* 9 1.41 LI
I I
0.03 16496+ 3 222+ 9 1.13 19000 20000 21000
0.01 16 500+ 3 220+ 20 1.09 FREQUENCY (cm ~)
Gas phase 16 550 FIG. 23. OA absorption spectra of thin films of NdC13, PrC13,
and HoCl3 aqueous solutions.

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


'C. K. N. Patel and A. C. Tam: PUlsed optoacoustic spectroscopy 535

The idea of external coupling of a transducer to a


sample for optoacoustic detection (Fig. 6) has been ex- HO&0~ (POWDER)
tended by Tam and Patel (1980) to construct an extreme- C5
CO
ly corrosion resistant optoacoustic cell for liquids.
Such arrangements have been shown by Voigtman ef +E. o
D
(1981) to be well suited to detect trace amounts of dyes, UJ

drugs, and biochemicals.


O
4. Powders and solids
I I I I I
) 8, 400 18,600 18,800
One attractive feature of GA spectroscopy is that ab- FREQUENCY (cm } ~
sorption spectra of highly light-scattering materials FIG. 24. OA absorption spectrum in the green-'blue region for
(e.g. , powders) can be examined. This advantage has Ho2G3 powder at 25 C. Prominent spectral lines are marked.
been frequently emphasized in conventional OA tech-
niques (Monohan and Nolle, 1977; Rosencwaig, 1977;
Rosencwaig, 1978). However, the "immunity" of con- ders shown in Figs. 25 and 26. The lines are due to
ventional OA techniques (utilizing an enclosed gas cell transitions between Stark components of the ground
containing the powder, a gas-phase microphone, state and Stark components of an excited state. We
chopped cw light beam, and phase-sensitive detection) have derived the ground-state Stark splittings and the
to light scattering is often exaggerated. In fact, strong excited-state Stark splittings in the powdered crystals
light scattering can cause optical absorption at cell at room temperature. These data are shown in Table
walls and at the microphone, producing spurious sig- V and Table VI for Dy, O, and E r, O„respectively.
nals (McClelland and Kniseley, 1976). Furthermore, More accurate splittings can be derived by cooling the
most of the conventional OA powder measurements powders. Our work seems to provide the first example
have been done with samples thick enough so that all that quantitative spectroscopy of ions in finely pow-
the incident light is absorbed. This poses the problem dered crystals can be performed by the OA technique.
of uncertain optical intensity dependence on the pene- We would like to mention, however, that while we have
trating depth, since strong Light scattering causes en- so far established a capability for obtaining high-
hanced light intensity nearer the top layer of the powder resolution absorption spectra of powders, we have left
sample. There are also subtle problems with chopping the questions of linearity range unanswered. Unlike
frequency dependence and signal saturation in optically the case of liquids where we have rigorously checked
thick samples. Thus, although numerous authors have to assure linearity of the optoacoustic signal a, s a func-
used the conventional OA technique to.examine powder tion of the absorption coefficient a (to make sure that
spectra, few quantitative data producing new informa- nl is small), no OA signal generation theory for powder
tion have been reported. For example, reported OA samples has yet been put forward. This area is likely
data on Ho20, powders have been numerous (Blank and to become very important, and additional work, both
Wakefield, 19'79; Eaton and Stuart, 19'78; Rosencwaig, theoretical and experimental, is needed.
1977), but only the recent data of Shaw (1979) are of Sam and Shand (19'79) have used a slightly different
respectable quality. technique for carrying out pulsed optoacoustic spectro-
During our efforts to overcome the above problems scopy of weakly absorbing solids. Instead of using a
we discovered (Tam and Patel, 1979c) that the thin- piezoelectric transducer that is directly attached to the
film apparatus shown in Fig. 6 is ideally suited for solid, they immerse the solid sample in a transparent
OA spectroscopy of thin layers of powders. The fine liquid and detect the optoacoustically generated pres-
powder suspension in a clear viscous liquid is clamped sure pulse using a submerged piezoelectric transducer,
between the quartz plates, and the OA signal generated as done by Patel and Tam (1979a) for the measure-
in the powder couples very efficiently into the liquid ments of weak optical spectra of liquids. This ex-
and then into the substrate. The powder film is op- tension of the Patel and Tam (1979a) technique, which
tically thin, so that we do not encounter complications they call the SLT (solid-liquid-transducer) technique,
due to total light absorption (like the signal saturation is reminiscent of the photoacoustic spectroscopy tech-
problems mentioned above). nique described by Harschbarger and Robin (1978) and
Examples of OA spectra of rare-earth oxide powders,
Ho, O~, Dy20„and Er, O„are shown in Figs. 24-26,
respectively. The laser bandwidth (2 cm ') at present
is sufficiently small to resolve all the spectral lines C9 Dyad (POWDER)
(of width -10 cm ') of the room-temperature
CO
powdered
C)
crystals in the spectral range investigated. More de- C)
e
tailed investigations can be made possible by cooling LU
t4
the powder to cryogenic temperatures (when linewidth
can be &O. l cm '), and performing OA spectroscopy LL
C) I
using a dye laser with suitable intracavity etalon to
t I I I

20, 800 21,000 2), 200


narrow the laser linewidth. FREQUENCY (Cm-&)
We have identified (Tam and Patel, 1979c) the lines FIG. 25. Optoacoustic spectrum in the blue-green region for
for Dy, O, (Table HI) and Er, O, (Table IV) oxide pow- Dy203 powder at 25 C. Identifiable peaks are indicated.

Rev. Mod. Phys. , Vol. 53, No. 3, July 198'l


536 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

Ey~Q~ (POWDFR)
TABLE IV. Lines of the 2 Hl f/g 4 If5/2 and 4 Fv/2 4 If5/2»sorp-
C9
tion multiplets presently observed in Er203 powder at 25 C.
(fl Notations are same as in Table III, except that I„,
H„, and F„
D 6 20 denote sublevels of 4 I~5/2, 2 H~~/2, and 4 FY/2, respectively, with
C)
lU 17 energy increasing for larger yz. These energy sublevels are
N
tabulated in Table VI.

O No. ~exp ~calc


I i I

) 9,000 19,200 20, 200 20, 400 20,600 1 18 943 12 10 18 944


Hg I4
FRFQUENCY (cm-&) 2 18 963 1p 8 H2 I3 18 963
FIG. 26. Optoacoustic spectrum in the blue-green region for 3 18 990 60 9 H3 I3 18 988
Er203 powder at 25 C. Prominent peaks are indicated. 19 025 4p 10 H3 I2 19 025
5 19 062 34 8 H3 Ig 19 063
6 . 19093 28 10 H4 I4 19 094
Rosencwaig (1973), which was an extension of the highly 7 19 107 5 H4 I3 19 107
successful gas-phase optoacoustic spectroscopy tech- 8 18 122 25 6 H5 I4 19 124
nique that utilized microphones as transducers (Kreu- 9 19 135 100 10 H5 I3 19 137
zer, 1971; Kreuzer and Patel, 1971) to condensed mat- 10 19 157 8 H6 I3 -. 19 157
ter. In the photoacoustic spectroscopy scheme (as it has 11 19 174 8 H5 I2 19 174
12 19 195 90 10 H6 19 194
come to be known) of Harshbarger and Robin (1973) and
13 19 232 14p 10 H6 Ig 19 232
Rosencwaig (1975), the photo-thermal signal generated 14 20 168 9 20
in a solid (for example) is coupled to a gas and detected 15 20 247 7 10 F& -I4 20 246
with a sensitive microphone, and is dubbed the SGM 16 20 259 30 12 Fg I3 2P 259
(solid-gas-microphone) technique. In the SGM case, 17 20 295 48 10 Fg —I2 20 296
there are two origins of sound generation: (a) primary 18 20334 60 14 Fg Ig 20 334
QA signal generation in the solid, with the signal only 19 20 395' 5 F2 Ig 20 396
20 20 426 70 14 F3 I3 20 425
very weakly transmitted across the gas-solid interface, 21 20 462 22 12 20 462
F3 Ig
and (b) secondary photoacoustic signal genera. tion in the 22 20 496 1p 20
gas due to thermal coupling with the illuminated solid 23 20 522 6 F4 I2 20 522
surface. In SGM detection, the detected signal is mainly 24 20 560 3 F4-4 20 560
due to source (b) (see, for example, Rosencwaig and
Gersho, 1976; McDonald and Wetsel, 1978), and the
poor acoustic coupling between solid and gas is not an
issue. However, in other schemes whe re primary
-

acoustic signals are monitored„acoustic impedances The reflection of a sound wave at an interface between
of the solids or liquids involved and the coupling two materials having dissimilar acoustic -impedances
schemes used are important factors to consider. For is given by
example,
Z„(1)—Z„(2)
Z~c =PV~ (52) .
Z. (1) +Z„(2) (53}

gives Thus we see that the acoustic reflections R~, and A, q

', gas-solid and liquid-solid interfaces, respectively,


Z (he„lium gas) =5 gmcm
are (typically)
Z„(solid)=8X10' gmcm ', A~, = 0. 999 91 (54)
Z„(liquid) = 2 x 10' gm cm '.
A, =0.56. (55}
TABLE III. Lines of the F~/2 —H~5/2 absorption multiplet
presently observed in Dy203 powder at 25 C. vexp is the ob-
served peak position (cm ), I is the intensity (arb. unit), ' W is
full width at half maximum (cm ), T is the transition identi-
TABLE V. Energy levels of the lower Stark sublevels in the
ground state 6 H~&/2 and in an excited state 4 Fe/2 of 'Dy 3+ in
fied, and v«&c is the calculated peak position (cm ) according
to the energy levels presently fitted, as given in Table V. Dy 203.

No. Energy (cm )


+exp &calc
Hender son
10 780 13 40 20 780
State Sublevel This work et gl. (1967)
Fg H3
20 868 6 31 F3 H3 20 866
20 971 20 4p 20 971 H~s/2 H2 67
Fg H2
20 994 10 22 20 994 H2 67 74
F2 Hg
21 038 10 25 Fg —Hg 21 038 H3 258 261
21 056 10 36 F3 Hg 21 057 4
FB/2 F~ 21 038 21 040
21 121 35 27 F3 — Hg 21 124- F2 21 061 21 064
21 260 13 43 F4 —H2 21 259 F3 21 124 21 131
21 325 16 34 21 326 F4 21 326 21 339

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

TABLE VI. Energy levels of the lower Stark sublevels in the WAVELENGTH (nm)
ground state I~5~2, and of the Stark sublevels in the excited 610 600 590
states H&&g2 and Fv(2 of Er ' in Er203.
1

—.25
Energy (cm ~) .020—
Gruber Nd: BEL
State Sublevel This work et al. (1966)
—.2
I-
(n .0 I 5—
I|5(2 Ig 0 0
I2 38 38
UJ
CI .l 5
I3 75 75
I4 88 88 o Olo
2
H~~t2 Hg 19 032 19 033 O
19 038 19 040
H3 19 063 19 067 .005 .05
H4 19 182 19 182
H5 19 212 19 213
H6 19 232 19 235 I I

F( 20 334 20 336 I 6400 i 6600 l6800 I7000


LASER FREQUENCY (cm ')
r

F 20 434 20 435
F3 20 500 20 500 FIG. 27. Optoacoustic absorption spectrum of Nd: BEL mea-
F4 20 560 20 561 sured using the solid-liquid-transducer technique. The sam-
ple was immersed in CC14 (after Sam and Shand, 1979).

Thus we can see that a sound wavehas a transmission -10~ cm ~

factor of —10 ' across a solid-gas interface, and ther- One advantageof the SLT scheme is the lack of scat-
mal coupling between the solid and gas generally tered light at the liquid-solid interfaces due to a good
dominates in the SGM scheme. On the other hand, in (but not perfect) refractive index match. Hence sur-
OA experiments where the primary acoustic signal is face preparation may not be very crucial. However,
measured, sound coupling must be optimized. In this the time gating scheme using a zig-zag substrate and
consideration, the SLT scheme is not as efficient as direct attached-transducer scheme described for the
the attached-transducer scheme since the acoustic study of powders (above} eliminates the scattering
signal generated in the solid has to be transmitted to problem anyway. Thus the refractive index matching
the liquid in which it is immersed, and further the capability may be outweighed by a more serious prob-
sound wave in the liquid has to be launched into the lem of finding sufficiently transparent liquids that do
piezoelectric transducer for detection. Thus we have not attack and dissolve the solid samples.
to contend with two solid-liquid interfaces giving a
total transmission coefficient of (1 R. ~)' 0.19 for B. Nonlinear spectroscopy
transferring an acoustic signal from the solid sample As mentioned in Sec. IV, the use of a pulsed laser
to the piezoelectric transducer. The attached-trans- source, with its high peak power and low duty cycle
ducer scheme, on the other hand, will have an acoustic (thus preventing significant heating of the sample},
signal transmission factor of ~0.9. Hence the SLT permits OA nonlinear spectroscopy to be readily
scheme is much better than the SGM scheme as far as performed. By nonlinear spectroscopy we mean
acoustic coupling is concerned, but still falls short of spectroscopic studies of processes involving inter-
the attached-transducer technique. Sam and Shand action of more than one photon. In these studies it is
(19'l9), however, suggest that the SLT scheme may be usually necessary to have high optical intensities. How-
simpler to use for relative and rapid characterization ever, the intensities must be kept sufficiently low so
of many different materials. that runaway self-action effects are not generated. For
An example of pulsed optoacoustic spectroscopy example, effects like self -f ocus ing or stimulated
based on the SLT scheme is the measurement of the Baman scattering originating from amplification of
absorption spectra of Nd:La~Be, O, (Nd:BEL) crysta. l spontaneous emissions need to be prevented in a study
with one atomic percent doping. The solid sample is of two-photon absorption or of Baman-gain spectro-
suspended in a liquid optoacoustic cell directly above scopy. Hence the choice of a flash-lamp-pumped dye
the piezoelectric transducer. The liquid, of course, laser as the light source for nonlinear OA spectroscopy
has to be sufficiently transparent in the spectral region is quite appropriate in many ca, ses. The laser pulse
of interes't. Sam and Shand used CC14, which has a energy of -1 mJ and duration -1 csee is usually
very low absorption coefficient =-10 ~ cm ' in the visi- focused to a. spot size -10 4 to 10 ' em' in our non-
ble region. Figure 27 shows one of the spectra re- linear spectroscopy studies. Thus 'the laser intensity
ported by Sam and Shand for Nd:BEL in the 16300 to is -10' to 10' % cm ', and is well below the onset of
17000 cm ' region. From the signal-to-noise ratio, self -action effects that may interfere with the studies.
they estimate a capability of measuring down to an Another effect that one needs to be cautious about when
absorption coefficient of -10 ' cm '; however, Fig. 27 dealing with a high-intensity laser beam is electro-
indicates that in a Nd:BEL/CC1, system it would be dif- striction (Bebchuk et al. , 1978, Brueck et a&. , 1980).
ficult to measure an absorption coefficient smaller than 'This effect is due to a positive or negative polarizability

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


538 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

of the molecules in the sample, causing the molecules


10
to move towards or away from regions of higher optical
intensities. The electrostriction effect is proportional
to the laser intensity, and may interfere with the non-
linear optical studies if we use laser pulses of shorter
duration but keep the pulsed energy constant. (See dis-
cussion of the electrostriction effect in Sec. II.D. )
Few nonlinear OA spectroscopy studies of condensed
matter have been reported in the literature. This may 2
be due to a lack of sensitivity (nonlinear absorption
b
effects are usually very weak) without the use of a 0 ——
(b) CIRCULAR
pulsed laser in the OA studies performed so far. We X POLAR ZAT ION
I

show here that weak nonlinear absorptions can be


readily detected with the pulsed OA technique, as
shown by our examples on two-photon absorption and
Raman-gain spectroscopy. Such OA investigations have
not been reported previously. It is clear that many 0— I

more new nonlinear spectroscopy investigations of con- 1 9000 20000 21000 22000
V (cm-I)
densed matte r are now possible, because of the high
sensitivity of the pulsed OA technique. Furthermore, FIG. 28. Spectra of the two-photon B» A~~ band in liquid
benzene. 0.(p) is the two-photon absorption cross section in
OA detection is frequently the preferred detection
10 cm sec molecules photon, and v is the single-photon
method for absorption in condensed matter, rather than frequency in cm . Note that the absorption peak at 18 810
fluorescence detection; this is because fluorescence cm is not a member of this band, but is the peak of the
is often quenched in condensed matter (but not in dilute seventh harmonic of the C-H stretch. The relative heights of
gases). the different vibrational transitions are quite different for (a)
linear laser polarization and (b) circular laser polarization.
1. Two-photon absorption
We have performed (Tam and Patel, 1979b) the first the transducer. This is verified for focal lengths in the
quantitative OA two-photon absorption spectroscopy range of 8 to 2 cm. The reason is that a shorter focal

absorption ar is ing from the '&, „-'A„


of liquids. We have studied the two-photon electronic
trans ition in
benzene, which is symmetry forbidden, but becomes
length results in a higher laser intensity but a corres-
pondingly shorter confocal distance, and these two ef-
fects nearly counterbalance each other. We have also
allowed in two-photon absorption by vibronic mixing. verified that the two-photon absorption peaks (all ex-
The forbidden nature of this transition results in a cept X) in Fig. 28 have intensities proportional to the
small two-photon absorption cross section 0 through- square of the laser pulse energy. Hence the data of
out this band, and o is always smaller than 10 ~' Fig. 28 have been normalized to the square of the pulse
cm' sec molecule ' photon '. Fully allowed two-photon energy. (Use of the minicomputer for digitizing signals
transitions usually have cross sections that are 2 —3 A and &, in Fig. 28, facilitates the normalizing of the
orders of magnitude larger. However, the optoacoustic OA signal by the square of the laser pulse energy. )
technique is adequate even for weak two-photon transi- The peak X varies linearly with the laser intensity and
tions. arises from the seventh harmonic absorption in ben-
Experimentally observed two-photon GA spectra are zene as discussed in Sec. V. A. 2.
shown in Fig. 28, for linear and circular laser po- Vibrational intervals and electronic origin can be de-
larizations. To obtain these data, the OA cell (Fig. 4) rived (Tam and Patel, 1979b) from the data of Fig. 28.
is filled with distilled spectro-grade benzene, and a These are summarized in Table VII, which identifies
lens with a focal length of 5 cm is used to focus the
dye laser beam at the center of the cell, as shown in 5/o BEAM-
Fig. 29. Multiple distillation of the commercially avail- SPLI TTER LIQUID FILLED CELL QUARTZ
WINDOW
able benzene was seen to be necessary to eliminate sus- PUL SED
LENS

pended particles which were seen to give rise to large DYE


LASER
extraneous noise output, if the particles are irradiated IL
PZT
TRANSDUCER
by a focused light beam. The laser focus should be ver- MP
2 METER
tically in front of the center of the transducer diaphragm SQUARING
SPECTROMETER II

for optimum detection efficiency. This is because the CIRCUIT AMPLI F I ER

source of acoustic energy is now nearly a point source IF

near the focus, rather than a line source as in the linear BOXCAR
A
SCAN
absorption cases. Experimentally, this fine adjustment DRIVE

is performed by moving the lens towards or away LASER SCAN


NOVA
COMPUTER
from the cell, so as to optimize the transducer signal
observed on an oscilloscope. PRIN TER VIDEO MONITOR
The two-photon OA s ignal intens ity is found to be FIG. 29. Schematic of experimental arrangement for measur-
nearly independent of the focal length of the lens used, ing two-photon optoacoustic spectra of liquids. The lens focal
if the laser focus is adjusted to be directly in front of length is typically 5 cm.

Rey. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tarn: Pulsed optoacoustic spectroscopy

TABLE VII. Observed peaks by Two-Photon Optoacoustic Spectroscopy, TPOAS, of the ~A& B2„—
band in liquid benzene. All values given are for linearly polarized laser light. Listed columns
show laser frequency and notation of an absorption peak, the exponent of its laser energy depen-
dence (+0.1), two-photon absorption cross section, vibrational energy in the B2„state {we set the
electronic origin at 37724 cm in the liquid phase), identification of the transition, and calcu-
lated position (above the electronic origin) of the absorption peak assuming gas-phase vibrational
constants.

Observed peaks Calculated '


Frequency Power Cross 2v —37 724 position
v (cm ) Legend dependence section ~ (c m Ident if ic ation (cm )
) ~

18 810 1-Ph 7th har m of C —H "


19 323 2. 1 18() 923
19 644 14() 1564
19 792 1.5 1860 18olo 1843
20 106 10 14o 1() . 2488
14() 10o 2731
20 241 1.8 2758
18~o12 2770
20 569 9.2 14ol() 3414
20 690 2.1 3656 14ol021 3656
21 031 9.0 4338 14olo 4333
21 047 1-ph 8th harm of C —H '
21 139 4554 14o10olo 4562
21487 4. 8 5250 14', 14, 5256
21 627 1.7 5530 olo~o13o 5485
21 957 3.1 6190 14', 1', 6179

10 cm molecule photon sec. The relative values of these cross sections are accurate to
+10 jp for the bigger peaks. The uncertainty in the absolute scale is a factor of 3 (see Tam and
Patel, 1979a).
"Using Wilson's notation {Wilson, 1934).
'Using Wunsch eI; at. 's constants (Wunsch, Neusser, and Schlag, 1975a, 1975b).
dPatel and Tam, 1979b.
'Patel, Tam, and Kerl, 1979.

the vibronic origin of each of the peaks and gives ac- vibrational transition lines. This dependence has not
curate cross-section values for them. While the vi- been previously observed for liquids, although such
brational intervals are nearly identical to those known strong dependence on polarization for gases is well
for benzene in the gas phase, the electronic origin for known (see Tam and Patel, 19'T9b, for details). How-
the band '&3„-'A,~ is quite different. The electronic ever, as expected, the one-photon absorption line at
orig in in the gas phase is known to be 38 086 cm ', in 18810 cm ' is found to be independent of laser polari-
contrast to our value of 37724 cm ' for the liquid ben- zations.
zene case (see Tam and Patel, 1979b, for details}. Our work in benzene demonstrates that multiphoton
No quantitative explanation for this large red shift in transitions in condensed matter can be readily ob-
the electronic origin is available at present. served by the pulsed OA technique. Multiphoton in-
We can also derive quantitative values for the two- vestigations are usefulbecause they map out states that
photon cross sections o for the present case by corn- are not one-photon accessible (e.g. , because of sym-
paring (Tam and Patel, 19'l9b) the intensity of the metry considerations). Furthermore, high-lying elec-
two-photon transition with that of the one-photon transi- tronic states or ionization states can be reached by
tion (seventh harmonic of the C-H stretch) at 18810 visible photons via multiphoton transitions. OA multi-
cm, shown in Fig. 28. The absorption coefficient at photon spectroscopy should permit the discovery of
18810 cm ' is already measured by us to be 4.3X 10 many new multiphoton transitions (strong or weak} in
cm ', as seen from Fig. 17. We thus obtain an abso- condensed matter.
lute scale for. the two-photon spectra shown in Fig. 28. Earlier attempts at observation of two-photon ab-
We may estimate from the signal-to-noise ratio in sorption by optoacoustic methods (Bonch-Bruevich,
"
Fig. 28 that two-photon cross sections as small as
10 cm4 sec molecule ' photon ' can be detected at
Razumova, and Starobogatov, 19'I'i) have shown that
two-photon absorption cross sections for anthracene
4' cm' sec photon ' mole-
present. . in alcohol at 6943 A are =10
We also note from Fig. 28 that strong dependence on cule ', while those for POPOP solution at 6943 A are
the laser polarization exists for the various two-photon =4 x10~' cm' sec photon ' molecule '. Figure 30

Rev. Mod. Phys. , Vol. 53, No. 3, JUIy &98&


540 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

800 Here, c is the velocity of light, N is the scatterer


density, da'/dQ is the differential Raman scattering
700 cross section per scatterer, h is Planck's constant,
6oo— n, is the refractive index at v„1/{1
+m) is the Boltz-
man factor for the ground state, and I is the spon-
C9
500— taneous Raman scattering linewidth (full width at half
C3
I- maximum). In Eq. (56) we have assumed that the
co 400 Raman gain g, is small -enough so that we can replace
C)
300— e'sr~' by 1+g, I~l. 'The conventional way to perform
I

O Raman-gain spectroscopy (Heritage et al. , 1979;


O
200— Levine and Bethea, 1979) is to measure the incident I,
I 00 and the exit (I, +LA, ). The difference between these two
quantities is then evaluated to obtain M, . Since M,
14,000 I3,000
l

I2,000
is usually small (e, g. , much less than 10 '), this
method of subtracting one large number from another
FUNDAMENTAL FREQUENCY (Crn-~)
t;o get a small difference is possible only if each num-
FIG. 30. Two-photon absorption spectrum of anthracene in ber can be very accurately measured; hence, stable and
alcohol measured by the pulsed laser optoacoustic technique.
The two-photon absorption cross section at the peak is -10
sophisticated laser sources are needed.
cm sec molecules »f photon (af",er Bonch-Bruevich et ag. , Our OA approach to this problem is to directly mea-
~. 977). sure the difference. For each photon conversion from
the pump beam to the signal beam, the amount of en-
ergy h(v~ —v, ) = hv& is lost from the optical beams and
shows a two-photon absorption-spectrum of anthracene is deposited into the medium. Let the signal incident
measured with the optoacoustic technique. These data pulse energy be E, and the increase in pulse energy be
do not show the detailed vibronic structure we saw AE, . Then the number of photons that have undergone
earlier for benzene, and further the peak two-photon conversion is AE, /(hv, ). Hence the total energy de-
absorption cross section of -10 ' cm' sec pho-
gal
posited in the medium is
ton ' molecule ' suggests that the two-photon transi-
~ =&&.('~/&;) . (56)
tion is a strongly allowed one, unlike the A„-&»
transition of benzene, which shows a maximum two- Integrating Eq. (56) over the duration of the laser
photon absorption cross section of -10 " pulses (the pump pulse and the signal pulse should oc-
cm~ sec photon 'molecule '. It is worth noting that the cur over the same time interval), we have
sensitivity of optoacoustic detection quoted by Bonch-
Bruevich et al. (1977) is -10 6 Z of absorbed single- AE, = g, IpE, /, (59}
pulse energy. This observation corresponds to three- where I~ is an average pump intensity during a pulse.
to -four order s of magnitude lower sensitivity than that Combining Eqs. (58) and (59), we have
reported by Patel and Tam (1979a).
&&z = ( 8 R/&s) Ip& s ~ (60)
2. Raman-gain spectroscopy Equation (60) is the basis of OA Raman-gain spectro-
There is strong current interest in Raman-gain scopy (OARS). The energy AE~ that is deposited in the
spectroscopy (e.g. , Heritage et a/. , 1979; Levine and medium due to the Baman gain usually converts rapidly
Bethea, 1979) because of its potentially high sensitivity into heat in condensed media, and can be detected di-
in measuring Baman frequencies and cross sections, rectly by the pulsed OA technique. Our work is the
and because it is applicable to luminescent samples and first demonstration of the use of OARS in liquids, al-
hostile environments (flames, discharges, etc. ), where though a similar technique in gases has recently been
spontaneous Baman scattering cannot be readily used. demonstrated by Ba.rrett and Berry (1979).
A block diagra. m of the arrangement (Patel and Tam,
The theory of Baman-gain spectroscopy is straight-
forward. A "pump" laser beam and a "signal" laser 1979c) for OARS in liquids is shown in Fig. 31. In
this setup, two flash-lamp-pumped dye lasers are
beam, of intensities I~ and I, and frequencies v~ and v„ synchronized. Laser I uses Coumarin 504 dye, while
respectively, are incident on a medium. The signal
beam is def ined to be the beam having the smalle r f re- Laser II uses Bhodamine 6G dye. Both operate at 10
quency. If v~ —v, is equal to a Raman frequency vR of pulses per second. The two laser beams are com-
the medium, then Baman gain occurs, i.e. , I& is attenu- bined by a dichroic mirror, and focused into the center
ated and I, is amplified. %e shall be concerned with of the OA cell. The red frequency v.. is fixed, while
small Raman gain only. In this case, we can show that the green frequency v~ is scanned. %hen the Haman
the increase &I, in the I, passing through the medium is resonance condition, v~ = v, + v&, is satisf ied for v&
equal to the Baman excitation frequency of the medium,
{56} the OA signal is detected by the piezoelectric trans-
where l is the gain path length, and g, is the Baman- ducer. The pulse energy of Laser I varies over its
scanning range. Hence the OA signal is normalized to
gain coefficient (normalized to pump intensity) given by
the pulse energy of Laser I to givea normalized OARS
16m 'c ~N(do /dQ) signal S(v~). Here uR is defined as v~ —v, . S(vz) is
kv,'m~(l +n)I"
(57) thus directly proportional to the Baman cross section

Rev. IVlod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. 7am: Pulsed optoacoustic spectroscopy

TRIGGER v„(Cm I)
I
3000 3)00 3200
SCANNING LASER I I
DRIVE ZE
I

~c
I ASER
I Ug
XDICHROIC
Vl

PYROELECTRIC MIRROR
PECTRO- PREAMP[
DETECTOR I
METER
l~»u R E Rl
T
TR IGGER BOXC AR COMPU TERIA
— BOX CAR
I
B B A A

LASER SCAN

v S= A/B

MON IT OR
8
PRI NTER

= transducer,
tivity beam-splitting.
I
FIG. 31. Experimental arrangement for OARS, C=cell, T
= beam-steering mirrors, and B = 5 jo reQec-
Laser I is the pump laser operating
with Coumarin 504 dye, while Laser II is the signal laser with
Rhodamine 66 dye.

I I I

19200 )9300 ) 9400


da'/dQ, as seen from Eqs. (60) and (57). Since v, is
vp (cm &)
not scanned, its pulse energy F-, should ideally remain
constant during the experimental scan (typically taking FIG. 32. Optoacoustic Raman-gain spectra of benzene for {a)
several minutes), for a stable enough laser. In our parallel and (b) orthogonal pqmp and signal polarizations.
case, 8, actually fluctuates by alOV& during the scan.
This constitutes a major source of noise in our OARS
studies so far. This source of noise could be elimi- cence poses no serious problem for the use of OARS.
nated either by having a much more stable Laser II, For cases where the Raman-excited vibrational modes
or by further normalizing S(vz) by F., (measured by decay slowly (e.g. , in liquid N, ), collisional quenching
another detector). Another important source of noise can be investigated by systematically adding foreign
in our OARS investigations is the relative time jitter molecules to the liquid and studying the increase in the
of the two laser pulses; the jitter is as large as +30% OARS signal. Furthermore, the experimental appara-
of the laser pulse width of -1 p. sec. Ideally, the two tus needed for OARS is simple (much less sophisticated
laser sources for OARS should be automatically than that needed for conventional Raman-gain measure-
synchronized. This could be done by using a two-color ments). Thus OARS should be extremely useful for
flash-lamp-pumped dye laser (one flash-lamp pumping
two dye cells simultaneously). Alternately, a laser-
pumped dye laser could be used, and part of the pump-
ing laser split off and combined with the dye laser to
perform OARS.
Some examples of OARS data for various liquids are
shown in Figs. 32 and 33. These data have been ob-
tained with the nonideal arrangement shown in Fig. 31,
and significant improvements in signal-to-noise should
be possible with better laser sources as discussed
above. We further note that the data of Fig. 32(a) are
obtained with the two laser polarizations parallel;
hence S(v~) is proportional to the polarized Raman
cross section. Totally different spectra have been ob-
tained (Patel and Tam, 1979c) with the two laser po-
larizations orthogonal, as seen in Fig. 32(b), in which
case S(vz) is proportional to the depolarized Raman
cross section. Further (Patel and Tam, 1979c), we
have shown a capability of measuring Raman gain as
small as 10 ' cm ' using the OARS technique. The
data reported in Fig. 33 were obtained for the case of
paralle 1 polarizations.
Additional desirable features of OARS include the
I

2800
i ~ i i I

2900
« l i I

3000
i i i i I

3 )00
following. Since pulsed lasers are used, transient V„(Cm-I)
OARS due to decaying species can be readily studied FIG. 33. Examples of OARS signal observed for various
in a time-resolved experiment. Background lumines- liquids. Laser polarizations are parallel for these signals.

Rev. Mad. Phys. , Vol. 53, No. 3, July 1981


542 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

liquids, including aqueous solutions and biological sam- shapes of the luminescence spectra of the sample and
ples. a standard.
Nechaev and Ponomarev (1975) have proposed the use
of an optoacoustic measurement technique for Raman B. Pulsed OA spectroscopy in strongly absorbing liquids
spectroscopy of gaseous media. The scheme described
Our primary interest is in the area of optoacoustic
is similar to the one used in the study of OARS spectra
of liquids described here, but utilizes cw lasers for the
tectionn
spectroscopy of very low-loss materials. But experi-
mental studies of Sigrist and Kneubuhl (1978), in which
study of low-pressure gases, and microphones for de-
of optoacoustic s ignals. Barrett and Her ry, they have investigated optoacoustic signal generation
in strongly absorbing liquids using a pulsed CO2 laser
1979, have recently demonstrated the effectiveness of
this technique for the study of Haman spectra of gases.
as the source (o. =1 —1000 cm '), provide some data
to support the theoretical considerations above. In the
Vl. OTHER OA SPECTROSCOPY STUDlES strongly absorbing case, we are dealing not with cylin-
drical, but with hemispherical geometry in the genera-
A. OA spectroscopy and luminescence spectroscopy tion of an OA signal. However, some comparisons
should be possible. In particular, the optoacoustic
The use of optoacoustic spectroscopy as a supplement
pulse shape which is predicted to consist of a positive
to luminescence spectroscopy in the study of energy
and a negative pulse separated by time of the order of
relaxation and energy-transfer processes in condensed
the duration of the laser pulse appears to have been
media is exciting, since now we can take into account
verified. In Fig. 34 we show the stress transient in
both the radiative and the nonradiative relaxation pro-
water for CO~ laser pulse energies below vaporization
cesses. Razumova and Starobogatov (19'77) have de-
scribed some preliminary experimental and theoretical threshold, as observed by Sigrist and Kneubuhl. In
studies using optoacoustic spectroscopy and lumines- agreement with the predictions of Eq. (38}, the opto-
acoustic signal consists of a compressional pulse
cence spectroscopy for the quenching "down" and
followed by an almost equal rarefaction pulse. (De-
quenching "up" of excited states in fluorescent organic
tailed theoretical arguments and computer solution of
solutions. They employed a pulsed ruby (including
the sound generation equation by Sigrist and Kneubuhl
frequency doubled radiation) laser. They measured the
optoacoustic signal (derived from an immersed piezo-
are also able to reproduce the experimentally ob-
served pulse shapes. } It should be emphasized, how-
electric transducer} as a function of the luminescence
ever, that because the authors have investigated only
signal in a three-level system, where level one is the
highly absorbing liquids at CO~ laser wavelengths, it
ground state, level two is the first excited state (from
is possible to deposit enough energy to vaporize the
which luminescence can originate), and level three
liquid. In this situation, considerably different pulse
is another excited state to which quenching transitions
shapes are observed.
are made through absorption of radiation, corres-
ponding to the transition between states two and three. C. Other phenomena at high intensities
The detailed theoretical description of rate equations
governing the luminescence and optoacoustic signal At very high laser intensities, it is possible to be
output (for an assumed model) was in good agreement confronted with breakdown and bubble formation in
with measured variation of the optoacoustic signal and liquids. For example, when the nonradiatively re-
the luminescence signals as the pumping, radiation and leased energy in water exceeds -2240 J cm ', there
the quenching radiation intensities were changed. It is is a possibility of vapor bubble formation. Since break-
clear that studies such as these where, in addition to down and associated explosive nonradiative relaxation
the intensities, the frequencies of the two radiation results in acoustic pulse generation, the optoacoustic
sources are varied tnot done in the Razumova and detection scheme can be utilized for gaining insights
Starobagotov (1977}studies] will provide significant into these phenomena. Teslenko (1977) has reported a
information regarding redistribution of excitation with-
in a system following excitation. Further, use of
picosecond pulse lasers and time gating of the excitation I I I I I I I

and quenching pulses (whether nanosecond or pico-


second duration) will provide additional information on bJ
the kinetics of excitation redistribution. As an ex-
o
tension of this technique, Starobogatov (1977) has used 0
the measured optoacoustic signal as a means of de-
termining luminescence efficiency of dye solutions. ~
V3 cn
This is straightforward, since the absorbed energy ~o
~o
(from the exciting beam} and energy released through
nonradiative processes (measured as the optoacoustic
signal) are measured in the experiment, and thus the I I I I I I I I

total luminescence output is the difference between TIME (l00 ns/div)


these two numbers. The advantage of this scheme over FIG. 34. Optoacoustic signal generated in water due to absorp-
the traditional optical measurement method is that tion of a pulse of CO2 laser radiation. Top: Laser pulse.
there is no need to introduce corrections for the spec- Bottom: Optoacoustic signal showing the compressional and
tral sensitivity of the photodetector and locations and the rarefaction pulses (after Sigrist and Kneubuhl, 1978).

Rev. Mod. Phys. , Vol. 53, No. 3, Juty 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 543

study of the optoacoustic and optohydrodynamic pa-


rameters of laser breakdown in mater, glycerine, and
benzene, using a 1 J, 10—70 nsec duration ruby laser. CO

For measurements of linear or nonlinear optical ab-


sorption using the optoacoustic technique under dis- LLI

cussion in this review, it is very unlikely that the ab-


sorbed energy will approach a value anywhere close to Nj
that necessary for shock wave and cavity formation. lK
LLI
C9
D. cw-source-based OA spectroscopy with piezoelectric
transducer \

One of the earliest demonstrations


of the use of a
piezoelectric transducer (either immersed in a liquid l I
0.$ 1.0 10 100
or attached to a solid) for measurements of optical
FREQUENC Y (k HZ j
absorption was that of Kohanzadeh et af. (1975). These FIG. 36. Optoacoustic signal as a function of laser modulation
O
authors used a cw argon ion laser at either 4880A or frequency (for the case of a cylindrical transducer-container
5145 A (1.5 W), a hollow cylindrical PZT transducer geometry shown in Fig. 35), showing the effect of transducer
which was filled with the liquid under investigation, and other resonances (after Kohanzadeh et al. , 1975).
and a mechanical chopper or an electro-optic modu-
lator (which gave a sinusoidal modulation). Since the of using a cm source, piezoelectric detector, and
wavelength of the laser was fixed, no spectroscopy was lock-in amplifier to measurement of small absorption
carried out, but some of the features of immersed- coefficients, it is clear that at the cw power level of
transducer optoacoustic spectroscopy were recognized. 1.5 %, even the lower limit of 0.005 cm ' was beyond
These included a linear dependence of the optoacoustic the capabilities of the technique. The limitations of the
signal on input power and absorption coefficient (varied method stemmed mainly from adherence to a cw source.
by use of methylene blue dye added to control the ab- It can be shown that with the same average power of
sorption), and the time delay of the acoustic signal 1.5 W, but in the form of pulses of -1-2.p. sec duration
measured by determining the phase delay detected with and 10—15 pulses per second (pps) repetition rate,
a lock-in amplifier. These authors further recognized sensitivity could have been much improved. In other
the important resonant response behavior of the piezo-
words, if we take the sensitivity achieved by boatel and
electric transducer for optoacoustic spectroscopy. Tam (19'79b) to be -10 6 —10~ cm ' using input pulse
They used a cylindrical piezoelectric ceramic that was energies of 1 mJ and peak powers of -1 kW with a
simultaneously the container for their liquids and the
pulse repetition frequency of -10 pps, we can extra-
detector for acoustic signals, as shown in Fig. 35. The polate to the case of 1.5 W peak power, pulse repetition
optoacoustic voltage output from their cyIindrical frequency of -100 kHz (square wa, ve modulation). We
transducer is plotted as a function of the laser modula- find the extrapolated sensitivity for the Kohanzadeh
tion frequency in Fig. 36. The absorption of the liquid
experiments to be
(acetone) was increased by the addition of a dye. The
range of absorption coefficients investigated was 0.005
to 0.5 cm '; however, the tower values of the absorp-
tion coefficient were measured not with the optoacoustlc
technique but with a thermal lens technique (Hu, 1973). where Pz and PPT are peak laser powers, fz, f are
Although no attempt mas made to extend the technique pulse repetition frequencies, and n„„„(K) „„„(PT)
PT
and n
are the minimum detectable absorption coefficients for
the Kohenzadeh and boatel- Tam experiments, respec-
tively. With n,
„(PT)=10 ~ cm ', PK=1.5 W,
lS1
P =10' W, f=10 Hz, and fK =10' Hz,
Z

=g x10~ cm '.
The discrepancy„ then, between the observed and the
calculated sensitivities is puzzling.
AIR In a subsequent paper, Fa, rrow et al. (1978), recog-
nizing the importance of acoustic impedance matching
for optimum detection of the eptoacoustic signal carried
further the earlier studies of Kohanzadeh et al. (1975),
still using cm argon ion lasers or dye lasers„and either
mechanically chopping or electro-optically modulating
the laser output (see Fig. 37). While they clearly ap-
\

FIG. 35. Optoacoustic cell consisting of a piezoelectric ma- preciated the versatility of the piezoelectric trans-
terial cylinder that doubles as the container for the liquid being ducers over the gas-phase microphone photoacoustic
studied (after Kohanzadeh et a/. , 1975). spectroscopy technique of Harshbarger and Robin

Rev. IVlod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

I I

450
~
I
MONO-
CHROMATOR = Xe ARC
LAMP
w
I
I

FRONT SURFACE MIRROR CHOPPER I


SAMPLE BEAM X CONTROL I
TRANSDUCER MODULATED LIGHT SPLITTER (STEPPING MOTOR)
~ I
BEAM I I

I
I
I DYE Ar-ION I
SIGNAL PHOTO-I LASER LASER I
DIODE "X CONTROL I
P. A. R.C. 183 I (STEPPING MOTOR) I

P HOTOMETRI C NTENSIT Y
I
I LIGHT SOURCES I
I
PREAMP' IF I FR NORMAL I ZER l I
CHANNEL
t STEPPING MOTOR DRIVE
PA RC 124A
LOCK- IN REFERENCE FREQUENCY — D. E.C. D. E. C.
AMPLIFIER LPS-
LSI - II =
VECTOR
GRAPHICS
(A. D. C) TERMINAL
SYNCHRONOUS SIGNAL LEVEL (D.C. ) f COMPUTER

[D.A. C, i = X5Y
PLOTTER

FIG. 37. Experimental setup for optoacoustic spectroscopy with attached transducer, using chopped or electro-optically modulated
laser radiation in radiation derived from a xenon arc-lamp-monochromator combination (after Farrow et aE. , 1978).

(1973), and Rosencwaig (1973), their conclusion was sured data. Hordvik and Skolnik (1977) estimate that
that the sensitivity achieved with the piezoelectric surface and bulk absorption losses in the 10 ' cm '
transducer scheme was similar to that achieved with range can be determined with cw laser powers of the
gas-phase photoacoustie spectroscopy. Thus it is order of a few hundred milliuatts and that the limitation
apparent from their published data that Farrow et al. to the sensitivity arises from false signals generated by
(1978) failed to achieve the ultra-low-absorption- scattered radiation reaching the transducer. Because
coefficient measurement capability afforded by the of the low-frequency chopping rates involved in these
pulsed laser, immersed or attached piezoelectric studies, Skolnik and his collaborators were not able to
transducer, gated detection technique that is the cen- use the time of flight or special geometrical schemes
tral subject of the present review. described in Sees. V.A. 3 and V. A. 4 which contribute
An extensive study of the cw-laser-based opto- significantly to the reduction adn/ or elimination of false
acoustic technique utilizing an attached piezoelectric signals arising from scattering when using the pulsed
transducer hag been carried out by Hordvik and his laser, immersed or attached piezoelectric transducer,
collaborators and reported in a number of elegant and gated detection. However, for good single crystal
papers (Hordvik, 1977; Hordvik and Schlossberg, materials such as CaF„SrF„BaF„MgO, Al, 03'
197'7; Hordvik and Skolnik, 1977). In their study of LiF, etc. , the cw laser scheme of Skolnik and colleagues
low-loss optical solids Hordvik and Schlossberg has allowed them to measure bulk absorption coefficients
(1977) used cw CO, CO2, or argon ion lasers with a '
as low as 10 cm ' as seen from the data for BaF„
piezoelectric transducer attached directly to the sam- SrF» and CaF2 in Fig. 39.
ple of optical material. Recognizing the importance of
false optoacoustic signals that may arise due to scat-
tered light from the bulk of the optical material being
absorbed by the attached piezoelectric transducer,
they placed a second transducer symmetrically with
respect to the optical beam, close to but not in ATTACHED
acoustical contact with the sample (see Fig. 38). CHOPPER TRANSDUCER
that by subtracting the unattached- -v' SAMPLE
They argued I
~
transducer signal from the attached-transducer signal, CW LASER
POWER
the effect of bulk scattered light, especially in low- UNATTACHED METER
absorption-loss materials, could be minimized. Fur- TRANSDUCER

ther, by the proper choice of sign and magnitude of the REFERENCE


piezoelectric constant of the transducer material, the
optoacoustic signal generated by bulk absorption could
be made to be exactly 180' out of phase with that
generated due to surface absorption. They also ob- LOCK- IN
AMPLI FI E R
served that the dependence of the optoaeoustie voltage
output on the distance between the transducer and the
FIG. 38. Attached-transducer optoacoustic technique for sub-
laser-illuminated region is different for signals arising tracting out scattered light signals for measurements of low-
from bulk and surface absorption losses. Thus it is optical-loss solids using chopped cw lasers as sources (after
possible to separate out the two quantities from mea- Hordvik, 1977).

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 545

made of cw tunable dye lasers of sufficient power out-


o CQF& put as sources; these are both more difficult and
~ Sr F&
more expensive to build than pulsed dye lasers of the
g BaF&
type used in the present review. The need for tunable
rather than fixed lasers arises from practical situations
in which more than one trace element is present, with
2— partially overlapping absorption features.
The pulsed laser scheme, using an immersed trans-
I
E
~0'
I- ducer and gated detection, is able to measure absorp-
LU
5— tion coefficients of -10~ cm '.From this detection
2— capability we can estimate the trace constituent con-
00 )Q
B centration that can be measured. For a trace con-
stituent peak absorption coefficient of -10'-10' cm ',
o we extrapolate to a detection capability of -1—10 ppt,
CL
2-— assuming the background absorption of the solvent is
smaller than -10 ' cm '.
C)
m )0~ This last requirement
hc 5 places rather stringent limits on the detectability of
trace constituents, s.ince we have seen that for pure
water the smallest optical absorption coefficient in the
visible region is -2&&10 4cm '. Thus the use of opto-
acoustic spectroscopy for trace detection in water will
be limited to -100 ppt unless the trace constituent
2 —Cpp Co'L has sharp absorption features and/or the trace con-
I DF HF
)0-5 il
stituent can be extracted into another liquid such as
i I l

1000 2000 5000


FREQUENCY (Cm-i) CC1, that has a very low absorption coefficient in the
FIG. 39. Infrared absorption spectra of BaF&, SrF2, and CaF2 visible region.
obtained using attached-transducer, cw laser (chopped) techni-
que. Lasers used for these measurements include various V I I. 0 ISCUSSI ONS
fixed-frequency transitions from the'CO2, CO, DF, and HF
lasers (after Hordvik, 1977). A. Summary of present status
We now summarize the advantages of our pulsed OA
spectroscopy technique, as compared to the conven-
F. Application of OA spectroscopy to trace detection tional OA technique using chopped cw laser beams, gas-
in liquids
phase microphones, and lock-in detections. These ad-
The ability of optoacoustic spectroscopy to measure vantages are
ultra-small absorption coefficients in gases (Patel and (a) High detection sensitivity. We can presently de-
Kerl, 1977; Kreuzer and Patel, 1971) has found ex- tect absorbed energy of -10 "J,in a 1 cm path length.
tensive applications in the area of trace gas detection, For a laser pulsed energy of 1 mJ, this corresponds to
i.e. , pollutant measurements (see, for example, Patel, a detection limit of -10 cm ' for absorption coefficient.
1978, and Pao, 1977). For liquids we have seen that Further improvements are possible, as discussed ear-
absorption coefficients as small as 10~ cm ' can be lier.
measured using the pulsed laser, immersed transducer, (b) The use of a, pulsed laser, with its high peak
and gated detection. Applications of the technique to power (-kW), permits the ready OA detection of non-
trace detection in liquids have been reported very re- linear absorption, like multiphoton absorption or
cently by Voigtman et al. (1981). There have also been stimulated Raman scattering. However, the duty cycle
reports of the use of cw-laser-based techniques, with of the laser is very low, so that the average power is
piezoelectric transducers for such measurements with small (-10 mW). Hence spurious effects due to heating
varying degrees of success. Lahmann et al. (1977) used of the liquid and convection currents are avOided.
the cw-laser technique first described by Kohanzadeh (c) The technique of time gating, possible only with
et al. (1975) in essentially unchanged form. They de- pulsed OA detection, is a very useful method for dis-
termined trace P carotene in chloroform using the criminating against spurious OA signals like those due
5145 A transition of the argon ion laser as a source of to light scattering or window absorptions.
radiation, and established a detection limit of 9&10' (d) High-resolution studies are possible by narrow-
molecules cm ' 12 parts per trillion (ppt)]. Oda et al.
t banding the laser with intracavity etalons.
(1978) applied essentially the same technique to the (e) Absolute calibration for liquids is readily ob-
detection of trace cadmium in penicillin ochro-chloron tained.
after its decomposition and extraction into chloroform. (f) There is no complex dependence of the OA signal
Using -500 mW of argon ion laser power at 5145 A., and on thermal diffusion length, on chopping frequency,
a chopping frequency of 185 Hz, a detection limit of etc. , as in the conventional method (Rosencwaig, 1973;
0.02 ng ml ' was deduced for cadmium (14 ppt). It is Rosencwaig and Gersho, 1976; Rosencwaig, 1977;
worth noting here that for reproducible detection of Rosencwaig, 1978).
trace elements in liquids, using the Lahmann et al. or We note that high sensitivity is the major advantage
Oda et al. schemes, eventually use will have to be of the pulsed laser OA technique using immersed or

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

attached piezoelectric transducer and gated detection. applications (e.g. , for optical fibers or for laser win-
High sensitivity results from high signal strength dows) can be tested.
and/or low noise. A strong signal is possible in our d. T~ace detection. Minute quantities of foreign
case because of the use of PZT transducers in direct molecules in a liquid or solid can be detected by
contact with the liquid or solid studied. The good choosing a laser frequency that is absorbed strongly
acoustic impedance matching and fast rise time of the only by the molecules of interest. Alternately, OARS
transducer result in a high signal level. Furthermore, can also be used to map out the vibrational frequencies
negligible thermal diffusion takes place during the of the foreign molecules.
laser pulse, so maximum expansion of the irradiated e. Eorbidden transitions. Many transitions that are
sample occurs. I ow noise is possible because we are "forbidden" (e.g. , symmetry-forbidden transitions or
detecting high-frequency ultrasonic waves, so that singlet-triplet transitions) are actually weakly allowed
bandpass filters (transmitting 100 —500 kHz) can be due to various perturbations, or they can be made
used in the preamplifier circuit. This is a great ad- allowed by externally applied fields. Our high-sensi-
vantage, since the laboratory environment is very tivity technique will allow the observation of many
noisy at low frequencies (dc-10 kHz), owing to mechani- forbidden transitions.
cal vibrations, multiples of line frequencies, etc. The f. Excited states. OA spectra and lifetimes of ex-
use of time gating is crucially important in reducing cited states can be readily obtained by the pulsed OA
the noise. Furthermore, the low light-scattering geo- technique, with variable time delay of the laser OA
metry and high polish of the inner surface of the cell pulses after the production of the excited species.
also minimize noise due to light scattering. (This, however, requires the use of nonresonant trans-
ducers. )
B. Examples of further experiments g. Monolayers. Electronic absorption spectra and
The advantages of the pu&sed OA technique, which in- Raman spectra of monolayers are subjects of strong
clude high sensitivity, high noise discrimination„and current interest for our understanding of the structure
applicability to linear and nonlinear spectroscopy, can of monolayers and surfaces. The pulsed OA method
clearly be exploited in many possible future investi- (linear absorption or OARS) seems quite adaptable for
gations or applications, and we enumerate a few be- these studies. OA two-photon spectroscopy of mono-
low. layers should also be a new and interesting area of
I
a. oats-temperature OA spectroscopy. Optical research.
absorption features in high-pressure gases at low tem- h. OA microscopy. Another field of strong interest
perature have been of special interest to the astro- is OA imaging and microscopy (Busse, 1979; Pouch
physics community far the elucidation of the composi- et al. , 1979; Wong et a/. , 1978) because of its im-
tion of the atmospheres of planets such as Jupiter, portant possible applications in science and technology.
Saturn, and Uranus. For example, collision-induced Both thick samples and thin-sliced samples could be
overtone absorptions in hydrogen have been identified studied by our pulsed OA technique. OA microscopy
in the above planetary spectra. But for the third and done with different laser frequencies should provide
higher overtones in hydrogen, no laboratory data exist, valuable information, e. g. , in diagnostics of inte-
since very long path lengths, low temperatures, and grated circuit wafers or in studying biological sam-
high pressures are needed for the measurement of ples.
weak absorptions using conventional techniques. The i. Other excitation sources. We have, in this review,
OA spectroscopic technique described in the present discussed only optical sources (tunable lasers, fixed-
paper is ideally suited for making many of the desired frequency lasers, etc. ) for providingexcitation through
measurements. Further, by going to yet lower tem- the absorption of the radiation. The technique of acous-
peratures, optical absorption features in simple cryo- tic detection of absorbed energy (especially when the
genic fluids (such as H2, HD, O„N„CH, ) arising from absorbed energy is released from the medium through
the overtone absorptions can be studied. Collisional nonradiative mechanisms) is a general one and is inde-
nonradiative decays of metastable vibrationally excited pendent of the specifics of the source of excitation.
homonuclear diatomic molecules in cryogenic liquids Thus it should be possible to-carry out electron energy-
can be readily studied over a wide temperature range. loss spectroscopy as well as x-ray absorption spec-
In addition, OA spectra of solids, powders, and films troscopy using the submerged (for a liquid) or attached
can now be readily obtained over a wide range of tem- (for a solid) piezoelectric transducer technique in a
peratures. manner similar to that described in this review. In
b. Higher-order Aaman processes. The ability of particular, availability of pulsed monoenergetic elec-
OARS to measure Raman gains as small as 10 ' cm ', tron beams and pulsed x-ray sources such as the
even today, points to a direct application of the OARS synchrotron radiation sources should make the transi-
to the study of higher-order Raman processes which tion from optoacoustic spectroscopy to electron en-
have smaller Raman scattering cross sections than ergy-loss acoustic spectroscopy and x-ray absorption
those presently studied. A specific example is hyper- acoustic spectroscopy possible in the near future.
Raman scattering, which is at present considered a
very difficult experimental subject. A preliminary C. Qther QA methods
estimate indicates that GARS could make these mea- In this article, we have been mainly concerned with
surements easier in many situations. the theory and applications of tQe pulsed OA spectro-
c. Materials testing. Materials for low-attenuation scopy method, for which many novel and quantitative

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy 547

uses have been demonstrated, and will be possible; method. In view of the usefulness of the conventional
we hope that this review mill be helpful for researchers OA technique in certain applications, and the fast
planning to use this simple but useful technique. growth of interest in OA measurements in condensed
However, we do not mean to imply that conventional matter, we have collected a list of recent relevant
optoacoustic techniques or photoacoustic techniques, papers, as outlined in'the Appendix. No claim can be
as they are often called (e.g. , Harschbarger and made for completeness in this list, but we hope that
Robin, 19V3; Rosencwaig, 19V3, 19V6, 19VV, 19V8) it will provide a helpful guide to current research
are not useful. They have sensitivities a few orders inte rests.
of magnitude less than that demonstrated using the
pulsed laser technique, with immersed (in the case of
a liquid) or attached (in the case of a solid) trans- AC KNOWL E DGME NTS
ducer and gated detection. Thus if high sensitivity is
desired, the pulsed optoacoustic technique is definitely We should like to thank R. J.Kerl for his assistance
indicated. However, in studies where only low sensi- in the constructi. on of the OA cells and preamplifiers.
tivities are required, (e.g. , in totally opaque samples), We also thank V. E. Bondybey, T. D. Harris,
the conventional gas-phase microphone (i.e. , solid- . S. Geschwind, E. I. Gordon, R. A. Lemons, S. L.
gas —microphone) method can be used. Indeed, ap- McCall, G. L. Miller, P. M. Hentzepis, M. B. Robin,
plications like depth-profiling of totally opaque sam- J. Stone, M. D. Sturge, and B. A. Wilson, for their
. ples by varying the chopping frequency of the cw light useful suggestions, enlightening comments, and
beam can only be performed with the conventional f ruitful discuss ions at various stages of this work.

APPENDIX: RECENT WORK IN THE OA STUDY OF CONDENSED MATTER

Subject References

1. Conventional OA cell Blank and Wakefield, 1979


Cahen and Garty, 1979
a. Exper iment
Callis et al. , 1969
Callis, 1976
Eaton and Stuart, 1978
Gray et al. , 1977
Kanstad and Nordal, 1978
McClelland and Kniseley, 1976
McDavid et a/. , 1978
Monahan and Nolle, 1977
Nordal and Kanstad, 197V
Quimby et al. , 19VV
Bosencwaig, 1973
Bosencwaig, 1977
Sham, 1979
Tam and Wong, 1980
b. Theory Aamodt et al. , 1977
Aamodt and Murphy, 1978
Bennett and Forman, 1977
McDonald and Wetsel, Jr. , 1978
Bosencwaig and Gersho, 1976
2. Applications of OA spectroscopy
a. Quantum yields Adams et a/. , 1977
Lahmann and Ludewig, 1977
Merkle and Powell, 197V
Murphy and Aamodt, 19VV
Quimby and Yen, 19V8
Bazumova and Starobogatov, 1977
Bockley and Waugh, 1978
Starobogatov, 19VV
b. Surface absorptions Adams et al. , 1978
Hordvik and Skolnik, 197V
Kerr, 1973
Parker, 1973
Patel and Tam, 1980

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


548 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

Subject References

c. Trace analysis Lahmann et al. , 1977


Qda et al. , 1978
d. Absolute absorption coefficient Patel and Tam, 1979b
Patel et al. , 1979
Patel and Tam, 1979d
Tam et al. , 1979
Tam and Patel, 1979b
Wetsel, Jr. , and McDonald, 1977
e. Dichroism Fournier et al. , 1978
Palmer et al. , to be published
f. Others Cahen, 1978
Gray and Bard, 1978
Robin and Kuebler, 1977
Tam and Patel, 1979c
3. Optoacoustic gener ation
a. Strong absorption, surface heating Askaryan et al. , 1963
Bell and Landt,
Bonch-Bruevich et al. , 19V5
Bushanam and Barnes, 1975
Emmony et al. , 1976
Gordienko et al. , 1978
Gournay, 1966
Von Gutfeld and Budd, 1979
Sigrist and Kneubuhl, 1978
Sladky et al. , 1977
Teslenko, 1977
White, 1963
b. Weak, bulk absorption Bonch-Bruevich, 1977
Bunkin and Komissarov, 1973
Egerev and Naugol'nykh, 1977
Farrow et al. , 1978
Gorodetskii et al. , 1978
Kasoev and Lyamshev, 1977
Lyamshev and Naugol'nykh, 1976
Naugol'nykh, 197V
Westervelt and Larson, 1973
c. Electrostriction Bebchuk et al. , 1978
Brueck et al. , 1980
4. Piezoelectric OA detection Burt, 1979
Farrow et al. , 1978
Hordvik and Schlossberg, 1977
Kohanzadeh et al. , 1975
Patel and Tam, 1979a
Rentzepis and Pao, 1966
Rosencwaig et al. , 1979
Wetsel and McDonaM, 1979

5. OA microscopy Busse, 1979


Pouch et al. , 1979
Wang et al. , 1979
Wickramasinghe et al. , 1978
Wong et al. , 1978

6. OA detection of nonlinear absorptions Bonch-Bruevich et al. , 1977


Cox, 1978
Patel and Tam, 1979c
Bentzepis and Pao, 1966

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

Subject References
Pockley and Devlin, 1977
Yam and boatel, 1979a
7. R eviews Harshbarger and Robin, 1973
Hordvik, 1977
Pao, 1977
Robin, 1976
Rosencwaig, 1978
Somoano, 1978

REFERENCES

Aamodt, L. C. , J.
C. Murphy, and J. G. Parker, 1977, J. Naugol'nykh, 1978, Kvant. Elektron (Moscow) 5, 2396 (Sov.
Appl. Phys. 48, 927. J. Quantum Electron. 8, 1345).
Aamodt, L. C. , and J. C. Murphy, 1977, J. Appl. Phys. 48, Gournay, L. S., 1966, J.
Acoust. Soc. Am. 40, 1322.
3502. Gray, R. C. , and A. J.
Bard, 1978, Anal. Chem. 50, 1262.
Aamodt, L. C. , and J. C. Murphy, 1978, J. Appl. Phys. 49, Gray, R. C. , V. A. Fishman, and A. J.
Bard, 1977, Anal.
3036. Chem. 49, 697.
Adams, M. J., B. C. Beadle, G. F. Kirkbright, and K. R. Gruber, J. J.
B., R. Henderson, M. Muramoto, K. Rajnak,
Menon, 1978, Appl. Spectrosc. 32, 430. and J. G. Conway, 1966, J.
Chem. Phys. 45, 477.
Adams, M. J., J.
G. Highfield, and G. F. Kirkbright, 1977, Hale, G. M. , and M. R. Querry, 1973, Appl. Opt. 12, 555.
Anal. Chem. 49, 1850. Harshbarger, W. R. , and M. B. Robin, 1973, Acc. Chem. Res.
Askaryan, G. A. , A. M. Prokhorov, G. F. Chanturiya, and 6, 329.
G. P. Shipulo, 1963, Zh. Eksp. Teol, Flz. 44, 2180 (Sov. Hass, M. , and J. W. Davisson, 1977, J. Opt. Soc. Am. 67, 622.
Phys. —JETP 17, 1463). Hayward, R. J., and B. R. Henry, 1976, Chem. Phys. 12, 387.
Barrett, J. J., and M. J. Berry, 1979, Appl. Phys. Lett. 34, Hayward, R. J., B. R. Henry, and W. Siebrand, 1973, J. Mol.
144. Spectrosc. 46, 207.
Bebchuk, A. S., V. M. Mizin, and N. Ya Salova, 1978, Opt. Henderson, J. R. , M. Muramoto, T. M. Henderson, and J. B.
Spektrosk. 44, 158 [Opt. Spectrosc. (USSR) 44, 92]. Gruber, 1967, J. Chem. Phys. 47, 5097.
Bell, C. E. , and J. A. Landt, 1967, Appl. Phys. Lett. 10, 46. Henry, B. R. , 1977, Acc. Chem. Res. 10, 207.
Bennett, H. S., and R, A. Forman, 1977, Appj. . Opt. 16, 2834. Henry, B. R. , and W. Siebrand, 1968, J. Chem. Phys. 49,
Blank, R. E., and T. Wakefield II, 1979, Anal. Chem. 51, 50. 5369.
Bonch-Bruevich, A. M. , T. K. Razumova, and I. O. Staroboga- Heritage, J. P. , J. G. Bergman, A. Pinzuk, and J. M. Wor-
tov, 1975, Pis ma Zh. Tech. Fiz. 1, 65 (Sov. Phys. — lock, 1979, "Picosecond Raman gain spectroscopy of cyanide
Tech. Phys. Lett. 1, 26). on silver" (unpublished).
Bonch-Bruevich, A. M. , T. K. Razumova, and I. O. Staroboga- Hordvik, A. , 1977, Appl. Opt. 16, 2827.
tov, 1977, Opt. Spektrosk. 42, 82 [Opt. Spectrosc. (USSR) Hordvik, A. , and H. Schlossberg, 1977, Appl. Opt. 16, 101.
42, 45]. Hordvik, A. , and L. Skolnik, 1977, Appl. Opt. 16, 2919.
Brilmyer, G. H. , A. Fujishima, K. S. V. Santhanam, and A. J. Hu, C. , and J.
R. Whinnery, 1973, Appl. Opt. 12, 72.
Bard, 1977, Anal. Chem. 49, 2057. Hulbert, E. O. , 1945, J.
Opt. Soc. Am. 35, 698.
Brueck, S. R. J., H. Kildal, and L. J. Belanger, 1980, Opt. Irvine, W. M. , and J. B. Pollack, 1968, Icarus 8, 324.
, Commun. 34, 199. Kanstad, S. O. , and P. E. Nordal, 1978, Opt. Commun. 26,
Bunkin, F. V. , and V. M. Komissarov, 1973, Akust. Zh. 19, 367.

305 (Sov. Phys. Acoust. 19, 203). Kasoev, S. G. , and L. M. Lyamshev, 1977, Akust. Zh. 23,
J.
Burt, J. A. , 1979, Acoust. Soc. Am. 65, 1164. —
890 (Sov. Phys. Acoust. 23, 510).
Bushanam, G. S. , and F. S. Barnes, 1975, J.
Appl. Phys. 46, Kerr, E. L. , 1973, Appl. Opt. 12, 2520.
2074. Kohanzadeh, Y. , J. R. Whinnery, and M. M. Carroll, 1975, J.
Busse, G. , 1979, in Technical Digest, Topical meeting in Acoust. Soc. Am. 57, 67.
photoacoustic spectroscopy, Ames, Iowa (unpublished). Kopelevich, O. V. , 1976, Opt. Spektrosk. 41, 666 IOpt. Spec-
Cahen, D. , 1978, Appl. Phys. I ett. 33, 810. trosc. (USSR) 41, 391].
Cahen, D. , and H. Garty, 1979, Anal. Chem. 51, 1865. Kreuzer, L. B., 1971, J. Appl. Phys. 42, 2934.
Callis, J. B., 1976, J. Res. , Nat. Bur. Stand. 80A, 413. Kreuzer, L. B., and C. K. N. Patel, 1971, Science 173, 45.
Callis, J. B., M. Gouterman, and J. D. S. Danielson, 1969, Lahmann, W. , and H. J. Ludewig, 1977, Chem. Phys. Lett.
Rev. Sci. Instrum. 40, 1599. 45, 177.
Cox, D. M. , 1978, Opt. Commun. 24, 336. Lahmann, W. , H. J. Ludewig, and H. Welling, 1977, Anal.
Eaton, H. E. , and J. D. Stuart, 1978, Anal. Chem. 50, 587. Chem. 49, 549.
Egerev, S. V. , and K. A. Naugol'nykh, 1977, Akust. Zh. 23, Landau, L. D. , and E. M. Lifshitz, 1959, Eluid Mechanics

738 (Sov. Phys. Acoust. 23, 422). translated from the Russian by J. B. Sykes and W. H. Reid
Emmony, D. C. , M. Siegrist, and F. K. Kneubuhl, 1976, Appl. (Pergamon, New York), Chap. VIII.
Phys. Lett. 29, 547. Levine, B. F., and C. G. Bethea, 1979, "Ultrahigh sensitivity
Farrow, M. M. , R. K. Burnham, M. Auzanneau, S. L. Olsen, stimulated Raman gain spectroscopy" (unpublished).
N. Purdie, and E. M. Eyring, 1978, Appl. Opt. 17, 1093. Long, M. E. , R. L. Swofford, and A. C. Albrecht, 1976, J.
Fournier, D. , A. -C. Boccara, and J. Badoz, 1978, Appl. Phys. Chem. Phys. 65, 179.
Lett. 32, 640. Lyamshev, L. M. , and K. A. Naugol'nykh, 1976, Akust. Zh.
Gordienko, V. M. , A. B. Reshilov, and V. I, Shmal'gauzen, —
22, 625 (Sov. Phys. Acoust. 22, 354).

1978, Akust. Zh. 24, 132 (Sov. Phys. Acoust. 24, 73). McClelland, J. F., and R. N. Kniseley, 1976, Appl. Opt. 15,
Gorodetskii, V. S. , S. V. Egerev, I. B. Esipov, and K. A. 2658; 2967.

Rev. Mod. Phys. , Vol. 53, No. 3, July 1981


550 C. K. N. Patel and A. C. Tam: Pulsed optoacoustic spectroscopy

McDavid, J. M. , K. L. Lee, S. S. Yee, and M. A. Afromowitz, Robin, M. B., and N. A. Kuebler, 1975, J. Am. Chem. Soc. 97,
1978, J. Appl. Phys. 49, 6112. 4822.
McDonald, F. A. , and G. C. Wetsel, Jr. , 1978, J. Appl. Phys. Robin, M. B., and N. A. Kuebler, 1977, J. Chem. Phys. 66,
49, 2313. 169.
McQueen, D. H. , 1979, J. Phys. D 12, 1673. Bockley, M. G. , and J. P. Devlin, 1977, Appl. Phys. Lett. 31,
Merkle, L. D. , and R. C. Powell, 1977, Chem. Phys. Lett. 46, 24.
303. Rockley, M. G. , and K. M. Waugh, 1978, Chem. Phys. Lett.
Monahan, E. M. , and A. W. Nolle, 1977, J. Appl. Phys. 48, 54, 597.
3519. Bosencwaig, A. , 1973, Opt. Commun. 7, 305.
Moses, E. I., and C. L. Tang, 1977, Opt. Lett. 1, 115. Rosencwaig, A. , 1975, Anal. Chem. 47, 592A.
Naugol'nykh, K. A. , 1977, Akust. Zh. 23, 171 (Sov. Phy's. — Rosencwaig, A. , 1977, Rev. Sci. Instrum. 48, 1133.
Acoust. 23, 98). Bosencwaig, A. , 1978, Adv. Electron. and Electron Phys. 46,
Nechaev, S. Yu. , and Yu. N. Ponomarev, 1975, Kvant. Elek- 207.
tron. (Moscow) 2, 1400 (Sov. J. Quantum Electron. 5, 752). Rosencwaig, A. , and A. Gersho, 1976, J. Appl. Phys. 47, 64.
Nelson, E. T. , and C. K. N. Patel, 1981, Opt. Lett. (in press). Bosencwaig, A. , J. L. Opsal, and J. F. Scheimer, 1979,
Nordal, P. E., and S. O. Kanstad, 1977, Opt. Commun. 22, "Photoacoustic Stress, " in Technical Digest, Topical Meeting
185. on Photoacoustic Spectroscopy, Ames, Iowa (unpublished).
Oda, S. , T. Sawada, and H. Kamada, 1978, Anal. Chem. 50, Sam, C. L. , and M. L. Shand, 1979, Opt. Commun. 31, 174.
865. Shaw, R. W. , 1979, Appl. Phys. Lett. 35, 253.
Palmer, R. A. , J. C. Roark, and J. C. Robinson, "Photo- Sigrist, M. K. , and F. K. Kneubuhl, 1978, J. Acoust. Soc. Am.
acoustic detection of natural circular dichroism in crystal- 64, 1652.
line transition metal complexes, " in ACS Symposium entitled Sladky, P. , R. Danielius, V. Sirutkaitis, and M. Boudys, 1977,
"The Stereochemistry of Optically Active Compounds of Czech. J. Phys. B27, 1075.
Transition Metals, " (to be published). Somoano, B. B., 1978, Angew. Chem. Int. Ed. Engl. 17, 238.
Pao, Y. H. , 1977, editor, Optoacolstic Spectroscopy and De- Starobogatov, I. O. , 1977, Opt. Spektrosk. 42, 304 [Opt. Spec-
tection (Academic, New York). trosc. (USSR) 42, 172].
Parker, J. G. , 1973, Appl. Opt. 12, 2974. Stone, J., 1978, Appl. Opt. 17, 2876.
Parker, J. G. , and D. N. Ritke, 1973, J. Chem. Phys. 59, Sullivan, S. A. , 1963, J. Opt. Soc. Am. 53, 962.
3713. Tarn, A. C. , and C. K. N. Patel, 1979a, Nature 280, 304.
Patel, C. K. N. , 1978, Science 202, 157. Tam, A. C. , and C. K. N. Patel, 1979b, Appl. Opt. 18, 3348.
Patel, C. K. N. , and R. J. Kerl, 1977, Appl. Phys. Lett. 30, Tam, A. C. , and C. K. N. Patel, 1979c, Appl. Phys. Lett. 35,
578. 843.
Patel, C. K. N. , and A. C. Tam, 1979a, Appl. Phys. Lett. 34, Tam, A. C. , C. K. N. Patel, and R. J. Kerl, 1979, Opt. Lett.
467. 4, 81.
Patel, C. K. N. , and A. C. Tam, 1979b, Chem. Phys. Lett. 62, Tam, A. C. and C. K. N. Patel, 1980, Opt. Lett. 5, 27.
511. Tam, A. C. and Y. H. Wong, 1980, Appl. Phys. Lett. 36, 471.
Patel, C. K. N. , and A. C. Tam, 1979c, Appl. Phys. Lett. 34, Teslenko, V. S., 1977, Kvant. Elektron. (Moscow) 4, 1732
760. (Sov. J. Quantum Electron. 7, 981).
Patel, C. K. N. , and A. C. Tam, 1979d, Nature 280, 302. Voigtman, E., A. Jurgensen, and J.
Winefordner, 1981, Anal.
Patel, C. K. N. , and A. C. Tarn, 1979e, "Pulsed opto-acoustic Chem. (in press).
spectroscopy of liquids, " in I xoceedings of the 4th Inte~a- Von Gutfeld, B-. J., and H. F. Budd, 1979, App1. . Phys. Lett.
tional Conference on Laser Spectroscopy, Aottach-Egest, 34, 617.
West Germany. Wang, T. T.,J. M. Mcoavid, and S. S. Yee, 1979, Appl. Opt.
Patel, C. K. N. , and A. C. Tam, 1980, Appl. Phys. Lett. 36, 7. 18, 2354.
Patel, C. K. N. , and A. C. Tam, and R. Kerl, 1979, J. J. Westervelt, P. J., and R. S. Larson, 1973, J. Acoust. Soc.
Chem. Phys. 71, 1470. Am. 54, 121.
Pouch, J. J.
, R. L. Thomas, Y. H. Wong, Schuldies, and J. Wetsel, G. C. , Jr. , and F. A. McDonald, 1977, Appl. Phys.
M. Srinivasan, 1979, ".Scanning photoacoustic microscopy, " Lett. 30, 252.
in Technical Digest, Topical Meeting on Photoacoustic Spec- Wetsel, G. C. , Jr. , and F. A. McDonald, 1979, "Piezoelectric
troscopy, Ames, Iowa {unpublished). "
detection of photoacoustic signals, in Technical Digest, the
Querry, M. R. , P. A. Cary, and'R. C. Waring, 1978, Appl. Topical Meeting on Photoacoustic Spectroscopy, Ames,
Opt. 17, 3587. Iowa (unpublished).
Quimby, R. S., P. M. Seizer, and W. M. Yen, 1S77, Appl. White, R. M. , 1963, J.
Appl. Phys. 34, 3559.
Opt. 16, 2630. Wickramasinghe, H. K. , B. C. Bray, V. Jipson, C. E. Quate,
Quimby, R. S. , and W. M. Yen, 1978, Opt. Lett. 3, 181. and J.
R. Salcedo, 1978, Appl. Phys. Lett. 33, 923.
Razumova, T. K. , and I. O. Starobogatov, 1977, Opt. Spekrosk. Wilson, E. B., Jr. , 1S34, Phys. Bev. 45, 706.
42, 489 [Opt. Spectrosc. (USSR) 42, 274]. Wong, Y. H. , R. L. Thomas, and G. F. Hawkins, 1978, Appl.
Beddy, K. V. , R. G. Bray, and M. J.
Berry, 1978, in Advances Phys. Lett. 32, 538.
in Laser Chemistry, edited by A. Zewail (Springer-Verlag, Wunsch, L. , H. J. Neusser, and E. W. Schlag, 1975a, Chem.
Berlin), p. 48. Phys. Lett. 31, 433.
Rentzepis, P. M. , and Y. H. Pao, 1966, J. Chem. Phys. 44, Wunsch, L. , H. J. Neusser, and E. W. Schlag, 1975b, Chem.
2931. Phys. Lett. 32, 210.
Robin, M. B., 1976, J. Lumin. 13, 131.

Rev. Mod. Phys. , Vol. 53, No. 3, July 't981

Вам также может понравиться