Вы находитесь на странице: 1из 12

AIP ADVANCES 1, 022154 (2011)

Thermal lens spectrometry: Optimizing amplitude and


shortening the transient time
Rubens Silva,
1,2
Marcos A. C. de Ara ujo,
1
Pedro Jali,
1
Sanclayton G. C.
Moreira,
2
Petrus Alcantara Jr.,
2
and Paulo C. de Oliveira
1,a
1
Departamento de Fsica, Centro de Ci encias Exatas e da Natureza, Universidade Federal da
Paraba, Jo ao Pessoa, 58051-970, Paraba, Brazil
2
Faculdade de Fsica, Instituto de Ci encias Exatas e Naturais, Universidade Federal do
Par a, Bel em, 66093-020, Brazil
(Received 28 April 2011; accepted 19 June 2011; published online 29 June 2011)
Based on a model introduced by Shen et al. for cw laser induced mode-mismatched
dual-beam thermal lens spectrometry (TLS), we explore the parameters related with
the geometry of the laser beams and the experimental apparatus that inuence the
amplitude and time evolution of the transient thermal lens (TL) signal. By keeping
the sample cell at the minimum waist of the excitation beam, our results show
that high amplitude TL signals, very close to the optimized value, combined with
short transient times may be obtained by reducing the curvature radius of the probe
beam and the distance between the sample cell and the detector. We also derive an
expression for the thermal diffusivity which is independent of the excitation laser
beam waist, considerably improving the accuracy of the measurements. The sample
used in the experiments was oleic acid, which is present in most of the vegetable oils
and is very transparent in the visible spectral range. Copyright 2011 Author(s). This
article is distributed under a Creative Commons Attribution 3.0 Unported License.
[doi:10.1063/1.3609966]
I. INTRODUCTION
Since the rst reports on the thermal lens effect in 1964 and 1965,
13
a great number of ap-
plications exploring this effect was developed, including the measurement of very low absorption
coefcients of transparent liquids, such as water, ethanol, etc..
36
Today this technique is known as
Thermal Lens Spectrometry (TLS). Another very important application is in the determination of
thermal diffusivities.
7, 8
Other studies such as the measurement of quantum efciency and dimeriza-
tion equilibria of laser dyes were also reported.
9, 10
In TLS experiments, a Gaussian TEM
00
excitation
beam is partially absorbed by the sample, creating a refractive index gradient in the radial direction.
When a probe beam travels along the heated path, phase differences appear between points of the
wavefront with different radial distances. These phase differences distort the wavefront, causing a
lenslike effect.
The rst systematic study of the inuence of beam geometric positions in dual-beam thermal
lens experiment was done by Berthoud et al..
11
They showed that TL signal is maximum when the
excitation beam is focused in the sample cell. Further, when the distance between the focus of the
probe beam and the cell is changed, the signal amplitude pass through a maximum at a distance
approximately

3 Z
c
, where Z
c
is the Rayleigh parameter of the probe beam. Recently, Marcano
et al. optimized the amplitude of TL signals, approaching the theoretical limit value, by expanding
and collimating the probe beam.
5
With this conguration, the TL amplitude was enhanced by 40%
but, on the other hand, the TL transient time increased almost two orders of magnitude. However, it is
a
Electronic mail: pco@sica.ufpb.br
2158-3226/2011/1(2)/022154/12 C
Author(s) 2011 1, 022154-1
022154-2 Silva et al. AIP Advances 1, 022154 (2011)
worth noting that all the above mentioned works investigated only the optimization of the amplitude
of the TL signals.
In this Letter, we present theoretical and experimental results showing that it is possible, for
certain geometrical congurations of the laser beams and experimental arrangement, to obtain high
amplitude TL signals with short transient times. The sample used in the experiments was oleic acid,
which is present in most of the vegetable oils and is very transparent in the visible spectral range.
II. THEORY
The rst models describing the thermal lens effect considered the temperature rise in the sample
as a parabolic function of r, and the thermal lens could be treated as perfect thin lens.
2, 12
Taking the
aberrant nature of the TLinto account Sheldon et al.
13
proposed an aberrant thermal lens model for the
single beam conguration. About ten years later, a model for cw laser induced mode-mismatched
dual-beam thermal lens spectrometry (TLS) was introduced by Shen et al..
14, 15
These models
take into account the Fresnel diffraction theory to the imaging of the probe beam at the detector
plane. A model with a different approach was reported by Bialkowski and Chartier,
16
where they
calculated cumulative electric-eld phase shifts produced by a series of Gaussian photothermal-
induced refractive-index perturbations. In this work we will base our discussions on the results
obtained by Shen et al..
The scheme of the laser beams is shown in Fig. 1(a) and 1(b). Fig. 1(a) shows the traditional
scheme, with the sample positioned at the minimum waist of the excitation beam, and a probe beam
which minimum waist is at a distance Z
1
from the sample cell. Fig. 1(b) shows a modied version
of the traditional scheme, recently introduced by Marcano et al..
5
The modication introduced in
Ref. 5 is that the probe beam is collimated (R
1p
= ) and the beam radius at the sample cell (
1p
)
is increased.
According to the model introduced by Shen et al., the intensity of the TL signal at the center of
the probe beam, in the plane of the detector, is given by
I (t ) = I (0)

1

2
arctan

2mV

V
2
+(1 +2m)
2

t
c
/2t + V
2
+1 +2m

2
(1)
where,
m =

2
1p

2
e
, (2)
=
P
e
L

p
k

dn
dT

, (3)
and,
V =

2
1p

1
R
1p
+
1
Z
2

. (4)
All parameters appearing in Eqs. (1)(4) are described in Table I. A few other parameters will be
described as they appear in the text.
The Eq. (1) is widely used to t the experimental data of the transient TL signals, giving the
values of the both, and t
c
parameters. The latter is being used to nd the thermal diffusivity,
obtained by
D =

2
e
4t
c
(5)
A. The amplitude of the TL signal
In our analysis, instead of working with the beam parameters
e
,
1p
,
0p
, and the confocal
parameters, we prefer to work with the parameters m and V described by Eq. (2) and (4). After nding
the m and V that maximize the TL signal, we analyze how they can be adjusted experimentally.
022154-3 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 1. Schema of the geometric position of the beams in a mode-mismatched TL experiment.(a) Traditional scheme,
introduced by J. Shen et al., and (b) scheme introduced by A. Marcano et al..
With the intensity of the light at the detector given by Eq. (1), the steady-state TL fractional
signal amplitude may be dened as
14
S =
I (0) I ()
I (0)
. (6)
After substitution of Eq. (1) into the above equation we get
S = 1

1

2
tan
1

2mV
1 +2m + V
2

2
. (7)
022154-4 Silva et al. AIP Advances 1, 022154 (2011)
TABLE I. Description of the symbols used in Eqs. (1)(4) and in Fig. 1.
Symbol Description
I(t) Intensity of the center of the probe laser at the detector

e
minimum excitation beam waist

1p
probe beam waist at the sample cell position
R
1p
curvature radius of the probe beam at the cell position
P
e
excitation laser power

p
probe beam wavelength
k thermal conductivity
absorption coefcient at the excitation beam wavelength
L sample thickness
dn/dT temperature coefcient of the refractive index
t
c
characteristic thermal lens time constant
D thermal diffusivity
Z
1
distance from the probe beam waist to the sample cell
Z
2
distance from the sample cell to the detector
From this equation, we may see that for a xed excitation power (xed ), the maximum
amplitude of the TL signal will be obtained when the argument of the arc tangent function is
maximum. I (t ) is a function of the two geometrical parameters m and V, and both depend on

1p
. For the sake of simplicity, we will analyze them as independent parameters. By analyzing
the dependence of S on V, we veried that the amplitude of the TL signal has a maximum when
V = V
opt
, with V
opt
given by
V
opt
=

1 +2m, (8)
where the negative solution corresponds to the situation where the sample is positioned before the
probe beam focus (R
1p
< 0).
An analysis of the dependence of S on m, keeping V at the maximum value, reveals that S
grows as m goes to innity, approaching a limit, given by
S
max
= 1

1

4

2
(9)
Experimentally m may reach very large values by increasing the relation between the probe and
excitation beams waists at the sample position. Marcano et al.
5, 19
optimized the TL signal, approach-
ing the theoretical limit value, by expanding and collimating the probe beam. In their experiments
the diameter of the probe beam at the sample position was about 6 mm, which corresponds to a value
of m of the order of 10000. From their data, V is estimated to be only about 30, while the optimized
value of V is around 120. This will not affect too much the amplitude of the TL signal. However,
as we will see in the next section, it will greatly affect the time necessary to reach the steady-state
regime.
In Fig. 2 we show the amplitude of the TL signals for various values of the parameter m as
a function of V. It is noticed that larger values of m gives rise to larger amplitudes of TL signals,
but this behavior saturates for m around a few hundreds. For example, the maximum TL signal for
m = 300 is very close to the TL signal for m = 10000, differing by only 3% from each other at
V = 30.
B. The time dependence of the TL signal
Another very important parameter in an experiment is how fast a physical phenomenon is, and
how long a measurement will take. In a single beam TLS, t
c
is the characteristic time of transient
signal. But in the dual-beam mode-mismatched TLS, it is well known that the larger the probe
022154-5 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 2. Amplitude of the TL signal divided by , for = 0.1, for various values of the parameter m (1, 5, 10, 100, 300,
10000) as a function of V.
beam diameter is, the longer the signal takes to reach the steady-state regime, and consequently, the
longer will be the measurement time. In order to optimize the TL signal, one must to increase the
amplitude without affecting too much the measurement time. To make a quantitative analysis of the
time expended in a dual-beam mode-mismatched TL experiment we dene a quantity that we call
the half-amplitude time (t
h
), obtained by
I (t
h
) =
I (0) + I ()
2
. (10)
Expanding Eq. (1), and dismissing terms proportional to
2
, we get
t
h
=
1
2

V
2
+(1 +2m)
2

2
+4m
2
V
2

4m
2
V
2

2
+4m
2
V
2

t
c
, (11)
where,
= V
2
+1 +2m. (12)
Since we have dismissed the terms proportional to
2
in the derivation of Eq. (11), a special
care must be taken when comparing t
h
given by this equation with the experimental results in cases
where is large ( 0.1). From numerical simulations of Eq. (1), we observed that a correction
factor approximately given by (1 /3) in the value of t
h
must be applied. For instance, in our
experiments an average = 0.132 was obtained, which gives a correction factor equal to 0.956 to
t
h
.
For V = V
opt
one may show that
t
h
=

1/2 +m
V
opt

t
c
(13)
022154-6 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 3. Half-amplitude time for the TL signal for various values of the parameter m (1, 10, 100, 1000, 10000) as a function
of V.
By plotting t
h
as a function of V (see Fig. 3) we veried that, when m >> 1, we may substitute
V
opt
by V in a wide interval around V
opt
, and Eq. (13) may be written as
t
h

m
V

t
c
(14)
This result shows that, by keeping m xed, we may decrease the half-amplitude time of the
transient TL signals by increasing the parameter V. In practice, this may be performed by reducing
the curvature radius of the probe beam(R
1p
), and also by reducing the distance (Z
2
) fromthe detector
to the sample cell.
The above results may be explained as follows: An increment in the mode-mismatching param-
eter m means that the probe beam will probe a larger area of the sample and the heat will take a
longer time to propagate to the border and to reach the steady-state regime. Consequently the TL
transient time will be longer. On the other hand, the amplitude of the TL signal will be larger due
to the increment in the phase difference. The parameter V is related to the initial phase difference
between the wave passing at the center and the wave passing at the probe beam waist. The larger is
the initial phase difference the faster will be the transient regime.
If instead of t
c
we use t
h
to estimate the thermal diffusivity, we may verify that a very important
simplication occurs. By substituting m and V given by Eqs. (2) and (4), into Eq. (14), and then into
Eq. (5), the result is
D =

1
R
1p
+
1
Z
2

t
h

1
(15)
which is independent of the excitation beamwaist
e
. Since the precise measurement of the excitation
beam waist usually takes a long time to be done and there is an error associated with its estimation,
the use of Eq. (15) will considerably reduce the measurement time, giving more reliable results than
Eq. (5). Another important issue is that t
c
is obtained by a tting process involving Eq. (1), that
depends on
e
. On the other hand t
h
may be obtained by a direct inspection of the experimental
curve.
022154-7 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 4. Experimental setup for thermal lens spectroscopy. M - mirror, P - polarizer prism, BS - nonpolarizer beam-splitter
prism, S - sample, PH - pinhole, PD - photodetector, L
1
( f
1
= 20 cm), and L
2
( f
2
= 10 cm) lenses.
III. THE EXPERIMENTS
We performed a set of experiments in order to demonstrate the above results. The experimental
setup is shown in Fig. 4. The excitation laser was an Ar
+
laser operating at a wavelength of 514 nm,
and focused by the lens L
1
, which has focal length f
1
= 20 cm. The excitation power was about
5 mW. The probe laser was a He:Ne laser emitting at 632.8 nm, and aligned counter-propagating
with the excitation laser. With this arrangement, we guarantee that the laser beams are completely
superposed throughout the sample extension, which thickness was 2 mm. Two polarizers with
orthogonal orientations were used at the output of each laser in order to eliminate light from one
laser to enter the cavity of the other. The sample material was oleic acid, which is transparent to the
probe laser but has a small absorption at the excitation laser wavelength.
The experiment was divided into two sets of measurements. The rst set of measurements was
done with the scheme of Fig. 1(a), where the probe beam was focused by a 10 cm focal length
lens (L
2
) and the sample cell was positioned at the minimum waist of the excitation laser (
e
=
36 m), that was about 20 cm from this lens. With this conguration the probe laser beam diameter
at the sample position was about 1.2 mm (
1p
= 613 m), which corresponds to the same diameter
of the original He:Ne laser beam, and R
1p
was 10.7 cm for this conguration. The second set of
measurements was done with the scheme of Fig. 1(b), where the lens L
2
was replaced by a pair of
collimating lenses, that were adjusted to maintain the same diameter of the rst set of measurements.
R
1p
was 1605 cm for this conguration.
The beam radii
e
,
1p
, and
0p
, and the Rayleigh parameters of the excitation and probe
beams, Z
ce
and Z
cp
, respectively, were measured by the knife-edge technique.
17, 18
For the rst
set of measurements R
1p
was calculated from R
1p
= (Z
2
1
+ Z
2
cp
)/Z
1
. For the second set, it was
calculated by a linear regression of the data of the beam radius taken at several positions around the
sample position. For both sets of measurements, the distance Z
2
, from the detector to the sample
cell was varied, and we have studied the thermal lens signal as a function of the parameter V that
is related with Z
2
through Eq. (4). Tables II and III show the equivalence between Z
2
and V for
each measurement. The TL signal was detected by a silicon photodiode, in front of which there was
a pinhole with a diameter of 200 m. The data acquisition was made by a Tektronix DPO 3012
digital oscilloscope, synchronized with the switching-on of the excitation laser, and an average of
128 measurements was made.
Fig. 5 shows transient TL signals obtained with the detector placed at different distances from
the sample cell. In this experiment, we have used the scheme of Fig. 1(b), where the probe beam
022154-8 Silva et al. AIP Advances 1, 022154 (2011)
TABLE II. Parameter V as a function of Z
2
, obtained from Eq. (4) for the geometry of Fig. 1(a), with
e
= 36 m,
1p
=
613 m and R
1p
= 10.7 cm.
Z
2
(cm) 145.7 28.7 21.7 15.5 11.2
V 15.6 19.9 21.7 24.5 28.4
TABLE III. Parameter V as a function of Z
2
, obtained from Eq. (4) for the geometry of Fig. 1(b), with
e
= 36 m,
1p
=
613 m and R
1p
= 1605 cm.
Z
2
(cm) 207.5 140.5 105.0 70.0 50.5 29.0 20.0 15.0
V 1.02 1.44 1.89 2.78 3.81 6.55 9.44 12.6
FIG. 5. Normalized transient TL signals for m = 290, and various values of the parameter V.
is collimated. The upper curve was taken with the detector at a distance of 207.5 cm from the
sample cell, which corresponds to V = 1.02, while the lower curve was taken at a distance of only
15 cm, which corresponds to V = 12.6 (see Tables II and III). From Fig. 5 one may notice that an
increasing of the amplitude and shortening of the transient time occurs simultaneously as we reduce
the distance of the detector from the sample cell. Since the probe beam diameter is of the order of
1 mm, and the shortest distance of the sample to the detector of about 15 cm, we believe that the far
eld approximation is still valid for our experiments.
The procedure for obtained the half-amplitude time from the thermal lens experimental data
consists in measuring both the intensity I (0) immediately before the opening of the shutter that
modulates the excitation laser and the intensity I () after stabilization of the transient regime, as
shown in Fig. 6. We then calculate the average between the intensities I (0) and I () and look
at which time it occurs. This is what we call the half-amplitude time (t
h
). This procedure is not
automated but it is very simple and accurate and do not require any tting. A photodiode is used to
detect the switching on of the excitation beam and dene the origin of the time axis.
The experimental results, showing the increase of the amplitude of the TL signal, as well as the
shortening of the transient time with V are shown in Figs. 7 and 8, respectively. Due to geometrical
022154-9 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 6. Measurement of the half-amplitude time of an experimental thermal lens signal.
limitations, the data points for V between 1 and 15 were obtained using the scheme of Fig. 1(b),
while the data points for V > 15 were taken using the scheme of Fig. 1(a). The probe beam radius
at the sample cell
1p
was set with the same value for both schemes. The experiments conrmed the
theoretical predictions for the amplitude and transient times.
In order to verify the applicability of Eq. (15) for the measurement of thermal diffusivities, we
made another set of measurements where we xed all the geometrical parameters, with exception of
the excitation laser beam diameter which increases with wavelength, as shown in Table IV. The Ar
+
laser was tuned at four different lines (476 nm, 488 nm, 496 nm, and 514 nm), while its waist varied
from 33 m to 38 m. For these measurements
1p
= 719 m, R
1p
= 13.2 cm and Z
2
= 31.7 cm.
With this conguration, we have m > 400, V = 27.6 and, during the experiments the exciting laser
power was adjusted such that <0.1, which guarantees that errors introduced by our approximations
may be neglected in comparison with the experimental ones. These measurements showed that
the values of thermal diffusivities, calculated with Eq. (15), stayed practically constant around
10.2 10
4
cm
2
/s, a value that agrees with that obtained by Dadarlat et al. (10.3 10
4
cm
2
/s)
using a photopyroelectric method.
20
The accuracy of the thermal diffusivity obtained by the use of
Eq. (15) is affected only by the accuracy of the measurement of R
1p
, Z
2
, and t
h
, which are estimated
to within 1%. On the other hand, an accurate estimation of the excitation beam waist
e
demands a
long time and special equipments. In this work no special care was taken to the measurement of the
excitation beam waist and an estimated error of its measurement was about 5%. It is worth noting
that the parameter m, appearing in Eq. (1), depends on the square of
e
, making it very sensitive to
uncertainties in
e
. These uncertainties will affect the accuracy of the characteristic time t
c
obtained
by the tting process and, consequently, the accuracy of the thermal diffusivity. An estimation of
the experimental error, using our method, gives a value around 1%, where errors based on the
traditional tting of the experimental thermal lens curve with Eq. (1) typically gives an error of about
5%.
IV. CONCLUSION
We have shown, theoretically and experimentally, that high amplitude TL signals, very close
to the optimized value, combined with short transient times, may be obtained by reducing the
022154-10 Silva et al. AIP Advances 1, 022154 (2011)
FIG. 7. Amplitude of the thermal lens signal as a function of the parameter V. Data points for V up to about 15 were obtained
in the collimated probe beam conguration, and above 15 were obtained in the focused probe beam conguration. The solid
line is the result given by Eq. (7) with m = 290 and = 0.132.
FIG. 8. Half-amplitude time of the thermal lens signal as a function of the parameter V. Data points for V up to about 15 were
obtained in the collimated probe beam conguration, and above 15 were obtained in the focused probe beam conguration.
The dashed line is the result given by Eq. (11), and the solid line is the approximated result given by Eq. (13), both with
m = 290 and t
c
= 3.08 ms.
022154-11 Silva et al. AIP Advances 1, 022154 (2011)
TABLE IV. Thermal diffusivities obtained from Eq. (15) for several pumping laser lines. Thermal diffusivities are
independent of pumping beam waist when calculated from Eq. (15).

e
(nm)
e
(m) t
h
(ms) D(10
4
cm
2
/s)
476 33 47.4 10.2
488 35 46.7 10.3
496 37 47.1 10.2
514 38 47.9 10.1
curvature radius of the probe beam and the distance between the sample cell and the detector. We
noticed that, by expanding the probe beam, the amplitude of the TL signal increases together with
the corresponding transient time. We also noticed that the focused probe beam conguration gives
rise to shorter transient times than the collimated probe beam conguration, for the same beam
diameter. The denition of a new parameter, that we called half-amplitude time or t
h
, was essential
for the model analysis presented here. We also showed that the measurement of t
h
is more accurate,
precise and easier to obtain than the characteristic time t
c
, and that it may be used to estimate the
thermal diffusivity of the samples. The major advantage of this method is that thermal diffusivity
will not depend on the excitation beam diameter for large m. We have successfully demonstrated
this applicability by estimating the thermal diffusivity of oleic acid, which agreed with the literature
value.
ACKNOWLEDGMENTS
The authors would like to thank Dr. M. L. Lyra for helpful discussions. This work was supported
by FAPESPA, FINEP, CNPq and CAPES, Brazilian agencies.
1
J. P. Gordon, R. C. C. Leite, R. S. Moore, S. P. S. Porto, and J. R. Whinnery, Long transient effects in lasers with inserted
liquids samples, Bull. Am. Phys. Soc. 9, 501 (1964).
2
J. P. Gordon, R. C. C. Leite, R. S. Moore, S. P. S. Porto, and J. R. Whinnery, Long transient effects in lasers with inserted
liquid samples, J. Appl. Phys. 36, 3-8 (1965).
3
R. C. C. Leite, R. S. Moore, and J. R. Whinnery, Low absorption measurement by mean of the thermal lens effect using
a He:Ne laser, Appl. Phys. Lett. 5, 141-143 (1964).
4
D. Solimini, Loss measurement of organic materials at 6328 , J. Appl. Phys. 37, 3314-3315 (1966).
5
A. Marcano, H. Cabrera, M. Guerra, R. A. Cruz, C. Jacinto, and T. Catunda, Optimizing and calibrating a mode-mismatched
thermal lens experiment for low absorption measurement, J. Opt. Soc. Am. B 23, 1408-1413 (2006).
6
R. A. Cruz, A. Marcano, C. Jacinto, and T. Catunda, Ultrasensitive thermal lens spectroscopy of water, Opt. Lett. 34,
1882-1884 (2009).
7
M. L. Baesso, J. Shen, and R. D. Snook, Time-resolved thermal lens measurement of thermal diffusivity of sodalime
glass, Chem. Phys. Lett. 197, 255-258 (1992).
8
M. Benitez, A. Marcano, N. Melikechi, Thermal diffusivity measurement using the mode-mismatched photothermal lens
method, Opt. Engineering 48, 043604(1-8) (2009).
9
M. Fischer, J. Georges, Fluorescence quantum yield of rhodamine 6G in ethanol as a function of concentration using
thermal lens spectrometry, Chem. Phys. Lett. 260, 115-118 (1996).
10
M. Fischer, J. Georges, Use of thermal lens spectrometry for the investigation of dimerization equilibria of rhodamine 6G
in water and aqueous micellar solutions, Chem. Phys. Lett. 260, 115-118 (1996).
11
T. Berthoud, N. Delorme, P. Mauchien, Beam geometry optimization in dual-beam thermal lensing spectrometry, Anal.
Chem., 1985 57, 1216-1219 (1985).
12
C. Hu and J. R. Whinnery, New Thermooptical Measurement Method and a Comparison with Other Methods, Appl.
Opt. 12, 72-79 (1973).
13
S. J. Sheldon, L. V. Knight, and J. M. Thorne, Laser-induced thermal lens effect: a new theoretical model, Appl. Opt.
21, 1663-1669 (1982).
14
J. Shen, R. D. Lowe, and R. D. Snook, A model for cw laser induced mode-mismatched dual-beam thermal lens
spectrometry, Chem. Phys. 165, 385-396 (1992).
15
J. Shen, A. J. Soroka, and R. D. Snook, A model for cw laser induced mode-mismatched dual-beam thermal lens
spectrometry based on beam prole image detection, J. Appl. Phys. 78, 700-708 (1995).
16
Stephen E. Bialkowski and Agn` es Chartier, Diffraction effects in single- and two-laser photothermal lens spectroscopy,
Appl. Opt. 36, 6711-6721 (1997).
17
M. A. C. de Ara ujo, R. Silva, E. de Lima, D. P. Pereira, and P. C. de Oliveira, Measurement of Gaussian laser beam radius
using the knife-edge technique: improvement on data analysis, Appl. Opt. 48, 393-396 (2009).
022154-12 Silva et al. AIP Advances 1, 022154 (2011)
18
J. M. Khosroan and B. A. Garetz, Measurement of a Gaussian laser beam diameter through the direct inversion of
knife-edge data, Appl. Opt. 22, 3406-3410 (1983).
19
A. Marcano, L. Rodriguez, and N. Melikechi, Thermal lensing in extended samples, Appl. Spectrosc. 56, 1504-1507
(2002).
20
D. Dadarlat, H. Visser and D. Bicanic, An improved inverse photopyroelectric cell for measurement of thermal effusivity:
application to fatty acids and triglycerides, Meas. Sci. Technol. 6, 1215-1219 (1995).

Вам также может понравиться