Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Spectrochimica Acta Part A 60 (2004) 1817–1823

Cw-laser thermal lens spectrometry in binary mixtures of water


and organic solvents: composition dependence of the
steady-state and time-resolved signals
Nadine Arnaud, Joseph Georges∗
Laboratoire des Sciences Analytiques, Bât. CPE-308 D, Université Claude Bernard—Lyon 1, 69622 Villeurbanne Cedex, France

Received 2 April 2003; received in revised form 1 September 2003; accepted 1 September 2003

Abstract

The thermal lens effect obtained in binary liquid systems composed of water and ethanol, propanol and acetonitrile has been investigated.
The dependence of dn/dT upon the solvent volume fraction follows polynomials up to sixth order and cannot be precisely predicted using the
additive rule. The sensitivity of the thermal lens method upon the addition of organic solvent in water varies as the temperature-dependent
refractive index gradient to thermal conductivity ratio of the mixture provided that the signal is sampled correctly. Otherwise, especially when
steady-state experiments are carried out, the thermally induced concentration gradient, known as the Soret effect, can change the thermo-optical
properties of the solution locally in the irradiated area and produce an additional signal. This effect depends on the solvent and is maximum
at low solvent composition. At the critical solvent volume fraction of 0.1–0.15, the Soret component may represent up to 25% of the pure
thermal lens signal and has a time constant which is 200–400 times greater than the characteristic time constant of the thermal lens.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Thermal lens spectrometry; Binary liquid systems; Refractive index; Temperature gradient; Concentration gradient

1. Introduction erties due to higher dn/dT and lower k values. Therefore, it


may be advantageous to use water–solvent mixtures to im-
Thermal lens spectrometry (TLS) results from the tem- prove the sensitivity of the TLS method [5,6]. The effect of
perature rise which is subsequent to absorption of optical ra- added methanol, acetone and acetonitrile on the sensitivity
diation and nonradiative relaxation of the excited molecules. and detection limits of chromium species using ion chro-
Due to the Gaussian profile of the excitation laser, the matography and TLS detection has been investigated [7].
temperature gradient produces a refractive index gradient For liquid mixtures, the composition dependence of the
which maximizes at the beam center and behaves like a photothermal signal is difficult to predict and one needs in-
converging or diverging lens depending on whether dn/dT is formation on the thermo-optical properties of the mixtures.
positive or negative, respectively [1–4]. The signal is mea- Most often, the relationship between sensitivity and solvent
sured as a change in the beam center intensity of a probe composition is estimated assuming that dn/dT and k are both
beam passing through the thermal lens. Besides the laser linear functions of the volume fraction of each solvent [6,7].
power and the solute concentration, the signal is sensitive However, the composition dependence of the refractive in-
to the thermo-optical properties of the medium, namely the dex gradient is generally a more complicated function [8]
temperature-dependent refractive index and the thermal con- and reliable values must come from experiments. Such de-
ductivity or heat capacity. Owing to its low dn/dT and high k termination of dn/dT, thermal conductivities or heat capaci-
values, water is a poor solvent for thermal lens spectrometry. ties have been made using the thermal lens method [9–14],
On the contrary, alcohols have better thermo-optical prop- but precise determination of dn/dT requires knowledge of
the thermal conductivity, the sample absorbance and a ref-
∗ Corresponding author. Tel.: +33-4-724-32627; erence medium to calibrate the experimental set-up if the
fax: +33-472-431-078. optical and geometrical parameters of this set-up are not
E-mail address: j.georges@cpe.fr (J. Georges). known [15]. In addition, it is necessary that the refractive

1386-1425/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.saa.2003.09.019
1818 N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823

index of the solution is not influenced by other factors than Table 1


the temperature. Especially, in solutions with more than one Thermo-optical parameters and physical constants for water, ethanol,
propanol and acetonitrile at 20 ◦ Ca
constituent, the temperature gradient induces thermal dif-
fusion and migration of molecules in the thermal gradient. Water EtOH PrOH ACN
This tendency of species to diffuse under the influence of a k (W m−1 K−1 ) 0.595 0.167 0.155 0.179
temperature gradient, known as the Soret effect, may result (dn/dT) × 104 (K−1 ) −0.9 −3.87 −3.8 −4.55
in the formation of a concentration gradient leading to a lo- Cp (J g−1 K−1 ) 4.18 2.41 2.31 2.14
cal change in the thermo-optical properties of the solution. ρ (g cm−3 ) 1 0.79 0.8 0.78
tc (ms) 2.95 4.8 5.0 3.9
The Soret effect has been observed in binary liquid mixtures
a Average of literature values.
near a consolute critical point [16–18], in micellar systems
[14,19,20] and in polymer solutions [21]. Depending on the
time scale of data recording, the concentration gradient can refractive index of water, ethanol, propanol, acetonitrile and
affect the thermal lens signal, which may lead to erroneous their mixtures was measured using a thermostated Abbé re-
interpretation of experimental data. fractometer over 1 ◦ C intervals between 14 and 26 ◦ C. The
The aim of this work was to investigate the composition dependence of the refractive index on temperature, dn/dT,
dependence of the thermal lens signal in binary systems and composition at room temperature, dn/dF, was deduced
composed of water and ethanol (EtOH), n-propanol (PrOH) by fitting the data to polynomial functions and then tak-
or acetonitrile (ACN). Time-resolved and steady-state ex- ing the derivative with respect to temperature and compo-
periments were carried out in order to separate the effects sition, respectively. Thermal conductivities of pure liquids
originating from a change in the temperature-dependent re- were taken from the literature (Table 1). For the binary sys-
fractive index due to a change in the macroscopic properties tems, the thermal conductivities have been estimated using
of the solution from the effects produced by the concentra- the Filippov equation which was shown to be better than the
tion gradient created locally in the irradiated region of the additive mass fraction model [22]. Then, the sensitivity of
solution. thermal lens spectrometry in the binary systems relative to
that in pure water was expected to vary as the dn/dT to k
ratio.
2. Experimental
2.3. Signal processing
2.1. Instrumentation
Under cw-laser excitation, the strength of the thermal lens
The dual-beam thermal lens spectrometer was the same as is defined by the inverse focal length:
that described and calibrated in a previous work [15]. It con- 1 P(1 − 10−A ) dn
sisted of an air-cooled argon-ion laser (λ = 488 nm) as the = (1)
f πkωoe
2 dT
excitation laser and a helium-neon laser as the probe beam.
Both beams were focused independently by two lenses and where P and ωoe are the power and the radius of the excita-
the set-up was arranged in a mode-mismatched configura- tion beam into the sample cell, A is the absorbance and k is
tion. The excitation beam was allowed to irradiate the sam- the thermal conductivity of the solution.
ple or blocked using a mechanical shutter. The probe beam The relation for the probe beam intensity change, derived
intensity change was measured through a 1 mm pinhole with from the theoretical model developed by Snook et al. [15,23],
a PIN photodiode and processed with a digital oscilloscope may be written as:
(Tektronix TDS 2002) connected to a PC computer. For each
sample, the signal was derived from the average of several I(t) = I(o)(1 − b(t)θ)2 (2)
recordings. b(t) is the time-dependent term which describes the forma-
tion of the thermal lens signal:
2.2. Reagents and procedures
 
1 −1 2mV
b(t) = tan
Thermal lens experiments were carried out with cobalt 2 [(1 + 2m)2 + V 2 ](tc /2t) + 1 + 2m + V 2
nitrate using a 2 mm quartz cuvette. Solutions of appropri- (3)
ate absorbances were prepared by successive dilutions from
stock solutions and absorbances were measured with a con- where m = (ωpc /ωoe )2 and V = Z1 /Zc are geometrical pa-
ventional spectrophotometer. The water–solvent mixtures rameters of the mode-mismatched dual-beam configuration,
were prepared by weighing and the composition was finally and tc = ωoe 2 ρC /4k is the characteristic time constant of
p
expressed as volume fraction of the organic solvent, F. De- the thermal lens effect; ωpc the radius of the probe beam in
pending on the mixture composition, the sample absorbance the cell, Z1 the distance from the waist of the probe beam
was adjusted between 0.004 and 0.06 in order that the ther- to the cell, Zc the confocal distance of the probe beam, and
mal lens signal was remaining approximately constant. The ρ and Cp are the density and specific heat of the medium,
N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823 1819

Table 2 where kT = DT /Dm is the thermal diffusion ratio, i.e. the


Experimental parameters used for the theoretical fits coefficient of thermal diffusion compared to that of mass dif-
Excitation laser spot size at cell, ωec 41 ␮m fusion. Following Eqs. (1)–(2) and assuming that dn/dF and
Probe laser spot size at cell, ωpc 100 ␮m dn/dT have the same effect on the propagation of the probe
Excitation laser power at cell 4 mW beam, the probe beam intensity change can be written as:
m 6.3
Zc 7.5 mm I(t) = I(o)(1 − b(t)θ − b (t)θ  )2 (6)
Z1 14 mm
where θ  is related to the amplitude of the Soret signal and
b (t) is the time-dependent term which describes the forma-
respectively. θ is a dimensionless parameter which indicates tion of the thermally induced concentration gradient. This
the strength of the thermal lens: term has been expressed as [25,26]:
(1 − 10−A )P dn  ∞  
θ=− (4)  4 (2n − 1)π
λp k 
dT b (t) = 1 − sin
(2n − 1)π 2
n=1
where λp is the wavelength of the probe beam. θ and b(t) can  
be calculated using the sample absorbance, the laser power in t
× exp −(2n − 1)2 (7)
the cell, the thermo-optical properties of the solution and the tD
geometrical parameters of the experimental set-up (Table 2)
which has been calibrated in a previous work [15]. where tD is the mass diffusion time.
In mixtures, the temperature gradient produced by re-
laxation of the excited species induces mutual migration
of both components which results in the formation of a 3. Results and discussion
composition-dependent refractive index gradient, dn/dF.
This gives rise to a concentration lens which adds to the 3.1. Changes in dn/dT and dn/dF
thermal lens. The inverse focal length of the new lens has
been expressed as [17,24]: The experimental values obtained for dn/dT are reported
  in Fig. 1. The data show that dn/dT as a function of the
1 P(1 − 10−A ) dn kT dn mixture composition does not follow a simple additive
= − (5)
f πkωoe2 dT T dF rule, but is best represented by fourth, fifth- and sixth-order

Fig. 1. Plots of the temperature-dependent refractive index gradient dn/dT × 104 (K−1 ) as a function of the solvent volume fraction for the water–solvent
binary systems. The dashed lines are polynomial fits to experimental data.
1820 N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823

Fig. 2. Plots of the composition-dependent refractive index gradient dn/dF at room temperature.

polynomials for ethanol, acetonitrile and propanol, respec- fractive index gradient, dn/dF. As shown in Fig. 2, the three
tively. The refractive index as a function of the mixture mixtures exhibit a different behavior: for water–propanol,
composition at room temperature followed second-, third- dn/dF decreases linearly as the solvent volume fraction
and fourth-order polynomials for propanol, ethanol and increases while for water–ethanol and water–acetonitrile
acetonitrile, respectively. The derivative of these functions systems, the variation is more complicated and exhibits a
allowed the calculation of the concentration-dependent re- sign change at a solvent volume fraction close to 0.7–0.8.

Fig. 3. Change in probe beam intensity upon step excitation for cobalt nitrate in water–propanol mixture (F = 0.1). The thin line is the fit of Eq. (2) to the
experimental curve within the 0–400 ms time range after the beginning of excitation. The thicker line is the fit of Eq. (6) to the overall experimental curve.
N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823 1821

3.2. Time-resolved signal processes can be easily separated because the Soret compo-
nent builds up with a time constant which is much greater
A typical curve for the probe beam intensity change un- than the characteristic time constant of the thermal lens.
der step excitation is shown in Fig. 3. The initial fast change
corresponds to the formation of the temperature-dependent 3.2.1. Relative sensitivity of the TLS method
refractive index gradient (thermal lens signal) while the sec- In pure solvents, where the Soret effect is not efficient,
ond part corresponds to the slower build-up of the ther- the fit of Eq. (2) using the calculated values of b(t) and θ
mally induced concentration gradient (Soret signal). Both was valid until steady-state is reached. On the contrary, in

Fig. 4. Plots of the relative sensitivity of the thermal lens method as a function of the solvent volume fraction for the (a) water–ethanol, (b) water–propanol
and (c) water–acetonitrile binary systems. The dotted lines represent the relative sensitivities expected from the dn/dT to k ratios.
1822 N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823

Fig. 4. (Continued ).

most binary mixtures, the fit was limited to the short period, of the transient steady-state intensity from which the ther-
typically 400 ms, within which mass diffusion is inoperative mal lens signal was obtained. The data, expressed as the
and the probe beam intensity change is only determined by response obtained in the binary systems relatively to that in
the temperature-dependent refractive index gradient. The fit pure water under the same experimental conditions, show
to the initial fast intensity change allowed the determination that the sensitivity adequately follows the dn/dT to k ratio

Fig. 5. Plots of the Soret signal, expressed as percent of the pure thermal lens signal, as a function of the solvent volume fraction.
N. Arnaud, J. Georges / Spectrochimica Acta Part A 60 (2004) 1817–1823 1823

of the binary system (Fig. 4). On the contrary, if the sig- where Dm is the coefficient of mass diffusion. Diffusion
nal was taken at the final steady-state, i.e. within a few sec- coefficients of 3.5 × 10−6 cm2 s−1 have been derived from
onds, the true thermal lens signal would be affected by the experimental tD values of 1.2 s for both water–ethanol and
concentration gradient. It is clear from Fig. 3 that the ef- water–propanol systems at solvent volume fractions of 0.15
fect of the concentration gradient depends on the excitation and 0.1, respectively. Such a data treatment has previously
duration and the time at which the signal is sampled af- been validated for micelles and polymers in aqueous so-
ter the beginning of excitation. It is expected that working lutions [21]. However, the values obtained in the present
with short irradiation time and fast data collection would work are about four times smaller than the mutual diffusion
prevent the effect of mass diffusion. However, it has been coefficients measured for the same mixtures at the same
shown in a previous work [27] that even upon chopped exci- composition using the Taylor dispersion technique [28,29].
tation and lock-in detection, a concentration gradient builds Further work is required to understand the reason of this
up slowly to a steady-state and changes the probe beam disagreement.
intensity.

3.2.2. Amplitude and kinetics of the Soret component References


The Soret signal was taken as the difference between the
transient probe beam intensity, as deduced from Eq. (2), [1] N.J. Dovichi, CRC Crit. Rev. Anal. Chem. 17 (1987) 357.
and the final steady-state intensity. As shown in Fig. 5, the [2] J. Georges, Talanta 41 (1994) 2015.
[3] R.D. Snook, R.D. Lowe, Analyst 120 (1995) 2051.
effect of mass diffusion is more important with ethanol
[4] S. Bialkowski, Photothermal spectroscopy methods for chemical anal-
and propanol than with acetonitrile. With both alcohols, the ysis, in: J.D. Winefordner (Ed.), Chemical Analysis, vol. 134, Wiley,
error produced by the concentration gradient may represent New York, 1996.
up to 25% of the pure thermal lens signal when the solvent [5] N.J. Dovichi, J.M. Harris, Anal. Chem. 51 (1979) 728.
volume fraction is within the range 0.05–0.25. This result [6] W.A. Weimer, N.J. Dovochi, J. Appl. Phys. 59 (1986) 225.
[7] M. Sikovec, M. Franko, M. Novic, M. Veber, J. Chromatogr. A 920
may be explained by two factors: (i) dn/dF is much smaller
(2001) 119.
with acetonitrile than with ethanol and propanol; (ii) as the [8] W.B. Li, P.N. Segre, R.W. Gammon, J.V. Sengers, J. Chem. Phys.
solvent composition increases, dn/dT increases while dn/dF 101 (1994) 5058.
decreases, which contributes to minimize the effect of mass [9] M.C. Gupta, S.-D. Hong, A. Gupta, J. Moacanin, Appl. Phys. Lett.
diffusion with respect to the thermal lens signal. At inter- 37 (1980) 505.
[10] R.D. Snook, R.D. Lowe, M.L. Baesso, Analyst 123 (1998) 587.
mediate solvent compositions, the concentration gradient
[11] K. Seidman, A. Payne, J. Chem. Educ. 75 (1998) 897.
had a reverse effect and the final steady-state intensity was [12] S.M. Colcombe, R.D. Snook, Anal. Chim. Acta 390 (1999) 155.
significantly greater than the transient steady-state intensity, [13] C.V. Bindhu, S.S. Harilal, V.P.N. Nampoori, C.P.G. Vallabhan, Opt.
which is reported by negative values in Fig. 5. The negative Eng. 37 (1998) 2791.
contribution of the Soret effect is unexpected within this [14] M. Franko, C.D. Tran, J. Phys. Chem. 95 (1991) 6688.
[15] M. Fischer, J. Georges, Anal. Chim. Acta 322 (1996) 117.
concentration range given that the measured dn/dF values
[16] F.D. Hardcastle, J.M. Harris, Appl. Spectrosc. 40 (1986) 606.
are positive. It is likely that the concentration gradient in- [17] M. Giglio, A. Vendramini, Appl. Phys. Lett. 25 (1974) 555.
duces local changes of the solution properties including [18] M. Giglio, A. Vendramini, Phys. Rev. Lett. 34 (1975) 561.
absorbance, the thermal conductivity and the refractive in- [19] J. Georges, T. Paris, Anal. Chim. Acta 386 (1999) 287.
dex. Although it is difficult to account for each of them [20] N. Arnaud, J. Georges, Spectrochim. Acta Part A 57 (2001) 1085.
[21] J. Georges, Spectrochim. Acta Part A 59 (2003) 519.
individually, these changes could result in a slight decrease
[22] G. Cai, H. Zong, Q. Yu, R. Lin, J. Chem. Eng. Data 38 (1993) 332.
of the thermal lens signal. [23] J. Shen, R.D. Lowe, R.D. Snook, Chem. Phys. 165 (1992) 385.
As shown in Fig. 3, the slow component is perfectly fitted [24] F.D. Hardcastle, J.M. Harris, Appl. Spectrosc. 40 (1986) 606.
by Eq. (6), retaining terms up to n = 4 in the summation [25] K.J. Zhang, M.E. Briggs, R.W. Gammon, J.V. Sengers, J. Chem.
of Eq. (7). The fit allowed the determination of the mass Phys. 104 (1996) 6881.
[26] K.J. Zhang, M.E. Briggs, R.W. Gammon, J.V. Sengers, J.F. Douglas,
diffusion time tD , which, by analogy to the thermal time
J. Chem. Phys. 111 (1999) 2270.
constant tc , has been defined as [21]: [27] N. Arnaud, J. Georges, Anal. Chim. Acta 445 (2001) 239.
[28] K.R. Harris, T. Goscinska, H.N. Lam, J. Chem. Soc., Faraday Trans.
ω2
tD = ec (8) 89 (1993) 1969.
4Dm [29] L. Hao, D.G. Leaist, J. Chem. Eng. Data 41 (1996) 210.

You might also like