Vous êtes sur la page 1sur 8

Nalda et al.

Vol. 19, No. 2 / February 2002 / J. Opt. Soc. Am. B

289

Limits to the determination of the nonlinear


refractive index by the Z-scan method
R. de Nalda, R. del Coso, J. Requejo-Isidro, J. Olivares, A. Suarez-Garcia, J. Solis, and C. N. Afonso
Instituto de Optica, Consejo Superior de Investigaciones Cientificas, Serrano 121, 28006 Madrid, Spain
Received May 14, 2001; revised manuscript received September 13, 2001
We analyze the limitations imposed by sample absorption on the determination of the nonlinear refractive index by the Z-scan technique. By using a nanostructured thin film consisting of Cu nanocrystals embedded in
a dielectric Al2 O3 matrix as an example, we show that thermo-optical effects appearing when linear absorption
is significant can be strongly misleading in the interpretation of the results of a Z scan. Even though this
effect is not new, the widespread use of the Z-scan technique during the past several years makes it necessary
to analyze explicitly the conditions under which the technique can be reliably applied and when more sophisticated techniques should be used instead. We discuss the contributions to the signal under different experimental conditions, several diagnostic techniques to discriminate true nonlinear effects from thermally induced
phenomena, and different methods to reduce the thermal contribution. 2002 Optical Society of America
OCIS codes: 000.2170, 160.4330, 190.3270, 190.4380, 190.4400, 190.4720, 190.4870.

1. INTRODUCTION
The characterization of the nonlinear optical properties of
materials is of utmost interest in several fields of physics,
both from the fundamental and the applied points of view.
In particular, great effort has been devoted to the determination of the third-order nonlinear optical susceptibility, ( 3 ) , responsible for phenomena such as thirdharmonic generation or optical phase conjugation. A
wide range of techniques have been used to measure ( 3 ) :
Z scan,1 degenerate four-wave mixing2,3 (DFWM), nearly
degenerate three- and four-wave mixing,4 nonlinear
interferometry,5 ellipse rotation,6 optical third-harmonic
generation,7 beam-distortion measurements,8 and photoacoustics experiments.9
The Z-scan method provides a sensitive and straightforward method for the determination of the sign and the
values of the real and the imaginary parts of ( 3 ) , respectively, proportional to the nonlinear refractive index and
the nonlinear absorption coefficient. The simplicity of
both the experimental setup and the data analysis has allowed the Z-scan method to become widely used by many
research groups. Extensions of the basic technique to allow the use of non-Gaussian-beam profiles,10,11 thick
samples,1214 or saturable Kerr media15 have been proposed, and certain improvements in sensitivity have been
achieved by measuring the total beam-profile
distortion12,16,17 or through the so-called eclipsing Z-scan
method.18,19
The other widely used technique, degenerate four-wave
mixing (DFWM), involves a far more complex experimental apparatus but provides several advantages. The fact
that the setup includes temporal and spatial overlapping
of three separate beams permits increased flexibility, such
as the possibility of measuring different tensor components of ( 3 ) , a straightforward study of temporal behavior, and less strict conditions on the beam profile and the
surface roughness.
When the Z scan was originally devised, it was used to
characterize the nonlinear susceptibility of transparent
0740-3224/2002/020289-08$15.00

bulk materials. Nevertheless, its use has now been extended to the study of a wide variety of samples. In particular, it is often used to study absorbing media. Among
the latter, its use for the assessment of materials consisting of semiconductor or metal crystallites of nanometer
size embedded in dielectric matrices has been very common. These composites are currently the object of intensive research, their most direct application being related
to their high third-order nonlinear susceptibility,
( 3 ) (; , , ), which makes them promising candidates for the development of all-optical switching
devices.20,21 In the case of metal nanocomposites, several
systems have been investigated, including Au,2224 Ag,20
Cu,25,26 Sn,27 or Ni26 crystallites embedded in dielectric
glasses, mostly SiO2 and Al2 O3 . Yet, the feature that
makes these materials exhibit large ( 3 ) values, that is,
the surface-plasmon resonance, also makes them efficient
absorbers of radiation in the wavelength region of interest.
The aim of this paper is to show the limitations of the
Z-scan method for the determination of the nonlinear refractive index of absorbing media. The limitations are
essentially caused by the presence of thermo-optical effects that can mask the true nonlinear response of the
material. Even though this problem has been documented before, the widespread use of the Z scan often
leads to results obtained from naively interpreted Z-scan
experiments that provide values for the nonlinear refractive index and the third-order nonlinear susceptibility,
which are simply not trustworthy. We use a metal nanocomposite film (Cu:Al2 O3 ) and we refer to the basic configuration of the Z scan28 in order to illustrate the difficulties that can arise. The results presented here
provide indications on how to perform the diagnosis of the
origin (thermo-optical, nonlinear, or a combination of
both) of the results obtained in a Z-scan experiment and a
survey of procedures that can either help to discriminate
between these two effects or to enhance the contribution
of one with respect to the other.
2002 Optical Society of America

290

J. Opt. Soc. Am. B / Vol. 19, No. 2 / February 2002

Nalda et al.

2. EXPERIMENT
The samples used throughout the experiments described
here are composite thin-film materials formed by Cu
nanocrystals embedded in a dielectric amorphous Al2 O3
matrix. The samples were synthesized with the alternate pulsed-laser deposition technique.29 The films consist of ten layers of Cu nanocrystals embedded in amorphous Al2 O3 with a total sample thickness of 110 nm.
The Cu content of the films was determined by Rutherford backscattering spectrometry (RBS) with a 2.0-MeV
4
He beam. The experimental spectra were simulated
by use of the RUMP code.30 Their structure was characterized by high-resolution electron microscopy31 and the
small-angle x-ray scattering technique under grazing
incidence (GISAXS).32,33 One sample was grown on
a 1-mm-thick glass substrate with a Cu content of
12 1016 at/cm2 and presenting oblate nanocrystals with
an in-plane average diameter of 6 1 nm. Then, a series of similar samples was deposited on three different
substrate materials (1-mm-thick glass, 1-mm-thick fused
silica, and 2-mm-thick sapphire) with relative thermal
conductivities of 1:1.05:30, respectively. The Cu content of these later samples was slightly smaller
(9.5 1016 at/cm2 ) than that of the first one, with a correspondingly smaller average diameter for the Cu nanocrystals (5 1 nm). Further details on the growth procedure and the characterization of the films can be found
elsewhere.25,32
The nonlinear response of the material was characterized by means of the DFWM and Z-scan techniques. The
laser beam used in both cases was the output of a cavitydumped, synchronously pumped mode-locked dye laser
tuned at 585 nm, which produced 12-ps pulses at a variable repetition rate from 80 kHz to 40 MHz.
Figure 1 shows the experimental setup for the DFWM
measurements, which were performed in the forward
folded-box scheme.2 The laser beam was split in three
branches, allowing separate control of delay and polarization. The beams were then focused with an achromatic
80-mm focal-length lens and overlapped in the focal region of the lens with a waist of 30-m diameter. The
maximum peak power34 attainable was 1 kW, which, at
focus, leads to a typical peak intensity of 100 MW/cm2.
In the overlap region, a fourth wave generates through
the nonlinear interaction. This conjugated beam was de-

Fig. 1. Experimental setup used for the forward folded-box degenerate four-wave mixing experiments.
NL, nonlinear
(sample).

Fig. 2. Experimental setup used for the Z-scan experiment:


Ref Pd, reference photodetector; FF PD, far-field photodetector;
NF PD, near-field photodetector. The plots show the typical
shape of the far-field Z scan for n 2 0 (upper graph) and the
typical near-field Z scan for 0 (lower graph).

tected and recorded with standard phase lock-in techniques. Choosing the appropriate combination of polar(3)
can be
izations, each of the tensor components of ijkl
measured. For the aim of this paper, we refer only to the
configuration in which all the beams have parallel polar(3)
ization, since that is the one that provides iiii
, the com(3)
ponent to be directly compared with the values determined through Z-scan measurements.
The Z-scan setup is shown in Fig. 2. It basically follows the configuration first reported in Ref. 28. The
sample, located on a translation stage, is moved along the
propagation path of the pulsed linearly polarized laser
beam, which is focused by means of a lens, and the optical
transmission is monitored by two different photodiodes.
In our particular case, the laser beam was focused with a
150-mm focal-length lens to a beam waist of 30 m and
with a Rayleigh range of 1 mm. The maximum peak
power density attainable at the focus of the lens under
these conditions was similar to the one used for the degenerate four-wave mixing experiment, except at 40 MHz,
for which it was one order of magnitude smaller. An additional set of experiments was performed with a shorter
focal-length lens ( f 75 mm) in order to attain peak
power densities around four times higher. As a consequence of the different power densities experienced by the
sample at different distances from the focus of the lens,
the total transmission of the sample and the beam divergence after the lens are modified, giving rise to changes in
the transmission of the sample as it moves. These
changes are monitored by the so-called near-field and far-

Nalda et al.

field detectors, respectively, and a third detector is used to


monitor the possible fluctuations of the beam intensity
before reaching the sample. The far-field detector, partially masked with a circular aperture of 0.1-cm diameter,
was located 45 cm away from the focus of the lens, thus
allowing only the central part of the beam to be measured. The near-field detector, not masked, measured
the transmission of the sample by collecting a fraction of
the total transmitted beam after going through the
sample. This detector was therefore sensitive only to absorptive nonlinearities. For the detectors, the signal evolution during the scanning of the sample was recorded
with standard lock-in amplification techniques. Typical
results for the far-field and near-field scans are also
shown in Fig. 2.

3. RESULTS
A. DFWM Measurements
Figure 3 shows the intensity of the conjugated beam as a
function of the total intensity of the pump pulses34 measured in the Cu:Al2 O3 samples on glass with Cu contents
of 12 1016 and 9.5 1016 at/cm2 . The dotted curve is
a fit of the experimental data to a cubic dependence, as
expected for a Kerr-type third-order nonlinearity. The
good agreement between the fit and the experimental
data allows one to discard the presence of saturation effects in the power-densities interval analyzed. It was
checked that the conjugated signal from a bare substrate
(glass, fused silica, sapphire) was in all cases below the
detection threshold limit. The linear absorption spectrum of the sample with the highest Cu content is shown
in the inset of Fig. 3. It shows an absorption band at
620 nm due to the surface-plasmon resonance of the Cu
nanocrystals in Al2 O3 . 25
Under the assumption of low conversion efficiency, ( 3 )
can be readily obtained from the cubic fit shown in Fig. 3,

Fig. 3. Intensity of the conjugated beam as a function of pump


intensity in the DFWM experiment for Cu:Al2 O3 samples on
glass with a Cu content of (stars) 12 1016 at/cm2 , (filled
squares) 9.5 1016 at/cm2 , and for (hollow circles) a 5-mm-thick
CS2 cell. The dotted curves correspond to cubic fits of the data.
The absorption spectrum of the sample with a Cu content of
12 1016 at/cm2 is shown in the upper-left-corner inset. The
surface-plasmon resonance can be observed at 620 nm. The
arrow indicates the excitation wavelength used in the Z-scan and
degenerate four-wave mixing experiments.

Vol. 19, No. 2 / February 2002 / J. Opt. Soc. Am. B

291

provided the temporal and spatial characteristics of the


pulses are known. The determination of ( 3 ) is, however,
subject to smaller errors when calculated by comparison
with the results obtained from a reference material. If
the reference sample is nonabsorbing, one can relate the
third-order nonlinear susceptibility of the sample with
that of the reference through the expression2

S3


nS

n ref

L ref

bS

LS

b ref

1/2

L exp L/2

1 exp L

3
ref
,

(1)

where the S subscript refers to the sample and the ref


subscript to the reference. In this formula n is the linear
refractive index, L is the interaction length, and b is a parameter related to the intensity of the conjugated wave
3
through I C b I pump
, I C and I pump being the intensities of the conjugated and pump beams, respectively. Finally, is the absorption coefficient of the sample. The
sensitivity of our experimental DFWM setup is given by
the minimum signal that can be discriminated from background scatter and that corresponds to a conjugation efficiency of 5 108 . For thick samples (thickness
bigger than the interaction length of the three incoming
(3)
beams, 0.3 mm), our limit in the determination of iiii
is
14
510
esu.
As is common practice, we have chosen CS2 as the ref(3)
2 1012 esu). 2 Figure 3 also inerence sample ( iiii
cludes the evolution of the conjugated signal from a
5-mm-thick CS2 cell measured under the same conditions
used for the Cu:Al2 O3 samples from which the b parameter for each sample is determined. Following this
analysis, we obtain values for the nonlinear susceptibility
(3)
of the Cu:Al2 O3 samples of iiii
(8 4) 108 esu
(3)
8
and iiii (3 1) 10 esu, respectively, for the
sample with the highest Cu content and for the series of
samples on different substrates having lower Cu content.
These values are among the highest ones ever reported, to
our knowledge, for this type of specimen.25,26,35 Additionally, we have checked that the buildup time of the conjugated signal was in all the cases shorter than the laser
pulse duration, whereas the decay time was of the order
of a few picoseconds,24 as we observed by delaying the arrival of one of the beams. Therefore the conjugated signal can be attributed to a true electronic nonlinear effect
in the nanocomposite samples.
B. Z-Scan Measurements
The nonlinear refractive index and absorption coefficient
of the Cu:Al2 O3 films on glass substrates were measured
by the Z-scan technique, performing several series of experiments in which different pulse peak powers and repetition rates were used. During the course of these experiments, it was clear that nonlinear absorption effects
in the samples were negligible in all cases for the power
densities used, as the transmission changes in the nearfield scans were 2%. From this limit, it is possible to
estimate that the value for the nonlinear absorption coefficient is smaller than 5 105 cm/W, which leads to an
imaginary part of the nonlinear susceptibility Im (3)
108 esu. These values agree well with earlier results reported in Cu nanocomposites29,36 in the vicinity of
the surface-plasmon resonance.

292

J. Opt. Soc. Am. B / Vol. 19, No. 2 / February 2002

Nalda et al.

The far-field scans corresponding to the sample with


the higher Cu content are shown in Fig. 4. They were obtained for power densities of 0.051 kW and repetition
rates in the 0.44-MHz range. The results for the lowest
repetition rate used, 80 kHz, with a total transmission
change in the far field close to the detection limit are not
shown in this graph for the sake of clarity.
The analysis of the results of the Z-scan experiments
was performed within the frame of the model developed
by Sheik-Bahae et al.1 The method consists of a decomposition of the complex electric field at the exit plane of
the sample into a summation of Gaussian beams.37 For
third-order nonlinearities, the real part of the refractive
index can be written as
n n 0 n 2 I,

(2)

where I is the intensity (in watts per centimeter squared)


of the laser pulse.
As a consequence of the Gaussian-beam profile of the
laser beam used, the sample behaves as a nonlinear lens
that modifies the intensity distribution in the far field.
Provided a number of assumptionsGaussian beam, thin
sample (l z 0 , where l is the thickness of the sample
and z 0 is the Rayleigh range), instantaneous nonlinearity,
negligible nonlinear absorption effects (Im (3) Re (3)),
small phase change, and far-field condition (d z 0 ,
where d is the position of the far-field detector)are satisfied for a z position, the relative on-axis transmittance
of the sample measured (at the small aperture of the farfield detector) is given by
T z, 0 1

4 0 z/z 0
z 2 /z 02 9 z 2 /z 02 1

(3)

where 0 is the phase change, given by


0

Fig. 5. Peak-to-valley transmittance change (T p-v) as a function of the mean power of the laser beam obtained from the Z
scans in the Cu:Al2 O3 sample on glass with a Cu content of
12 1016 at/cm2 . The mean power was varied either by changing the repetition rate while keeping the peak power constant
(filled symbols) or by changing both the repetition rate and peak
power (hollow symbols): (solid squares) 80 kHz, 1 kW; (solid triangles) 400 kHz, 1 kW; (solid circles) 800 kHz, 1 kW; (hollow diamonds) 4 MHz, 0.26 kW; (hollow inverted triangles) 4 MHz, 0.44
kW; (hollow squares) 40 MHz, 0.042 kW; and (hollow circles) 40
MHz, 0.055 kW.

where I 0 is the peak intensity, is the wavelength of the


radiation, and L eff is the effective length, given by L eff
1 exp(L)/, where is the absorption coefficient.
The peak-to-valley transmittance change (vertical distance between maximum and minimum in each of the
scans of Fig. 4) and the peak-to-valley separation in z
(horizontal distance between maximum and minimum)
are then given by
T p-v 0.406 0 ,

L effn 2 I 0 ,

(4)

(5)

Expressions (4) and (5) provide a straightforward method


to determine the nonlinear index of refraction from the
measurement of the on-axis far-field transmittance
change, once the characteristic parameters of the sample
and the laser pulses are known.
The real part of the third-order nonlinear optical susceptibility, ( 3 ) , can then be readily calculated as2
Re 3 esu

Fig. 4. Experimental far-field Z-scan results in the Cu:Al2 O3


sample on glass with a Cu content of 12 1016 at/cm2 . The corresponding repetition rates and peak powers are the following:
(filled triangles) 400 kHz, 1 kW; (filled circles) 800 kHz, 1 kW;
(hollow inverted triangles) 4 MHz, 0.44 kW; (hollow circles) 40
MHz, 0.055 kW. The curves are guides to the eye. The dashed
curves corresponds to the Z-scan profile calculated by assuming
the (3) value measured through DFWM (3) (8 4)
108 esu .

z p-v 1.7z 0 .

n 02
0.0395

n 2 cm2 /W .

(6)

Applying this analysis to the results presented in Fig. 4


and depending on the repetition rate and pulse peak
power used, values of ( 3 ) of 107 105 esu are obtained,
which are clearly inconsistent with the one obtained by
DFWM (8 4 108 esu). Furthermore, these values
follow a clear trend: the higher the repetition rate of the
laser, the larger the apparent value of ( 3 ) . This is
shown in Fig. 5, where the peak-to-valley transmittance
change (T p-v) is plotted as a function of the laser mean
power. It can be seen that it depends linearly on the
mean power regardless of the peak power. In particular
for the lower repetition rates (kilohertz range), a constant
peak power is used, and thus it is deduced that T p-v increases linearly as a function of the laser repetition rate.

Nalda et al.

This is in contrast with the prediction of Sheik-Bahaes


model [Eq. (5)], from which one can expect a constant
value of T p-v for an electronic nonlinear effect and a
fixed peak power. The deviation manifests itself even
clearer for the higher repetition rates because T p-v becomes larger, even though the peak power is significantly
lower.
In order to make a clearer comparison between the results derived from both techniques, we performed a simulation of the far-field scan expected from a sample with
the value of ( 3 ) determined by DFWM. The result is
plotted in Fig. 4 as a dashed curve. It can be seen that
the peak-to-valley transmittance change corresponding to
the simulated scan is below the resolution of the experiment (T p-v 2%). This latter result may seem surprising considering that a value of ( 3 ) of 8 108 esu is
considerably high for most nonlinear media and that actually much lower values for ( 3 ) have been reported with
the Z-scan method.1,38 It has to be considered, though,
that the key parameter that establishes the detection
threshold is the product I 0 L eff (3). In our case, the
factor defining this threshold is the small effective length
of the sample. Moreover, it is important to note that no
advantages would be obtained with thicker samples.
The reason is that in Eq. (2), L eff 1 exp(L)/, L eff
increases with the thickness of the sample until
1 exp(L) 1, and above this limit there is no further dependence on the sample thickness. This is intuitively equivalent to the idea that, in absorbing materials,
the major contribution to the nonlinear lens arises from
the region of the sample close to the surface where the intensity has not dropped enough to suppress the nonlinear
effect. The rest of the thickness results only in a stronger absorption.
To further analyze the role of thermal effects, the
Z-scan experiments were repeated in a series of similar
samples grown on different substrate materials (glass,
fused silica, and sapphire) in order to cover a wide range
of thermal conductivities. The Z-scan experiments were
performed under a higher-intensity regime, by use of a

Fig. 6. Far-field Z-scan results obtained in the Cu:Al2 O3


samples with lower Cu content (9.5 1016 at/cm2 ) grown on different substrates: (solid circles) sapphire, (hollow squares)
glass, and (hollow triangles) silica, with a peak power of 1 kW
and a repetition rate of 4 MHz. The peak power density was
P peak 4 108 W/cm2 .

Vol. 19, No. 2 / February 2002 / J. Opt. Soc. Am. B

293

shorter focal-length lens (75 mm). The corresponding results are shown in Fig. 6 for a repetition rate of 4 MHz
and a peak power of 1 kW. The samples grown on fused
silica and glass show a similar behavior, whereas that
grown on sapphire, with a thermal conductivity 30 times
larger than that of the other substrates, presents a drastic change. In this latter case, the total transmittance
change is below the detection threshold.

4. DISCUSSION
A. Thermal-Lens Effect
The dependences observed in Fig. 5, in which T p-v scales
linearly with the mean power of the laser, as well as in
Fig. 6, in which T p-v is below the detection limit for the
film on the substrate with the highest thermal conductivity, clearly suggest that thermal-induced effects play a
dominant role in the results of the Z-scan experiments.
In the presence of non-negligible linear absorption, the
sample temperature will increase leading to the formation of a thermal lens when the refractive index of the
sample is temperature dependent (dn/dT 0). This
thermal lens can contribute to the signal of the far-field
detector along with the true nonlinear response of the
sample.
According to Eq. (4), we see that, when the mechanism
responsible for the buildup of a lens in the material is a
pure electronic nonlinearity, the phase shift, and similarly the peak-to-valley transmittance change, T p-v ,
scales linearly with the peak power of the pulses. The
thermal-lens effect depends on the total amount of heat
transferred to the sample and should scale roughly linearly with the mean power.39,40 Furthermore, the behavior plotted in Fig. 5, where the intercept of the linear fit
with the y axis is compatible with zero, is the evidence for
a purely thermo-optical effect, with a contribution from a
nonlinear term below the resolution limit. The thermal
nature of the Z-scan far-field signal is also supported by
the results shown in Fig. 6 for composite films grown on
three substrates having different thermal conductivities.
The dependence of the observed signal on the thermal
properties of the substrate cannot be explained by an electronic nonlinearity, but it is directly related to a thermallens effect. This thermal lens contributes to modifying
the divergence of the beam in a Z-scan experiment and
thus produces a signal that is superimposed on the nonlinear one. There are two situations for which this thermal effect can be of importance. The first is when long
pulses (200 ps) are used, because each of them can at
the same time induce and see the thermal lens. When
shorter pulses are used, this situation does not hold, since
the electronphonon coupling to the matrix is estimated
to have a characteristic time of 50200 ps.41 Instead, a
cumulative effect can happen when high repetition rates
(10 kHz) are used.
More precisely, the cumulative thermal effect occurs
whenever the spacing between pulses is shorter than the
characteristic thermal-diffusion time, t c w 2 /4D, where
w is the laser spot size and D (cm2 s1) is the thermaldiffusion coefficient of the sample material. Since D
is expected to be in the range of 1 103 to 6
103 cm2 s1 , 39,42 and typical beam waists in these ex-

294

J. Opt. Soc. Am. B / Vol. 19, No. 2 / February 2002

periments are 1050 m, t c should be greater than 40 s;


thus the above condition is met whenever the laser repetition rate is higher than a few tens of kilohertz. In this
regime, the sample does not return to the equilibrium
temperature in the time interval between consecutive
pulses, and a stationary state is reached once the heat absorbed in the illuminated region of the sample equals the
heat diffused in all directions. A quasi-permanent spatial temperature distribution is then created, which manifests itself as a lens, for a refractive index dependent on
temperature. Along the Z-direction, this lens produces a
far-field scan where the prefocal and the postfocal transmittance changes have opposite signs.
In our case, the shape of the Z scans observed is consistent with the formation of a positive lens, that is, dn/dT
0. It is important to note that this effect is a purely
linear one, although its effect is so similar to what is expected for a third-order nonlinear susceptibility that it is
not uncommon to find texts in the literature where the
thermal-lens effect is simply ignored.43,44 In some other
cases, even though the thermal origin of the beam distortion is recognized, it is described as a thermal nonlinearity, and the authors give thermal ( 3 ) values that are
meaningless.4547 The thermal-lens effect will play a significant role in all materials having a significant absorption in the spectral region of the laser radiation. Additionally, it has also been reported that it may occur
through multiphoton absorption48 in transparent materials. In contrast, in the case of DFWM, even when the
sample gets heated under irradiation, this does not contribute to the intensity of the conjugated beam. In our
experiment, this has been verified by measuring the temporal response of the conjugated signal in the case of the
Cu:Al2 O3 film. In this way, it is possible to ensure that
the observed conjugated signal is not caused by a longliving refractive-index inhomogeneity formed in the
sample through heating.46
B. Diagnosis and Alternatives
From the preceding paragraphs it is clear that, in those
cases in which the repetition rate of the laser can be easily changed, one should plot T p-v as a function of the
mean power, as done in Fig. 5, in order to determine the
true origin (thermal or nonlinear) of the far-field signal.
If the same peak intensity is used, no change in the Z
scan should be observed if the mechanism is purely nonlinear, whereas the thermal cumulative effect will yield Z
scans with T p-v scaling linearly on the repetition rate, as
was seen in our example. In fact, for any given system
and sufficiently short pulses, one should find a repetition
rate that is low enough that thermal-lens effects are negligible, allowing a direct measurement of Re (3) by means
of Z scan,49 and provided that the value of T p-v for the
nonlinear behavior is above the experimental discrimination threshold.
Another possibility that was explored in this paper and
that can be applied in the case of thin films relates to the
change of the substrate material. As we have seen, the
characteristic thermal-diffusion time, t c , is a critical parameter that determines the magnitude of the thermal effect. A means to reduce t c , and therefore, to increase the
highest allowed repetition rate for which the thermal ef-

Nalda et al.

fect is small, is to increase the thermal-diffusion coefficient D that is related to the thermal conductivity K
through D K/ C, where is the density and C, the specific heat. In our case, the 30-fold increase of the thermal
conductivity in the sapphire substrate has suppressed the
thermal contribution. This approach can thus be used to
measure the electronic nonlinearity whenever this effect
produces a signal that is above the experimental discrimination threshold. The reason why there is no measurable peak-to-valley transmittance change in the film on
sapphire, when the shorter focal-length lens was used, is
the smaller value of the nonlinear refractive index of the
sample due to its smaller Cu content. In this respect, it
is worth noting that the expected contributions from the
nanocomposite films with lower Cu content and the sapphire substrate to the Z-scan signal (T p-v 2% and
0.01%, respectively) are either at the limit of or well below
the experimental resolution.
Recently, a smart, novel technique has been proposed
to distinguish thermal from nonlinear lensing in Z-scan
experiments when either cw or high-repetition-rate lasers
are used.48 The method is based on the use of a mechanical chopper to induce an irradiation rise time of 15 s
and recording of the on-axis far-field intensity with time
resolution. For the first 100 s after the onset of illumination, the Z-scan far-field signal is dominated by the
electronic nonlinear effect, owing to the relatively slow
buildup time of the thermal lens. For later times, the
thermal effect sets the features of the Z-scan trace. This
method should, however, be applied with care, since the
buildup time of the observation window depends on the
thermal properties and absorption of the sample, and
thus the time scales necessary to obtain reliable data can
be, in many cases, much shorter.
Since the possibility of changing the laser repetition
rate, recording time-resolved Z-scan data, or changing the
thermal conductivity of the substrate are not always
readily available, some attempts have been made to separate the thermal from the electronic contributions in the
data-analysis stage.36,45,50 The idea underlying Yangs
approach49,50 is that the behavior of the electronic lens as
a function of z differs from that of the thermal one, and
hence it is possible, in principle, to extract the contribution from each effect by a careful analysis of the Z-scan
trace. However, several difficulties are not solved in this
simplified model, and it does not allow a reliable quantitative separation of thermal and electronic effects, as we
have checked with the results presented here, even if the
qualitative trends of both are established.29 A reliable
thermal-lens model to be used in Z-scan data analysis
would need to take into account an aberrant thermal-lens
theory,39,40 the dynamics of the cooling of the material between pulses (both included recently by Falconieri,42) and
the two-dimensional aspect of the heat conduction for thin
absorbing samples.

5. CONCLUSIONS
It has been shown that the measurement of the thirdorder nonlinear optical susceptibility of absorbing
samples by the Z-scan method can be hindered by thermooptical effects arising from cumulative heating, even

Nalda et al.

when short pulses are used. Two experimental methods


are proposed to overcome this problem. The first one,
which has been proposed earlier in the literature and is
not always valid, is to reduce the laser repetition rate.
The second method, appropriate for thin films, is the use
of substrates with higher thermal conductivity.

ACKNOWLEDGMENTS
This work has been partially supported by the Comunidad Autonoma de Madrid (project 07T/0038/98) and by
the Comision Interministerial de Ciencia y Tecnologia
(TIC99-0866). R. de Nalda and R. del Coso acknowledge
financial support of the Comunidad Autonoma de Madrid
(Spain), and A. Suarez-Garca is thankful to Ministerio de
Ciencia y Tecnologa (Spain) for funding under the PN99
10894396 fellowship. J. Olivares wishes to thank the Consejo Superior de Investigaciones Cientificas (Spain) for
funding his position. We wish to thank J. M. Ballesteros
for fruitful discussion and M. Mejas for technical support.

Vol. 19, No. 2 / February 2002 / J. Opt. Soc. Am. B

17.

18.

19.

20.

21.
22.

23.

REFERENCES AND NOTES


1.

2.
3.
4.
5.

6.
7.
8.

9.
10.
11.
12.
13.
14.
15.
16.

M. Sheik-Bahae, A. A. Said, T. Wei, D. J. Hagan, and E. W.


Van Stryland, Sensitive measurement of optical nonlinearities using a single beam, IEEE J. Quantum Electron.
26, 760769 (1990).
R. L. Sutherland, ed., Handbook of Nonlinear Optics (Marcel Dekker, New York, 1996).
S. R. Friberg and P. W. Smith, Nonlinear optical glasses for
ultrafast optical switches, IEEE J. Quantum Electron. QE23, 20892094 (1987).
R. Adair, L. L. Chase, and S. A. Payne, Nonlinear
refractive-index measurements of glasses using three-wave
frequency mixing, J. Opt. Soc. Am. B 4, 875881 (1987).
M. J. Moran, C. Y. She, and R. L. Carman, Interferometric
measurements of the nonlinear refractive-index coefficient
relative to CS2 in laser-system related materials, IEEE J.
Quantum Electron. QE-11, 259265 (1975).
A. Owyoung, Ellipse rotation studies in laser host materials, IEEE J. Quantum Electron. QE-9, 10641071 (1973).
P. D. Maker and R. W. Terhune, Study of optical effects due
to an induced polarization third order in electric field
strength, Phys. Rev. 137, A801A818 (1965).
W. E. Williams, M. J. Soileau, and E. W. Van Stryland, Optical switching and n 2 measurements in CS2 , Opt. Commun. 50, 256260 (1984).
Y. Bae, J. J. Song, and Y. B. Kim, Photoacoustic study of
two-photon absorption in hexagonal ZnS, J. Appl. Phys. 53,
615619 (1982).
S. M. Mian, B. Taheri, and J. P. Wicksted, Effects of beam
ellipticity on z-scan measurements, J. Opt. Soc. Am. B 13,
856863 (1996).
Y. L. Huang and C. K. Sun, Z-scan measurement with an
astigmatic Gaussian beam, J. Opt. Soc. Am. B 17, 4347
(2000).
P. Chen, D. A. Oulianov, I. V. Tomov, and P. M. Rentzepis,
Two- dimensional z-scan for arbitrary beam shape and
sample thickness, J. Appl. Phys. 85, 70437050 (1999).
P. B. Chapple, J. Staromlynska, and R. G. McDuff, Z-scan
studies in the thin- and the thick-sample limits, J. Opt.
Soc. Am. B 11, 975982 (1994).
M. Sheik-Bahae, A. A. Said, D. J. Hagan, M. J. Soileau, and
E. W. Van Stryland, Nonlinear refraction and optical limiting in thick media, Opt. Eng. 30, 12281235 (1991).
S. Bian, M. Martinelli, and R. J. Horowicz, Z-scan formula
for saturable Kerr media, Opt. Commun. 172, 347353
(1999).
A. Marcano, O. H. Maillotte, D. Gindre, and D. Metin, Pi-

24.
25.

26.

27.

28.

29.

30.
31.

32.

33.

34.

295

cosecond nonlinear refraction measurement in single-beam


open z-scan by charge-coupled device image processing,
Opt. Lett. 21, 101103 (1996).
F. E. Hernandez, A. Marcano, and H. Maillotte, Sensitivity
of the total beam profile distortion z-scan for the measurement of nonlinear refraction, Opt. Commun. 134, 529536
(1997).
T. Xia, D. J. Hagan, M. Sheik-Bahae, and E. W. Van Stryland, Eclipsing z-scan measurement of /104 wave-front
distortion, Opt. Lett. 19, 317319 (1994).
A. Marcano, F. E. Hernandez, and A. D. Sena, Two-color
near-field eclipsing z-scan technique for the determination
of nonlinear refraction, J. Opt. Soc. Am. B 14, 33633367
(1997).
K. Uchida, S. Kaneko, S. Omi, C. Hata, H. Tanji, Y. Asahara, A. J. Ikushima, T. Tokizaki, and A. Nakamura, Optical nonlinearities of a high concentration of small particles
dispersed in glass: copper and silver particles, J. Opt.
Soc. Am. B 11, 12361243 (1994).
D. Ricard, P. Roussignol, and C. Flytzanis, Surfacemediated enhancement of optical phase conjugation in
metal colloids, Opt. Lett. 10, 511513 (1985).
H. B. Liao, R. F. Xiao, J. S. Fu, and G. K. L. Wong, Large
third order nonlinear optical susceptibility of Au Al2 O3
composite films near the resonant frequency, Appl. Phys. B
65, 673676 (1997).
R. H. Magruder III, L. Yang, R. F. Haglund, Jr., C. W.
White, L. Yang, R. Dorsinville, and R. R. Alfano, Optical
properties of gold nanocluster composites formed by deep
ion implantation in silica, Appl. Phys. Lett. 62, 17301732
(1993).
F. Hache, D. Richard, C. Flytzanis, and U. Kreibig, The optical Kerr effect in small metal particles and metal colloids:
the case of gold, Appl. Phys. A 47, 347357 (1988).
J. M. Ballesteros, J. Sols, R. Serna, and C. N. Afonso,
Nanocrystal size dependence of the third order nonlinear
optical response of Cu:Al2 O3 thin films, Appl. Phys. Lett.
74, 27912793 (1999).
M. Falconieri, G. Salvetti, E. Cattaruzza, F. Gonella, G.
Mattei, P. Mazzoldi, M. Piovesan, G. Battaglin, and R. Polloni, Large third order optical nonlinearity of nanoclusterdoped glass formed by ion implantation of copper and nickel
in silica, Appl. Phys. Lett. 73, 288290 (1998).
Y. Takeda, T. Hioki, T. Motohiro, and S. Noda, Large thirdorder optical nonlinearity of tin microcrystallite-doped
silica glass formed by ion implantation, Appl. Phys. Lett.
63, 34203422 (1993).
M. Sheik-Bahae, A. A. Said, and E. W. Stryland, Highsensitivity, single-beam n 2 measurements, Opt. Lett. 14,
955957 (1989).
J. M. Ballesteros, R. Serna, J. Sols, C. N. Afonso, A. K.
Petford-Long, D. H. Osborne, and R. F. Haglund, Jr.,
Pulsed laser deposition of Cu:Al2 O3 nanocrystal thin films
with high third order optical susceptibility, Appl. Phys.
Lett. 71, 24452447 (1997).
L. R. Doolittle, Algorithms for the rapid simulation of Rutherford backscattering spectra, Nucl. Instrum. Methods B
9, 344351 (1985).
R. Serna, C. N. Afonso, C. Ricolleau, Y. Wang, Y. Zheng, M.
Gandais, and I. Vickridge, Artificially nanostructured
Cu:Al2 O3 films produced by pulsed laser deposition, Appl.
Phys. A 71, 583586 (2000).
A. Naudon and D. Thiaudie`re, Grazing-incidence smallangle scattering. Morphology of deposited clusters and
nanostructure of thin films, J. Appl. Crystallogr. 30, 822
827 (1997).
A. Naudon, D. Babonneau, D. Thiaudie`re, and S. Lequien,
Grazing-incidence small-angle X-ray scattering applied to
the characterization of aggregates in surface regions,
Physica B 283, 6974 (2000).
Peak power refers to the maximum power of the laser pulse
and the mean power to the total power of the train of
pulses. Intensity refers to the energy density, i.e., the
power of a laser pulse (peak or total) divided by the area of
the laser beam.

296
35.

36.

37.
38.

39.

40.
41.

42.
43.

J. Opt. Soc. Am. B / Vol. 19, No. 2 / February 2002


R. F. Haglund, Jr., L. Yang, R. H. Magruder, III, J. E. Wittig, K. Becker, and R. A. Zuhr, Picosecond nonlinear optical
response of a Cu:silica nanocluster composite, Opt. Lett.
18, 373375 (1993).
L. Yang, D. H. Osborne, R. F. Haglund, Jr., R. H. Magruder,
C. W. White, R. A. Zuhr, and H. Hosono, Probing interface
properties of nanocomposites by third order nonlinear optics, Appl. Phys. A 62, 403415 (1996).
D. Weaire, B. S. Wherrett, D. A. B. Miller, and S. D. Smith,
Effect of low-power nonlinear refraction on laser-beam
propagation in InSb, Opt. Lett. 4, 331333 (1974).
T. Kawazoe, H. Kawaguchi, J. Inoue, O. Haba, and M.
Ueda, Measurement of nonlinear refractive index by timeresolved z-scan technique, Opt. Commun. 160, 125129
(1999).
M. L. Baesso, J. Shen, and R. D. Snook, Mode-mismatched
thermal lens determination of temperature coefficient of optical path length in soda lime glass at different wavelengths, J. Appl. Phys. 75, 37323737 (1994).
S. J. Sheldon, L. V. Knight, and J. M. Thorne, Laser induced thermal lens effect: a new theoretical model, Appl.
Opt. 21, 16631669 (1982).
T. Tokizaki, A. Nakamura, S. Kaneko, K. Uchida, S. Omi, H.
Tanji, and Y. Asahara, Subpicosecond time response of
third order optical nonlinearity of small copper particles in
glass, Appl. Phys. Lett. 65, 941943 (1994).
M. Falconieri, Thermo-optical effects in z-scan measurements using high-repetition-rate lasers, J. Opt. A: Pure
Appl. Opt. 1, 662667 (1999).
S. Vijayalakshmi, H. Grebel, Z. Iqbal, and C. W. White, Ar-

Nalda et al.

44.

45.

46.
47.

48.

49.

50.

tificial dielectrics: nonlinear properties of Si nanoclusters


formed by ion implantation in SiO2 glassy matrix, J. Appl.
Phys. 84, 65026506 (1998).
X. Zhu and Z. Meng, The optical nonlinearity and structure for a PbO, TiO2 , SiO2 and K2 O quaternary glass systems, J. Appl. Phys. 75, 37563760 (1994).
A. A. Andrade, E. Tenorio, T. Catunda, M. L. Baesso, A. Cassanho, and H. P. Jenssen, Discrimination between electronic and thermal contributions to the nonlinear refractive
index of SrAlF5 :Cr3 , J. Opt. Soc. Am. B 16, 395400
(1999).
Y. M. Cheung and S. K. Gayen, Optical nonlinearities of
tea studied by z-scan and four-wave mixing techniques, J.
Opt. Soc. Am. B 11, 636643 (1994).
B. Taheri, A. F. Munoz, W. D. St. John, J. P. Wicksted, R. C.
Powell, D. H. Blackburn, and D. C. Cranmer, Effects of the
structure and composition of lead glasses on the thermal
lensing of pulsed radiation, J. Appl. Phys. 71, 36933700
(1992).
M. Falconieri and G. Salvetti, Simultaneous measurement
of pure-optical and thermo-optical nonlinearities induced
by high-repetition-rate, femtosecond laser pulses: application to CS2 , Appl. Phys. B 69, 133136 (1999).
R. F. Haglund, Handbook of Optical Properties. Vol. II:
Optics of Small Particles, Interfaces and Surfaces, R. E.
Hummel and P. Wissmann, eds. (CRC Press, Boca Raton,
Fla., 1997).
L. Yang, Nonlinear optics of noble metal crystallites embedded in a dielectric matrix, Ph.D. dissertation (Vanderbilt University, Nashville, Tenn., 1993).

Vous aimerez peut-être aussi