You are on page 1of 10

Polarization dependence of Z-scan measurement:

theory and experiment


Xiao-Qing Yan, Zhi-Bo Liu, Xiao-Liang Zhang, Wen-Yuan Zhou, Jian-Guo Tian*
The Key Laboratory of Weak Light Nonlinear Photonics, Ministry of Education, Teda Applied Physics School, Nankai
University, Tianjin 300457, China
*Corresponding author: jjtian@nankai.edu.cn

Abstract: Here we report on an extension of common Z-scan method to


arbitrary polarized incidence light for measurements of anisotropic third-
order nonlinear susceptibility in isotropic medium. The normalized
transmittance formulas of closed-aperture Z-scan are obtained for linearly,
elliptically and circularly polarized incidence beam. The theoretical analysis
is examined experimentally by studying third-order nonlinear susceptibility
of CS2 liquid. Results show that the elliptically polarized light Z-scan
method can be used to measure simultaneously the two third-order nonlinear
susceptibility components χxyyx (3) and (3) .
χ xxyy Furthermore, the elliptically
polarized light Z-scan measurements of large nonlinear phase shift are also
analyzed theoretically and experimentally.
©2009 Optical Society of America
OCIS codes: (190.0190) Nonlinear optics; (190.4710) Optical nonlinearities in organic
materials, (190.3270) Kerr effects.

References and links


1. I. Fuks-Janczarek, B. Sahraoui, I. V. Kityk, and J. Berdowski, “Electronic and nuclear contributions to the
third-order optical susceptibility,” Opt. Commun. 236, 159 (2004).
2. G. Boudebs, M. Chis, and J. P. Bourdin, “Third-order susceptibility measurements by nonlinear image
processing,” J. Opt. Soc. Am. B 13, 1450–1456 (1996).
3. P. D. Maker, R. W. Terhune, and C. M. Savage, “Intensity-dependent changes in the refractive index of
liquids,” Phys. Rev. Lett. 12, 507–509 (1964).
4. M. Lefkir and G. Rivoire, "Influence of transverse effects on measurement of third-order nonlinear
susceptibility by self-induced polarization state changes," J. Opt. Soc. Am. B 14, 2856-2864 (1997).
5. M. Sheik-Bahae, A. A. Said, T. H. Wei, D. J. Hagan, and E. W. Van Stryland, “Sensitive measurement of
optical nonlinearities using a single beam,” IEEE J. Quantum Electron. 26, 760–769 (1990).
6. P. B. Chapple, J. Straromlynska, J. A. Hermann, T. J.Mckay, and R. G. McDuff, ‘‘Single-beam Z-scan:
measurement techniques and analysis,’’ J. Nonlinear Opt. Phys. Mater. 6, 251–293 (1997).
7. E. W. Van Stryland and M. Sheik-Bahae, “Z-scan Measurements of Optical Nonlinearities,” in
Characterization Techniques and Tabulations for Organic Nonlinear Materials, M. G. Kuzyk and C. W.
Dirk, Eds., page 655-692, Marcel Dekker, Inc., 1998.
8. J. A. Hermann and R. G. McDuff, “Analysis of spatial scanning with thick optically nonlinear media,” J.
Opt. Soc. Am. B 10, 2056-2064 (1993).
9. J. G. Tian, W. P. Zang, C. Z. Zhang, and G. Y. Zhang, “Analysis of beam propagation in thick nonlinear
media,” Appl. Opt. 34, 4331-4336 (1995).
10. T. Xia, D. J. Hagan, M. Sheik-Bahae, and E. W. Van Stryland, “Eclipsing Z- Scan Measurement of λ/104
Wavefront Distortion,” Opt. Lett. 19, 317-319 (1994).
11. J. Wang, M. Sheik-Bahae, A. A. Said, D. J. Hagan, and E. W. Van Stryland, “Time-Resolved Z-Scan
Measurements of Optical Nonlinearities,” J. Opt. Soc. Am. B 11, 1009-1017 (1994).
12. L. Demenicis, A. S. L. Gomes, D. V. Petrov, C.B. de Araújo, C. P. de Melo, C. G. dos Santos, and R. Souto-
Maior, “Saturation effects in the nonlinear-optical susceptibility of poly(3-hexadecylthiophene),” J. Opt.
Soc. Am. B 14, 609 (1997).
13. M. Sheik-Bahae, J. Wang, J.R. DeSalvo, D. J. Hagan and E. W. Van Stryland, “Measurement of
Nondegenerate Nonlinearities using a 2-Color Z-Scan,” Opt. Lett. 17, 258-260 (1992).
14. R. Bridges, G. Fischer, and R. Boyd, “Z-scan measurement technique for non-Gaussian beams and arbitrary
sample thicknesses,” Opt. Lett. 20, 1821-1823 (1995).
15. W. Zhao and P. Palffy-Muhoray, “Z-scan measurements of χ3 using top-hat beams,” Appl. Phys. Lett. 65,
673-675 (1994).

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6397
16. R. DeSalvo, M. Sheik-Bahae, A. A. Said, D. J. Hagan, and E. W. Van Stryland, “Z-scan measurements of
the anisotropy of nonlinear refraction and absorption in crystals,” Opt. Lett. 18, 194 (1993).
17. S. J. Wagner, J. Meier, A. S. Helmy, J. S. Aitchison, D. Modotto, M. Sorel, and D. C. Hutchings,
"Polarization-Dependent Nonlinear Refraction in GaAs/AlAs Superlattice Waveguides," in Frontiers in
Optics, OSA Technical Digest (CD) (Optical Society of America, 2006), paper FWE2.
18. J. Liang, H. Zhao, and X. Zhou, “Polarization-dependence effects of refractive index change associated with
photoisomerization investigated with Z-scan technique,” J. Appl. Phys. 101, 013106 (2007).
19. R. W. Boyd, Nonlinear Optics, 2nd ed. (Academic Press, San Diego, 2003).
20. Z. B. Liu, X. Q. Yan, J. G. Tian, W. Y. Zhou, and W. P. Zang. “Nonlinear ellipse rotation modified Z –scan
measurements of third-order nonlinear susceptibility tensor,” Opt.Exp. 15, 13351 (2007).
21. Z. B. Liu, X. Q. Yan, W. Y. Zhou, and J. G. Tian, “Evolutions of polarization and nonlinearities in an
isotropic nonlinear medium,” Opt. Express 16, 8144 (2008).
22. S. Q. Chen, Z. B. Liu, W. P. Zang, J. G. Tian, W. Y. Zhou, F. Song, and C. P. Zhang, “Study on Z-scan
characteristics for a large nonlinear phase shift,” J. Opt. Soc. Am. B 22, 1911 (1997).
23. Z. B. Liu, Y. L. Liu, B. Zhang, W. Y. Zhou, J. G. Tian, W. P. Zang, and C. P. Zhang , “Nonlinear absorption
and optical limiting properties of carbon disulfide in a short-wavelength region,” J. Opt. Soc. Am. B 24,
1101 (2007).
24. K. Kiyohara, K. Kamada, and K. Ohta, “Orientational and collision-induced contribution to third-order
nonlinear optical response of liquid CS2,” J. Chem. Phys. 112, 6338, (2000).
25. R. L. Sutherland, Handbook of Nonlinear Optics (Second Edition); (Marcel Dekker: New York, 2003).
26. R. Volle, V. Boucher, K. D. Dorkenoo, R. Chevalier and X. N. Phu, “Local polarization state observation
and third-order nonlinear susceptibility measurements by self-induced polarization state changes method,”
Opt. Commun. 182, 443 (2000).
27. R.A. Ganeev, A.I. Ryasnyansky, M. Suzuki, N. Ishizawa, M. Turu, S. Sakakibara, and H. Kuroda,
“Nonlinear refraction in CS2,” Appl. Phys. B 78, 433 (2004).
28. http://en.wikipedia.org/wiki/CS2
29. B. Gu, J. Chen, Y. Fan, J. Ding, and H. Wang, "Theory of Gaussian beam Z scan with simultaneous third-
and fifth-order nonlinear refraction based on a Gaussian decomposition method," J. Opt. Soc. Am. B 22,
2651-2659 (2005).
30. R. A. Ganeev, M. Baba, M. Morita, A. I Ryasnyansky, M. Suzuki, M. Turu, and H. Kuroda, “Fifth-order
optical nonlinearity of pseudoisocyanine solution at 529 nm,” J. Opt. A: Pure Appl. Opt. 6, 282–287 (2004).

1. Introduction
There are a variety of experimental methods for measurements of third-order nonlinear
susceptibility, such as degenerate four-wave mixing [1], nonlinear imaging techniques [2],
nonlinear ellipse rotation [3,4] and Z-scan method [5]. Among the methods used for
determination of third-order nonlinear susceptibilities, the very efficient technique is
degenerate four-wave mixing, which allows not only to determine the different tensor
components, but also effectively to separate the electronic and nuclear contribution for
isotropic materials including the CS2. However, this method is much complicated compared
with single beam Z-scan method. The well-known Z-scan method has become a standard tool
for determining nonlinear parameters of various materials because of its simplicity,
sensitivity, accuracy, and the ease of separation between nonlinear refraction (NLR) and
nonlinear absorption (NLA) [5-7]. Since the pioneering experimental work of M. Sheik-Bahae
et al [5], there have been various theoretical and experimental modifications of this method,
such as thick sample Z-scan [8,9], eclipsing Z-scan [10], pump-probe Z-scan [11,12], two-
color Z-scan[13], non-Gaussian beams Z-scan [14], top-hat beam Z-scan[15], etc. However,
among these Z-scan methods the linearly polarized lights are always used as incidence beams,
and there is no theoretical and experimental work to study the Z-scan method with elliptically
and circularly polarized lights.
For the effect of polarization on Z-scan measurements, R. DeSalvo et al presented the
work on polarization dependence of nonlinear refractive index in crystals [16], in which they
measured the transmittance by altering the linear polarization direction of incidence beam
relative to crystallographic axis when crystal was stationary in the Z direction. Sean J. Wagner
et al studied the polarization-dependent nonlinear refraction and two-photon absorption in
GaAs/AlAs superlattice waveguides [17]. However, the incidence lights still were linearly
polarized in their studies. Recently, Liang et al used different polarized beams to study the
polarization dependence effects of refractive index change (RIC) associated with
photoisomerization in poly (methyl methacrylate) (PMMA) film by Z-scan method [18], in

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6398
which the peak-valley difference ∆Tp-v with linearly polarized light was larger than those with
circularly and elliptically polarized lights. They tried to explain it by decomposing the
polarized light into two perpendicular linear polarized components, but they did not
quantitatively analyze the effect of the RIC due to the appearance of perturbation of
polarization ellipse and did not give the relationship between normalized transmittance and
nonlinear phase shift as in Ref. 5.
In an isotropic medium, the third-order susceptibility tensor has only two independent
third-order susceptibility components, namely χxxyy
(3) , and (3) . For molecular orientation and
χ xyyx
nonresonant electronic nonlinearities, RIC is greatly dependent on polarization state due to the
anisotropy of third-order nonlinear susceptibility [19]. By contraries, for thermal induced
nonlinearity, excited state nonlinearity and electrostriction, RIC is non-dependent on
polarization state. Therefore, in order to obtain more information of third-order nonlinear
susceptibility tensor and understand the nonlinear mechanisms from RIC, elliptically and
circularly polarized lights should be used in Z-scan measurements. We have demonstrated a
method that combines the Z-scan technique with nonlinear ellipse rotation to measure third-
order nonlinear susceptibility components [20,21]. However, to the best of our knowledge,
there is no report on analytical expression of Z-scan normalized transmittance under the
condition of elliptically or circularly polarized lights. In this paper, we theoretically and
experimentally study the polarization dependence of Z-scan measurements. The normalized
transmittance formula of closed-aperture Z-scan is obtained for linearly, elliptically and
circularly polarized incidence beams. The normalized transmittances of CS2 for linearly,
elliptically and circularly polarized incidence beams were measured and analyzed
theoretically. The experimental results agree well with theoretical analyses. Further, the
elliptically polarized light Z-scan was used to study intensity dependence of normalized
transmittance.

ϕ1
f
BS
D2
s
Sample

polarizer λ/4 Lens z Aperture

D1

Fig. 1. Experimental arrangements for polarization changeable Z-scan

2. Theory
The modified Z-scan experimental arrangement is shown in Fig.1. Compared with normal Z-
scan method, a quarter-wave plate is placed between the polarizer and the sample to create
different polarized lights. When a linearly polarized beam passes the λ/4 plate with angle ϕ1
( -π/2≤ϕ1<π/2) which is the angle between the linear polarization direction and the slow axis
of the λ/4 plate, it can be converted into a polarized beam (ϕ1=-π/2, 0 for linearly polarized
beam, ϕ1=±π/4 for circularly polarized beam, and others for elliptically polarized beam).
Assuming a TEM00 Gaussian beam of waist radius w0 traveling in the +z direction, we can
write E as
  w r2 iκ r 2
E ( z, r , t ) = E0 (t ) 0 exp(− 2 − − iφ ( z , t )) . (1)
wz wz 2R( z)

where wz2 = w02 (1 + z 2 z02 ) is the beam radius, R( z ) = z (1 + z02 z 2 ) is the radius of curvature of the
wave front at z, z 0 = κ w02 2 is the diffraction length of the beam, and κ = 2π λ is the wave

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6399

vector. E0 (t ) denotes the radiation electric field vector at the focus and contains the temporal
envelope of the laser pulse. The exp[ −iφ ( z , t )] term contains all the radically uniform phase
variations.
If we define the slow axis of the λ/4 plate as x-axis, after passing the λ/4 plate, the electric
field can be written as
 w r2 κr2 (2)
E ( z, r , t ) = ( E (t )cos ϕ xˆ + E (t )sin ϕ e−iδ1 yˆ ) 0 exp(− − i
0 1 0 1 − iφ ( z, t )) ,
wz wz2 2R( z )

where δ1 is the phase retardation, x̂ and ŷ are unit vectors. δ1=π/2 for a λ/4 plate. The
electric field vector of such a beam can always be decomposed into a linear combination of
left- and right-hand circular components as [19]

E ( z, r , t ) = E+ ( z, r , t )σˆ + + E− ( z, r , t )σˆ −
. (3)
w r2 π r2
= ( E+,0 (t )σˆ + + E−,0 (t )σˆ − ) 0 exp(− 2 − i − iφ ( z , t )),
wz wz λ R( z )

where σˆ + = ( xˆ + iyˆ ) / 2 and σˆ − = ( xˆ − iyˆ ) / 2 are the circular-polarization unit vectors,


E+ = ( Ex − iE y ) / 2 , E− = ( Ex + iE y ) / 2 . The other two parameters can be expressed as
 E+ ,0 (t ) = E0 (t )(cos ϕ1 − ie−iδ1 sin ϕ1 ) 2. (4)
 − iδ
 E−,0 (t ) = E0 (t )(cos ϕ1 + ie 1 sin ϕ1 ) 2

The ellipticity used to describe the polarization geometry is defined as


i.e. e = sin δ sin 2ϕ /(1 + 1 − sin δ sin 2ϕ ) . e is 1 for circularly
2 2
e = E+ − E− /( E + + E − ) , 1 1 1 1

polarized light, and 0 for linearly polarized light.


Because of the nonlinear effect in medium, the refractive indexes of the two circular
components in the medium are different and given by [19]:
2π 
A E± + ( A + B) E∓  , (5)
2 2
n± = n0 +
n0  

(3) and
where n0 is the linear refractive index, A = 6 χ xxyy (3) . Substituting Eq. (4) into Eq.
B = 6 χ xyyx
(5), we can rewrite Eq. (5) as follows:
2π  1 ± sin δ1 sin 2ϕ1 
B  E ( z, r , t ) . (6)
2
n± ( z, r , t ) = n0 + ∆n± = n0 + A+
n0  2 
Due to the change in refractive index δn± is different for two components, the left- and
right-hand circular components propagate with different phase velocities, thus, the
polarization direction will rotate as the elliptically polarized beam propagates through the
nonlinear medium. This is called as nonlinear ellipse rotation [19,3].
Considering a thin medium and using the slowly varying envelope approximation, we can
separate the beam propagation equations with the two circular components into a couple of
equations, one for the phase and another for the irradiance, as follows:
 d ∆φ
 dz ' = κ∆n( I ) , (7)

 dI = −α ( I ) I
 dz '

where z ' is the propagation depth in the sample. If nonlinear absorption of the medium is
negligible, the phase shifts at the exit surface for the two circular components can be obtained
by substituting Eq. (6) into Eq. (7) and solving the equations:

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6400
 2r 2
∆φ+ ( z, r , t ) = ∆φ+ ,0 ( z, t ) exp(− 2 ) , (8)
 wz
 2
∆φ ( z, r , t ) = ∆φ ( z, t ) exp(− 2r )
 − − ,0
wz2

with
∆φ± ,0 ( z , t ) = ∆φ± ,0 (t ) (1 + z 2 z02 ) . (9)

where ∆φ+ ,0 (t ) and ∆φ− ,0 (t ) are on-axis phase shift at the focus of left- and right-hand circular
components, respectively. They are defined as
∆φ± ,0 (t ) = κ∆n± ,0 (t ) Leff . (10)

where Leff=[1-exp(-αL)]/α is the effective length, L is the sample length, α is the linear
absorption coefficient and ∆n± ,0 (t ) = 2π [ A + (1 ± sin δ1 sin 2ϕ1 ) B / 2] E0 (t ) 2 / n0 are the on-axis RICs
of the two circular components at the focal plane.
Ignoring Fresnel reflection losses and diffraction effect, we can obtain the resultant
complex electric field pattern
 at the exit surface of the medium as [5]
Ee ( z, r , t ) = Ee,+ ( z, r , t )σˆ + + Ee,− ( z, r , t )σˆ − . (11)
= E+ ( z , r , t )e−α L / 2e−i∆φ+ ( z ,r ,t )σˆ + + E− ( z, r , t )e−α L / 2e−i∆φ− ( z ,r ,t )σˆ −
We use Gaussian decomposition (GD) method to calculate the complex field pattern at the
aperture plane. The nonlinear phase term exp[ −i∆φ± ( z, r , t )] in Eq. (11) can be expanded as
m
∞  −i∆φ± ,0 ( z, t )  2
e−i∆φ± ( z ,r ,t ) = ∑   exp(− 2mr ) .
2
(12)
m =0 m! wz

The complex field pattern at the aperture is expressed as



Ea ( r , t ) = Ea , + (r , t )σˆ + + Ea ,− (r , t )σˆ −
m
. (13)
∞  −i∆φ+ ,0 ( z , t )  wm 0 r 2 iκ r 2
= E+ ( z , r = 0, t )e −α L / 2

m=0 m! wm
exp( − 2 −
wm 2 Rm
+ iθ m )σˆ +
m
∞  −i ∆φ
− ,0 ( z , t ) 
 wm 0 exp(− r − iκ r + iθ )σˆ
2 2
+ E− ( z , r = 0, t )e −α L / 2 ∑  2 m −
m=0 m! wm wm 2 Rm
Defining d as the propagation distance in free space from the medium to the aperture
plane, and g = 1 + d / R ( z ) , the other parameters in Eq. (13) are expressed as
wm2 0 = wz2 /(2m + 1) (14)
d m = kwm2 0 / 2 (15)
g −1 (16)
Rm = d (1 − )
g 2 + d 2 / d m2

θ m = tan −1 (
d / dm
)
(17)
g
wm2 = wm2 0 ( g 2 + d 2 / d m2 ) (18)

The normalized transmittance can be calculated as


∞ ra
∫ dt ∫ Ea (r , t ) rdr
2
, (19)
T (ra , z , t ) = ∞
−∞

0
2
S ∫ dt ∫ Ea ( r , t , ∆φ± ,0 = 0) rdr
−∞ 0

where ra is the radius of the aperture, S = 1 − exp(−2ra2 / wa2 ) is the aperture linear transmittance,
wa is the beam radius at the aperture in the linear regime. Under the far-field condition of
z0<<d, the normalized transmittance of Eq. (19) can be calculated as:

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6401
1 ∞ ∞ 1 π π
∑∑ (20)
2
T= Gmn {cos(cmn ) − e − bmnYa ⋅ cos(cmn + d mnYa2 )}
S m =0 n = 0 (m + n + 1) 2 2
with
m+ n m+ n
1 − sin δ1 sin 2ϕ1  ∆φ+,0 ( z, t )  1 + sin δ1 sin 2ϕ1  ∆φ−,0 ( z, t )  (21)
Gmn = +
2 m!n! 2 m!n!
2m + 1 2n + 1 (22)
bmn = (1 + x2 )[ + ]
x 2 + (2m + 1)2 x2 + (2n + 1)2
4 x(1 + x 2 )(m − n)(m + n + 1) (23)
d mn =
[ x 2 + (2m + 1)2 ][ x 2 + (2n + 1)2 ]
cmn = m − n (24)
S = 1 − exp(−2Ya2 ) (25)

where Ya=ra/D ω0 is a dimensionless aperture radius, x=z/z0 is the dimensionless sample


position, and D=d/z0 is the dimensionless distance from the sample to the aperture plane.
From Eq. (21), it is seen that there is no nonlinear phase shifts coupling terms of left- and
right- hand lights, which means the two circular components independently propagate and
reach the detector. The transmittance at the detector plane is the weighted average of the two
components. By using L’Hôpital’s rule, the normalized transmittance for closed-aperture
(CA) case can be obtained from Eq. (20) as
1 ∞ ∞ Gmn π π . (26)
TCA = ∑∑
2 m=0 n=0 (m + n + 1)
{bmn cos(cmn
2
) + d mn sin(cmn
2
)}

Under first-order approximation, TCA can be written as


4x 1 − sin δ1 sin 2ϕ1 1 + sin δ1 sin 2ϕ1
TCA = 1 + [ ∆φ+,0 (t ) + ∆φ−,0 (t )] (27)
( x 2 + 1)( x 2 + 9) 2 2
4x
= 1+ 2 ∆φ
( x + 1)( x 2 + 9) eff

where ∆φeff = kπ [2 A + (1 − sin 2 δ1 sin 2 2ϕ1 ) B] E0 (t ) 2 Leff / n0 is effective nonlinear phase shift at the
focus on the Z axis. If the first quarter-wave plate is taken away or the angle ϕ1 is set at 0 deg,
above formula will return to the normal Z-scan transmittance formula in Ref. 5. Only the
weighted average RICs of the two components can be obtained if Eq. (27) is used to analyze
the experimental data no matter what polarization state of the incidence beam is. If we want to
separate A and B values from one experiment, one possible way is to abandon first-order
approximation by taking more terms in Eqs. (26) and (20) as normalized transmittance fitting
formula because the coefficients of high-order terms are non-correlation. According to Chen
et al’s works [22], normalized transmittance formulas of Eq. (20) obtained by GD method can
also be available for large nonlinear phase shifts.
As in Ref. 5, the steady-state (the same as cw situation) results can be extended to transient
effects induced by pulsed radiation. For a Gaussian pulse, the average refractive index is
〈∆n± ,0 (t )〉 = ∆n± ,0 / 2 (28)

where ∆n±,0 is the peak-on-axis index change at the focus.


3. Experimental result and discussions
In our experiments, a frequency-doubled Continuum Model PY61 mode-locked pulsed laser
was used to generate 35 ps pulses at 532 nm with a repetition rate of 10 Hz. The laser pulse
focused by a lens of 150 mm focal length to produce a beam waist w0 of 18 µm was incident
to a 1 mm quartz cell containing CS2. The on-axis peak intensity I0 was 5.23 GW/cm2. The
extensively studied isotropic nonlinear medium, CS2, was chosen as the sample. CS2 is
transparent and has very small linear absorption coefficients α0 (<10-3 cm-1) in the visible
region [23], and the nonlinear absorption coefficient is less than 10-11 cm/W, so we can ignore

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6402
the nonlinear absorption in our experimental conditions. The physical mechanism leading to
nonlinear refraction of CS2 in subnansecond regime is molecular orientation, and the ratio of
B to A should be 6 [19]. In order to make the reference light be synchronous with the signal
light, we placed the beam splitter between the λ/4 plate and the lens. Polarization states could
be changed by altering the λ/4 plate. These experiments were carried out at room temperature.
First, polarization states dependence of Z-scan were examined. From Eq. (26), it can be
seen that the normalized transmittance T of Z-scan is dependent on the angle ϕ1, i.e., the
ellipticity e. Figure 2 gives the CA Z-scan curves of linear, circular, and elliptical
polarizations of CS2. The angle ϕ1 was set 78 degree (i.e. e=0.2126) to create an elliptically
polarized light. It can be seen that the NLR is largest for a linearly polarized beam and
smallest for a circularly polarized beam, which is consistent with Liang’s experimental result
[18]. The lines are least-squares fit of Eq. (26) (terms were taken until m+n=3) to
experimental results, the theoretical fits are in good agreement with experimental results.
1.3
Lin
Lin-fit
Ell
Ell-fit
1.2
Cir
Cir-fit

1.1
Normalized transmittance

0.9

0.8

0.7

0.6
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Sample position x

Fig. 2. Normalized transmittance curves for linearly, circularly and elliptically polarized
(ϕ1=78deg) incidence beams.

Since the RIC of linearly polarized light is given by ∆ n = π (2 A + B ) E 2 / n0 [19], we can


not distinguish the B and A values (only 2A+B is obtained) from linearly polarized light Z-
scan. The ( A + B / 2) from the fit of linear polarization component was 9.232×10-12 esu. For
circular polarization case, from Eq. (5) the RIC equals to 2π A E / n0 for both left- and right-
2

hand lights, thus, only one value A can be obtained from the fit. The A from the fit in Fig. 2 is
2.241 ×10-12 esu (i.e. χ xxyy
(3)
= 3.735 × 10−13 esu ). Comparing the two cases, we could obtain B
=1.398×10-11 esu (i.e. χ xyyx (3)
= 2.330 × 10−12 esu ). The ratio B to A is 6.240, which agrees very well
with theoretical value for molecular orientation. The ratio of B to A is 6 for molecular
orientation, 1 for nonresonant electronic response, and 0 for electrostriction [19].
However, the RICs of elliptically polarized light are different for the two circular
components, inducing different nonlinear phase shifts. The nonlinear phase shifts of the two
components could be separately obtained by using Eq. (26) to fit the experimental data,
further the B and A values could be determined from Eq. (10). The fitting values of B and A
for elliptically polarized case is 1.313×10-11 and 2.182×10-12 esu respectively (i.e.
χ xyyx
(3)
= 2.189 × 10−12 esu , and χ xxyy
(3)
= 3.637 × 10 −13 esu ). The ratio B to A was 6.019. The A and B
values from these fits agree well with theoretical simulation [24] and previous measurements
[3, 26].

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6403
Although the total effective RIC obtained by using first-order approximation Eq. (27) is
different for three kinds of polarized lights, the third-order nonlinear susceptibility ( χ = ( 3)

2 χ + χ ) obtained from any polarized light experiment is identical. Indeed the polarization
( 3)
xxyy
( 3)
xyyx

state of the light just changes the contribution of B to nonlinear refractive index as shown in
Eq. (6) since both polarization state and nonlinear refractive index are determined by angle ϕ1.
As shown in above, if the incidence light is elliptically polarized, the A and B values could
be simultaneously obtained from single experiment by taking more terms in Eq. (26) during
experimental data analysis. Furthermore the ratio of B to A can be used to determine the
nonlinear mechanism. However the values of A and B can not be simultaneously obtained
from circularly or linearly polarized light Z-scan no matter how many terms are taken for the
transmittance fitting formula. Thus, circularly and linearly polarized light have to be used
together in Z-scan experiments as in Ref. 20 if one wants to measure χ xxyy and χ xyyx . It can be
( 3) ( 3)

easily understood from Eq. (21), the condition for determining A and B is that ∆φ+ ,0 (t ) and
∆φ− ,0 (t ) should be simultaneously obtained from fitting and they must be different, only
elliptically polarized light can satisfy this requirement. Compared with circularly and linearly
polarized lights Z-scans, the advantage of elliptically polarized light Z-scan is that two
nonlinear susceptibility components can be simultaneously determined in a single Z-scan
measurement.
1.25 1.4 1.8
Experiment
1.2 1.3 1.6 From best fitting
Normalized transmittance

Normalized transmittance

Normalized transmittance
1.15 From fixed B fitting
1.2
1.4
1.1
1.1
1.05 1.2
1
1 1
0.9
0.95
0.8
0.8
0.9

0.7 0.6
0.85

0.8 0.4
-8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8
Sample position x Sample position x Sample position x

(a) (b) (c)


1.6 1.8 1.8

1.6 1.6
1.4
Normalized transmittance

Normalized transmittance

Normalized transmittance

1.4 1.4
1.2
1.2 1.2

1 1 1

0.8 0.8 0.8

0.6 0.6
0.6
0.4 0.4
0.4
0.2 0.2

0.2 0 0
-8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8
Sample position x Sample position x Sample position x

(d) (e) (f)


Fig. 3. Normalized transmittance curves for different peak on-axis radiation intensities: (a)
4.26, (b) 7.93, (c) 12.13, (d) 17.59, (e) 20.02, (f) 24.83 GW/cm2. The polarizer angle ϕ1 is 78
deg. Owing to the blue line covers the red dashed line we can not watch red lines in (a) and (b).

Figure 3 gives the normalized transmittance curves for different peak on-axis irradiance
intensities (4.26, 7.93, 12.13, 17.59, 20.02, 24.83 GW/cm2). The angle ϕ1 was set 78 deg for
these experiments, and the linear transmittance S is 15%. It is seen that the anti-symmetry of
the peak and valley has been destroyed by large intensities. Blue lines are the fits using Eq.
(20) (terms are taken to m+n≤ 30 in order to eliminate the errors from missing terms [22])
when the B value is identical for all the intensities and the ratio B to A is 6. B=1.372×10-11 esu

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6404
chosen for this fitting was from best fit result of Fig. 3a. The B value is set to be identical for
all the intensities because it is usually assumed to be independent of intensity in the χ xyyx (3)

measurements [3].
Note that the scattering and nonlinear absorption are ignored in the fitting. The theoretical
simulations agree well with experimental data for two smaller intensity cases (i.e, Fig. 3(a),
Fig. 3(b)), but the fits are not appropriate for the latter four intensities. And the larger the
intensity is, the more evident the discrepancy between fitting and experimental results is.
These discrepancies are more obvious from the comparison of the peak-valley transmittance
differences (∆Tp-v) as shown in Fig. 4(a). The experimental data do not match the fitting
results for higher intensities. It can be seen that the peak-valley difference for fixed B is not
proportional to the light intensity but is reaching its saturation value. According to Eq. (27),
the effective nonlinear phase shift at the focus is proportional to the intensity, so the ∆Tp-v
should also be proportional to the intensity as the green dash line shown, which has been
widely used as a criterion to judge the nonlinear type [18]. The appearance of nonlinear
relationship between ∆Tp-v and intensity means that the first-order approximation cannot be
used for large nonlinear phase shift case. For large nonlinear phase shift case, although
∆ T p − v ∼ I 0 may present a saturation, the nonlinear phase shift (so does the RIC) still increases
with intensity. In order not to wrongly regard large nonlinear phase shift case as saturated
nonlinear effect, the proportional relationship between ∆Tp-v and intensity under first-order
approximation could not simply treated as criteria to determine whether it is a Kerr nonlinear
effect or not. In this case, more terms in Eqs. (20) and (26) should be taken to analyze the
experimental data and get the nonlinear phase shifts, and then judge the nonlinear type of the
medium.
-11
x 10
2.5
Experiment 1.5 From best fit
From best fit From fixed B fit
2 From fixed B fit 1.4
First-order approximation
B Value (esu)

1.3
1.5
∆Tp-v

1.2

1 1.1

1
0.5
0.9

0
0 5 10 15 20 25 0 5 10 15 20 25

I (GW/cm2) I (GW/cm2)
0 0
(a) (b)
Fig. 4. Peak-to-valley transmittance difference (a) and B value used in above fits (b) as a
function of peak on-axis radiation intensity.

From the last four fits shown as blue lines in Fig. 3, it is seen that fitting results which did
not take into account scattering and nonlinear absorption have lower valley and higher peak
than experimental curves, so the discrepancy could not mainly come from scattering or
nonlinear absorption. If scattering played an important role in higher intensity Z-scan
experiments, there would be less energy into the detector. Both the valley and the peak would
become lower than the case without scattering effect, and the valleys of experimental curves
should be deeper than the valleys of fitting curves. But, the valleys of experimental curves are
shallower than the valleys of our fitting results shown in Fig. 3. Therefore, the contribution of
scattering to the discrepancy between fixed B fitting and experimental results can be ignored.
Moreover, during our experiments, no obvious scattering was observed. According to the
literatures [7,25,27], the nonlinear absorption is negligible for intensity less than 100 GW/cm2.
Since the intensities (<25 GW/cm2) used in our experiments are much smaller than the critical

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6405
intensity, the discrepancy between fixed B fitting and experimental results should not mainly
arise from nonlinear absorption. No nonlinearity for the quartz cell was observed in our
experiments. In addition, the self-defocusing effects of acoustic generation and heat
accumulation can also be excluded, because the repetition rate of picosecond pulses in our
experiments was too low to bring this type self-defocusing effects [27].
One origin of the discrepancy between fixed B fitting and experimental results may be the
temperature changes because of absorption. According to references [19,24], the B value is
inverse to temperature. The temperature was higher for larger radiation intensity, so the B
value should be smaller for larger radiation intensity. In order to check whether the
discrepancy could come from the decreasing of B value, the fits with different B were carried
out. In the new fitting (terms in Eq. (20) is taken until m+n=30), the B values are altered to
get the best fit for different intensities. The criterion of best fit is that the peak-valley
transmittance difference from fit is nearly identical to the experimental ∆Tp-v, when the ratio
of B to A was fixed at the theoretical value 6. The new fitting curves are shown as red dashed
lines in Fig. 3, and the ∆Tp-v and B values obtained from the new fits are shown in Fig. 4.
Comparing the two fitting results in Fig. 3, we find that the new fits are much better than the
former ones (fixed B fits), and the B value varies with intensity. The B value used in new fits
decreases with the intensity as shown in Fig.4b, which is consistent with theoretical prediction
[19,24]. Note the analysis of the B value is just qualitative, from which we can conclude that
the B value decreases with intensity. If one wants to quantitatively determine the relationship
between B value and intensity, the temperature change of the liquid should be studied fully.
If the discrepancy is completely induced by the decreasing of B value, the temperature of
liquid at 24.83 GW/cm2 should be larger than 400 K [19], which is much higher than the
boiling point 319 K of CS2 [28]. Thus, it is probable that there are other processes
contributing to the discrepancy. Another origin for the discrepancy may be higher-order
nonlinearities. If CS2 has a negative fifth-order nonlinear refraction, the valley will be less
deep and the peak be lower than pure third-order nonlinearity [29,30], which are consistent
with the discrepancies for higher intensity cases in Fig. 3. It is possible that there is an effect
of fifth-order nonlinearity refraction of Z-scan experiments at high intensity cases. Additional
source of the discrepancies may be related to simultaneous multi-photon processes and
saturation absorption processes, multi-photon processes make the peak lower and saturation
absorption processes make the valley shallower. In summary, from the two fitting results, It
can be confirmed that the Eqs. (20) and (26) can also be available for a large nonlinear phase
shift.
4. Conclusion
We have presented the theoretical and experimental analyses of Z-scan measurements by
arbitrary polarized lights in isotropic medium. The normalized transmittance formulas for
arbitrary aperture radius were obtained. The CA case of Z-scan was used to analyze the
linearly, circularly, elliptically polarized lights Z-scan data in CS2 medium. From the analysis,
we found that the elliptically polarized light Z-scan could be used to simultaneously measure
the two third-order nonlinear susceptibility components χxyyx (3) and (3) . Furthermore, the ratio of
χ xxyy
B to A was determined to be about 6, which agrees very well with theoretical value for
molecular orientation. The normalized transmittance formulas of elliptically polarized light Z-
scan were also examined for large nonlinear phase shifts.
Acknowledgments
This research was supported by the Natural Science Foundation of China (grant 60708020),
Chinese National Key Basic Research Special Fund (grant 2006CB921703), the Research
Fund for the Doctoral Program of Higher Education of China (No. 20070055045), and 111
Project (B07013).

#107050 - $15.00 USD Received 4 Feb 2009; revised 15 Mar 2009; accepted 15 Mar 2009; published 2 Apr 2009
(C) 2009 OSA 13 April 2009 / Vol. 17, No. 8 / OPTICS EXPRESS 6406

You might also like