You are on page 1of 7

Fluctuations near a phase transition in liquid crystals

Thomas Moses, Jason Reeves and Paulo Pirondi

Citation: American Journal of Physics 75, 220 (2007); doi: 10.1119/1.2410017


View online: https://doi.org/10.1119/1.2410017
View Table of Contents: https://aapt.scitation.org/toc/ajp/75/3
Published by the American Association of Physics Teachers

ARTICLES YOU MAY BE INTERESTED IN

The Fréedericksz transition in liquid crystals: An undergraduate experiment for the advanced laboratory
American Journal of Physics 66, 49 (1998); https://doi.org/10.1119/1.18807

Superconductivity, broken gauge symmetry, and the Higgs mechanism


American Journal of Physics 87, 436 (2019); https://doi.org/10.1119/1.5093291

An optical n-body gravitational lens analogy


American Journal of Physics 89, 11 (2021); https://doi.org/10.1119/10.0002117

Scaling laws at the critical point


American Journal of Physics 74, 441 (2006); https://doi.org/10.1119/1.2173273

Neutron stars for undergraduates


American Journal of Physics 72, 892 (2004); https://doi.org/10.1119/1.1703544

Resource letter CPPPT-1: Critical point phenomena and phase transitions


American Journal of Physics 69, 255 (2001); https://doi.org/10.1119/1.1333102
Fluctuations near a phase transition in liquid crystals
Thomas Moses,a兲 Jason Reeves, and Paulo Pirondi
Department of Physics, Knox College, Galesburg, Illinois 61401
共Received 26 June 2006; accepted 17 November 2006兲
The increase in molecular orientational fluctuations as a phase transition is approached is an
interesting and dramatic phenomenon, but associated undergraduate experiments are rare. We
present an advanced undergraduate experiment on light scattering from the molecular orientational
fluctuations in liquid crystals. In the high-temperature, disordered phase of a liquid crystal, small
clusters of parallel-oriented molecules form spontaneously and scatter light. The size of the
fluctuations and the consequent light scattering increase dramatically as the temperature is reduced
approaching the weakly first-order phase transition to the ordered liquid crystal phase. We discuss
a simplified theory that captures the essential physics. © 2007 American Association of Physics Teachers.
关DOI: 10.1119/1.2410017兴

I. INTRODUCTION strong increase in light scattering due to density fluctuations


in the high-temperature gas phase near the critical point. In a
The statistical mechanics of phase transitions is an excit- liquid crystal it is molecular orientational fluctuations rather
ing area of contemporary research, but is not well repre- than density fluctuations that cause the light scattering, but
sented in typical undergraduate laboratory curricula. We de- the essential physics is similar.
scribe an undergraduate experiment in which the fluctuations Several experiments involving light scattering4–6 and two
near a phase transition in a liquid crystal are important. The papers involving critical phenomena in liquid crystals have
experimental apparatus is inexpensive and the experiment is appeared recently.7,8 The advantages of the light scattering
not technically difficult or lengthy to perform. The choice of experiment reported here is the simplicity of the experimen-
liquid crystals as a system to investigate pretransitional fluc- tal apparatus and the convenience of liquid crystal samples.
tuations is convenient because of their large, easily detect- The pretransitional fluctuations in the liquid crystal are a
able fluctuations and near room temperature phase transition. straightforward but interesting application of the statistical
A liquid crystal is composed of rod-like molecules. At mechanics of fluctuations, making the experiment ideal for
high temperatures, the molecular rods point in random direc- an advanced laboratory experiment.
tions and the liquid is isotropic. As the temperature is low-
ered there occurs a transition to the nematic phase, where the II. FLUCTUATIONS IN LIQUID CRYSTALS
rods align approximately parallel to each other 共see Fig. 1兲. A. Order parameter and Landau free energy
The phase transition to the orientationally ordered nematic
phase results from anisotropic interactions between neigh- We present a brief, simplified analysis of the temperature-
boring molecules. Our experiment measures the intensity of dependence of fluctuations in a liquid crystal. Our approach
light scattered from the high-temperature isotropic phase as follows Collings and Hird,9 who provide more details. A
the temperature approaches the nematic phase. In the isotro- more rigorous treatment of the fluctuations in a liquid crystal
pic phase, random fluctuations of the molecular orientations must take into account their tensor character. An especially
cause the temporary formation of microscopic domains of clear treatment is given in the text by Vertogen and de Jeu.3
parallel-oriented molecules that scatter light, because the Phase transitions, including the orientational transition in
nematic-like domains are birefringent and optically distinct liquid crystals, can be described quantitatively by an order
from the average isotropic background. As the temperature parameter, which is zero in the disordered 共isotropic兲 phase
approaches the isotropic-nematic transition temperature TNI, and nonzero in the ordered 共nematic兲 phase. We define the
the number and size of the nematic-like domains grow. This order parameter S as
growth is due to the fact that near the phase transition, the S = 具 21 共3 cos2 共␪兲 − 1兲典 , 共1兲
free energies of isotropic and nematic ordering become com-
parable, so the free energy cost of a nematic-like fluctuation where ␪ is the angle between an axis of a rod-like molecule
is reduced. Consequently, the probability of a nematic-like and the local average direction of the rods 共called the direc-
fluctuation is increased near the phase transition.1 tor兲; the brackets indicate an average over a region contain-
The increase in nematic-like fluctuations near the phase ing many molecules. In all known nematic liquid crystals,
transition affects various properties of the liquid crystal, such ␪ = 0 is equivalent to ␪ = ␲; the nematic phase has overall
as its heat capacity, viscosity, and its ability to scatter light.2,3 “head-tail” symmetry, even though the individual molecules
To understand the effect on light scattering, we note that the usually do not. It can be readily checked that S = 0 in the
size of the nematic-like domains typically increases from a isotropic phase and S = 1 in a perfectly aligned nematic
few molecular lengths to a few tens of molecular lengths as phase.10
the phase transition is approached. As the size of the domains We next consider a simple model for a second-order phase
increases 共though remaining much smaller than the optical transition1 due to Landau, modified to describe the isotropic-
wavelength兲, the scattering efficiency increases and we ob- nematic transition in liquid crystals by de Gennes.2 In the
serve a strong temperature dependence of the light scattering Landau–de Gennes picture, the Gibbs free energy g per unit
intensity. This increased scattering is analogous to critical volume in the isotropic phase is represented as a power series
opalescence at the liquid-gas phase transition, which is a in the order parameter S and its spatial derivatives:

220 Am. J. Phys. 75 共3兲, March 2007 http://aapt.org/ajp © 2007 American Association of Physics Teachers 220
where S共x兲 represents the varying order parameter in the
sample, is given by

P共S兲 ⬀ exp 共− 共Gtot − Go兲/kBT兲, 共6兲

where Gtot is given by Eq. 共4兲 and Go = goV, with V = AoL the
sample volume.
We proceed by expanding the order parameter S共x兲 in a
Fourier series,

S共x兲 = 兺 S共q兲 exp 共iqx兲, 共7兲


q

Fig. 1. Isotropic 共a兲 and nematic 共b兲 phases. The rods represent molecules; where the sum over q runs over the values q = 2␲n / L for n
the shortened rods represent molecules pointing into or out of the page. = 0 , ± 1 , ± 2 , . . . and S共q兲 represents the Fourier coefficients.

冉 冊
Although S共x兲 is a real function, the Fourier coefficients S共q兲
2
dS are in general complex and are given by
g = go + AS2 + B . 共2兲


dx L
1
Here go, A, and B are material-dependent constants, indepen- S共q兲 = dxS共x兲 exp 共− iqx兲. 共8兲
L 0
dent of S, but possibly functions of temperature and pressure.
Because S is small in the isotropic phase 共zero, if not for From Eq. 共8兲 we see that S共−q兲 = S共q兲*, where * denotes the
fluctuations兲, the series expansion of g is truncated after the complex conjugate. If we substitute the Fourier series of Eq.
first few terms. There is no term linear in S because the free 共7兲 into Eq. 共2兲 and calculate the free energy Gtot of Eq. 共4兲,
energy g共S兲 is a minimum at S = 0 in the isotropic phase. We we find that the S2 term becomes VA兺q 兩 S共q兲兩2, and the
can approximate A by its Taylor expansion as 共dS / dx兲2 term can be written VB兺qq2 兩 S共q兲兩2. We substitute
A = a共T − T*兲, 共3兲 these two terms into Eq. 共6兲 and find
*
where a and T are material-dependent constants and pres-
sure dependence is neglected because the pressure remains
constant in the experiment. The material-dependent constant
P ⬀ exp − 冋 V
兺 兩S共q兲兩2共A + Bq2兲
k BT q 册 共9兲

T* has dimensions of temperature and is usually of the order


for the probability of a fluctuation.
of 1 K below the phase transition temperature TNI. The last
We wish to find an expression for 具兩S共q兲兩2典, the mean
term in Eq. 共2兲 expresses the energy cost of a spatial varia-
tion in the order parameter S共x兲. This term might be thought square amplitude of the fluctuations of the order parameter
of as expressing the energy cost of a deformation in analogy component with wavenumber q. If we evaluate the desired
to the potential energy kx2 / 2 of a compressed spring. mean value from the probability distribution in Eq. 共9兲, we
In Eq. 共2兲 and in the rest of this paper we will consider a obtain the result11
one-dimensional model and take S = S共x兲 as a function of x k BT k BT
only.9 This approach simplifies the mathematics and pre- 具兩S共q兲兩2典 = 2 = . 共10兲
serves the essential physics of the problem. The total free A + Bq a共T − T*兲 + Bq2
energy of the sample is given by an integral of the free en-
Equation 共10兲 gives the temperature-dependence of the order
ergy density:
parameter fluctuations. Note that as T approaches T*,
Gtot = 冕0
L
dxAog共S兲, 共4兲
具兩S共q兲兩2典 increases to a maximum—this pretransitional in-
crease in the order fluctuations is responsible for the ob-
served increase in scattered light intensity as the transition is
where L is the length of the sample and Ao is the cross- approached.12
sectional area of the sample. The intensity I共q兲 of light scattered with scattering
wavevector q 共q is the change in the wavevector of the scat-
B. Temperature-dependence of fluctuations tered light兲 is proportional to 具兩S共q兲兩2典. From Eq. 共10兲 we
obtain:
We next consider spontaneous fluctuations of the molecu-
lar orientation in the isotropic phase in a liquid crystal main- T
tained at constant temperature and pressure. We focus on a I共q兲 ⬀ . 共11兲
a共T − T*兲 + Bq2
small volume element ⌬V in which a fluctuation in the mo-
lecular orientations has resulted in a temporary nonzero order A brief derivation of the relation I共q兲 ⬀ 具兩S共q兲兩2典 is given in
parameter S. A standard result in statistical mechanics gives the Appendix ; more details may be found in Ref. 3. The
the probability P⌬V共S兲 of a fluctuation with order parameter plausibility of the result can be understood by noting that an
S in the volume element ⌬V 共Ref. 11兲: order parameter fluctuation involves a local ordering of mol-
P⌬V共S兲 ⬀ exp 共− 共g共S兲 − go兲⌬V/kBT兲, 共5兲 ecules, and the locally ordered domain comprises an optical
inhomogeneity in an isotropic background, causing the scat-
where g共S兲 is given by Eq. 共2兲 and kB is Boltzmann’s con- tering of light. Hence increased fluctuations in the order pa-
stant. The overall probability P共S兲 of a fluctuation S共x兲, rameter should lead to increased scattering of light.

221 Am. J. Phys., Vol. 75, No. 3, March 2007 Moses, Reeves, and Pirondi 221
Fig. 3. Molecular structure of 5 CB.

cap with epoxy because traces of liquid crystal will prevent a


secure epoxy bond; a cotton swab dipped in acetone is effec-
tive, though care is needed to avoid contaminating the liquid
Fig. 2. Schematic of the optical arrangement. L⫽polarized HeNe laser, crystal sample with acetone.
S⫽liquid crystal sample in a thermally-controlled chamber, A⫽aperture, The use of round glass tubing is convenient and inexpen-
Le⫽lens, BS⫽beam stop, B⫽barrier for rejecting stray light, PMT sive, but in some respects not ideal for a high-precision ex-
⫽photomultiplier tube detector, DMM⫽digital multimeters, and periment, because the round sample acts as a strong cylindri-
C⫽computer. cal lens of the scattered light and the less than optical quality
glass surfaces may backscatter more light than precision op-
tical surfaces would. In our experiments, these effects were
III. EXPERIMENTAL APPARATUS not significant and the data agreed well with theoretical pre-
A. Optical setup and alignment dictions. We also used precision miniature square cuvettes
with flat sides of high-optical-quality glass to avoid the lens-
A diagram of the experimental arrangement is shown in ing effect and minimize backscattering from the glass cell,
Fig. 2. A linearly polarized helium-neon laser 共2 mW兲 was and obtained essentially identical data.
used as the light source. The sample cell is a sealed glass In our experiments we used the liquid crystal
tube containing a liquid crystal and is housed in a homemade 4⬘-n-pentyl-4-cyanobiphenyl 共5CB兲 共see Fig. 3兲.13 5CB has
thermally controlled chamber. Light transmitted through the become a standard liquid crystal for experiments due to its
sample without scattering is allowed to propagate 1 m from chemical stability, low cost, thoroughly tabulated physical
the sample before being blocked to minimize light scattered properties, and conveniently accessible isotropic-nematic
back to the detector. Light scattered from the sample at 90° transition temperature 共35.0 ° C兲. Our sealed sample tubes
passes through an aperture and is collected by a lens 共10 cm have shown no degradation in 3 years.
focal length兲 and focused on a photomultiplier tube detector. It is necessary to control the temperature of the liquid
The aperture selects light scattered from the liquid crystal crystal sample during the experiment. The simplest approach
sample and rejects light scattered from the sides of the exit is to house the sample tube in a hole drilled in a copper block
orifice of the sample chamber housing or other extraneous 共chosen for its high thermal conductivity and heat capacity兲.
sources. A black cloth barrier was placed around the detector The copper block will need additional holes for the incident
to reject stray light from sources other than the liquid crystal and transmitted light beams and the scattered beam 共we
sample. The experiment was done in a dark room with care chose to measure the light scattered at 90°兲 and to accom-
taken to minimize sources of stray light. Optionally, a laser- modate a heating resistor14 and a thermistor15 for measuring
line filter mounted at the entrance aperture of the photode- the temperature 共see Fig. 4兲. The copper block is heated to
tector can be used to reject stray light at frequencies other the desired highest temperature and allowed to decrease
than the laser frequency. The detector output voltage was freely through the isotropic-nematic transition while the tem-
measured using a high quality voltmeter interfaced to a com- perature and scattered light intensity data are recorded. For
puter. Another option is to chop the laser beam and use a the 750 g copper block we used, the temperature decreased
lock-in amplifier as a detector. In combination with a laser- at a rate of ⬃1 ° C / min while a computer acquired the resis-
line filter, lock-in detection makes the experiment essentially tance data from a digital multimeter connected to the ther-
insensitive to stray light and allows the room lights to be on. mistor. An alternative approach is to actively control the
The vertical polarization direction 共perpendicular to the sample temperature with feedback; a simple and inexpensive
scattering plane兲 of the incident light was chosen primarily controller for this purpose is described in Ref. 7. Unlike
for purposes of reproducibility. The use of a polarized laser many experiments on critical phenomena, highly accurate
is convenient because polarization fluctuations in an unpolar-
ized laser may cause undesired fluctuations in scattered light
intensity.

B. Sample materials and sample cell


In our experiments the liquid crystal material was en-
closed in a small piece of circular glass tubing and housed in
a thermally controlled chamber. The tubing dimensions were
3.2 mm 共inner diameter兲 ⫻25 mm long. The tube was sealed
with homemade aluminum caps epoxied in place. After the
bottom end of the tube is sealed, the tube can be conve-
niently filled with liquid crystal using a syringe. A few mil-
limeters of air space should be left at the top of the tube to
allow for thermal expansion of the liquid crystal as the Fig. 4. Top view of the sample cell and thermally controlled housing.
sample temperature is changed. If necessary, the top surface S⫽sample tube, H⫽heater resistor, and T⫽thermistor. Note the access ports
of the tube should be cleaned carefully before sealing the top for the incident beam, the unscattered beam, and the light scattered at 90°.

222 Am. J. Phys., Vol. 75, No. 3, March 2007 Moses, Reeves, and Pirondi 222
Fig. 6. A plot of T / I versus T shows the linear dependence predicted by the
Landau-de Gennes theory, Eq. 共11兲. The values of the free energy parameter
T* = 1.6± 0.2 ° C was determined from the slope and intercept values of the
Fig. 5. Scattered light intensity versus temperature. Note the strong increase plot.
in light scattered by the fluctuations as the isotropic liquid crystal is cooled
toward the isotropic-nematic transition.
of T / I versus ␧ will be linear, with T* given by
temperature control is not required for this experiment be- T* = TNI − b/m, 共13兲
cause the light scattering intensity is easily measured up to where m and b are fit parameters 共slope and intercept兲. The
10 ° C from the phase transition; temperature precision as data and linear fit are shown in Fig. 6. The good agreement is
crude as ±0.1 ° C would suffice. Regardless of the method consistent with the theory of order parameter fluctuations in
used to vary the sample temperature, it is important to mea- liquid crystals we have described. For the best fit values m
sure all temperatures relative to the observed phase transition = 6.67± 3% and b = 10.9± 20%, Eq. 共13兲 gives T*
temperature. This procedure will 共mostly兲 compensate for = 1.6± 0.2 ° C in reasonable agreement with the published
errors such as a possible offset in the thermometer calibra- value of 1.4 ° C.16 Some investigators have reported a small
tion or a temperature differential between the measuring but significant deviation of the experimental results from the
thermometer and the liquid crystal sample. Landau–de Gennes theory at temperatures near TNI 共within a
few tenths of a degree, depending on the liquid crystal ma-
IV. EXPERIMENTAL RESULTS terial兲 and have explained the discrepancy by the known in-
adequacy of mean-field theory to account fully for the effects
Experimental data were collected by measuring the inten-
of fluctuations in the vicinity of the phase transition.16 This
sity of scattered light and the sample temperature as the iso-
inadequacy may explain the small deviation of the data
tropic liquid crystal sample temperature approached the
points in Fig. 6 from the theoretical line at temperatures
isotropic-nematic phase transition from above. A sample of
the raw data 共light intensity versus temperature兲 is shown in within 0.5 ° C of TNI.
Fig. 5. The sharp pretransitional increase in the light scatter-
ing due to nematic-like fluctuations can be clearly seen. The V. SUMMARY
transition temperature TNI was clearly identifiable as the in-
Our experiment on light scattering from liquid crystals
tensity of scattered light increased by more than an order of
introduces students to the physics of fluctuations and phase
magnitude as the sample changed from isotropic to a multi-
transitions. Unlike many experiments on critical phenomena,
domain nematic.
sample preparation is easy and expenses are minimal. The
We can replot the data to make a convenient comparison
typical results presented here show the dramatic increase in
with the theoretical prediction of Eq. 共11兲. The term Bq2 in
fluctuations near the phase transition. The experimental re-
the denominator of Eq. 共11兲 is negligible for our experimen-
sults are well described by the simplified Landau-type theory
tal conditions. To determine its magnitude, we evaluate
of a weakly first-order phase transition, which predicts a
Bq2 / a共T − T*兲 at the minimum attainable isotropic-phase
strong pre-transitional increase of the scattered light intensity
temperature T = TNI = 35.0 ° C. If we use q = ki − k f , q as the temperature of the disordered, isotropic phase is
= 2ki sin 共␪ / 2兲 for elastic scattering 共ki = k f 兲, and ␪ = 90° and cooled toward the transition to the nematic phase.
ki = 2␲ / ␭ with ␭ = 633 nm for the HeNe laser, we find q
= 14.0 ␮m−1. The published values for the material constants
ACKNOWLEDGMENTS
of 5CB are a = 6.0⫻ 104 JK−1 m−3 and
16,17
−12
B = 4.6⫻ 10 J / m, soBq2 / a共T − T*兲 = 0.01Ⰶ 1, justify- Support from the Ford Foundation and the McNair Foun-
ing the neglect of the Bq2 term. dation are gratefully acknowledged.
We can then rewrite Eq. 共11兲 as
T/I = k共T − T*兲 = k共␧ − ␧*兲, 共12兲 APPENDIX: ORDER FLUCTUATIONS AND LIGHT
SCATTERING
where k is a constant of proportionality, ␧ = T − TNI, and ␧*
= T* − TNI. We have expressed the temperatures on the right- We present a brief derivation of the relation of the mean
hand side relative to the observed phase transition tempera- square fluctuations of the order parameter to the light scat-
ture TNI, which provides a convenient correction for possible tering. We define ␦␧, the anisotropic part of the dielectric
offsets in the thermometry. Equation 共12兲 predicts that a plot tensor of the liquid crystal, as

223 Am. J. Phys., Vol. 75, No. 3, March 2007 Moses, Reeves, and Pirondi 223
␦␧ = ␧ − ␧iI, 共A1兲 P = n具␣典E = ␧o␹eE, 共A9兲
where ␧ is the dielectric tensor, ␧i is the dielectric constant in where n is the number of molecules per unit volume, E is the
the isotropic phase, and I is the identity matrix. We note that electric field, and ␹e is the electric susceptibility tensor. As
␦␧ vanishes in the isotropic phase in the absence of fluctua- usual, the dielectric tensor ␧ is given by
tions and a nonzero ␦␧ describes a nematic-like fluctuation. D = ␧ o␧ E with ␧ = I + ␹e . 共A10兲
For a nematic-like region in a uniaxially birefringent liquid
crystal, the dielectric tensor is given by If we combine Eqs. 共A8兲–共A10兲, we obtain

冢 冣
␧⬜ n n
⌬␧ = ␧zz − ␧xx = 共具␣zz典 − 具␣xx典兲 = 共␣储 − ␣⬜兲
␧= ␧⬜ , 共A2兲 ␧o ␧o
␧储 ⫻具 21 共3 cos2 ␪ − 1兲典 ⬀ S, 共A11兲
where the local director 共direction of the rod-like molecules兲 which justifies Eq. 共A5兲.
is along the z-axis and ␧⬜ and ␧储 are the components of the There is a subtle problem in our discussion: the electric
dielectric tensor perpendicular and parallel to the director. field in the middle equality of Eq. 共A9兲 is the microscopic,
The dielectric constant in the isotropic phase can be written internal field seen by one molecule 共the sum of the external
as ␧i = 31 共2␧⬜ + ␧储兲, so by combining Eqs. 共A1兲 and 共A2兲 we applied field and the field produced by molecular dipole mo-
obtain ments other than the molecule under observation兲, whereas
the field in the rightmost equality of Eq. 共A9兲 and in Eq.

冢 冣
− 1
3
共A10兲 is the average macroscopic field in the medium.
A detailed calculation of the internal field in a system of
␦␧ = ⌬␧ − 1
3 , 共A3兲 interacting dipoles is difficult and approximations must be
2
3
made. The effect of the most commonly used approximate
treatments is only to change the proportionality constant in
where ⌬␧ = ␧储 − ␧⬜ is the dielectric anisotropy. In an arbitrary Eq. 共A11兲, and thus the proportionality relation remains
coordinate system, we can find the new form of ␦␧ by ap- valid.19
plying appropriate rotation matrices, but it is clear from Eq. The intensity of light scattering can be related to fluctua-
共A3兲 that each component of ␦␧ remains proportional to ⌬␧: tions of the dielectric tensor. Specifically, the Fourier com-
ponent I共q兲 of the intensity of light scattered with wavevec-
␦␧ ⬀ ⌬␧. 共A4兲
tor q is given by
We will next show that ⌬␧ is proportional to the scalar
Iif 共q兲 ⬀ 具兩␦␧if 共q兲兩2典. 共A12兲
order parameter S defined in Eq. 共1兲:
Here i and f represent the incident and final polarization
⌬␧ ⬀ S = 具 21 共3 cos2共␪兲 − 1兲典 . 共A5兲 directions and q = ki − k f is the scattering wavevector, with ki
Equation 共A5兲 is plausible because ⌬␧ = 0 and S = 0 in the and k f representing the wavevectors of the incident and scat-
isotropic phase in the absence of fluctuations, and ⌬␧ ⫽ 0 and tered radiation. An accessible derivation of Eq. 共A12兲 is
S⫽ 0 in an orientationally ordered, birefringent nematic given in Ref. 20. Because the proportionality in Eq. 共A12兲
phase 共or within a nematic-like fluctuation in the isotropic holds regardless of the polarization states, we may drop the
phase兲. We define the molecular polarizability tensor ␣ for a subscripts. Finally, we combine Eqs. 共A4兲, 共A5兲, and 共A12兲
single molecule, approximated as a cylindrically symmetric and find
rod oriented along the z-axis: I共q兲 ⬀ 具兩S共q兲兩2典, 共A13兲

冢 冣
␣⬜ justifying the result in Eq. 共11兲.
␣= ␣⬜ . 共A6兲
␣储
a兲
Electronic mail: tmoses@knox.edu
1
The isotropic-nematic transition in liquid crystals is a weak first-order
We can find the average molecular polarizability tensor 具␣典 transition, with an very small latent heat 关1.42 J / g for 5CB 共Ref. 16兲,
compared to, for instance, 333 J / g for water兴, which is why pretransi-
for a collection of molecules by expressing Eq. 共A6兲 in a tional fluctuations are so prominent.
coordinate system rotated by Euler angles ␾, ␪, ␺ 共about the 2
P. G. de Gennes, “Short range order effects in the isotropic phase of
original z-axis, the new x-axis, and the final z-axis, in suc- nematics and cholesterics,” Mol. Cryst. Liq. Cryst. 12, 193–214 共1971兲.
cession兲 and then averaging over angles:18 See also P. G. de Gennes and J. Prost, The Physics of Liquid Crystals, 2nd
ed. 共Clarendon, Oxford, 1993兲.
具␣典 = 具R␺R␪R␾␣R␾−1R␪−1R␺−1典, 共A7兲 3
G. Vertogen and W. H. de Jeu, Thermotropic Liquid Crystals, Fundamen-
tals 共Springer-Verlag, New York, 1988兲.
4
where R represents a rotation matrix and the brackets indi- Noel A. Clark, Joseph H. Lunacek, and George B. Benedek, “A study of
cate the average over angles ␾, ␪, ␺. The result is Brownian motion using light scattering,” Am. J. Phys. 38 共5兲, 575–585
共1970兲.
具␣xx典 = 具␣yy典 = 具 21 共1 + cos2 ␪兲典␣⬜ + 具 21 sin2 ␪典␣储 , 共A8a兲
5
R. T. Schumacher, “Brownian motion by light scattering revisited,” Am.
J. Phys. 54共2兲, 137–141 共1986兲.
6
W. I. Goldburg, “Dynamic light scattering,” Am. J. Phys. 67共12兲, 1152–
具␣zz典 = 具sin2 ␪典␣⬜ + 具cos2 ␪典␣储 共A8b兲 1160 共1999兲.
7
Thomas Moses, Brian Durall, and Gregory Frankowiak, “Magnetic bire-
with the off-diagonal elements vanishing. The polarization P fringence in a liquid crystal: An experiment for the advanced undergradu-
of the medium is given by ate laboratory,” Am. J. Phys. 68共3兲, 248–253 共2000兲.

224 Am. J. Phys., Vol. 75, No. 3, March 2007 Moses, Reeves, and Pirondi 224
8 15
A. J. Nicastro, “Experiment in critical phenomena at phase transitions in Thermistors are available from Yellow Springs Instruments 共part no. YSI-
liquid crystals,” Am. J. Phys. 52共10兲, 912–915 共1984兲. 44106兲, distributed by Newark Electronics.
9 16
Peter J. Collings and Michael Hird, Introduction to Liquid Crystals: H. J. Coles and C. Strazielle, “The order-disorder phase transition in
Chemistry and Physics 共Taylor and Francis, Bristol, PA, 1997兲. liquid crystals as a function of molecular structure. I. The alkyl cyanobi-
10
We can calculate explicitly the average in Eq. 共1兲 by weighting each phenyls,” Mol. Cryst. Liq. Cryst. 55, 237–250 共1979兲.
orientation angle ␪ by its probability P共␪兲. In the isotropic phase, P共␪兲 17
W. Chen, L. J. Martinez-Miranda, H. Hsiung, and Y. R. Shen, “Orienta-
= 1 / ␲, because all values of ␪ between 0 and ␲ are equally likely. Hence tional wetting behavior of a liquid-crystal homologous series,” Phys. Rev.
we calculate the average value of S in the isotropic phase as S Lett. 62共16兲, 1860–1863 共1989兲.
= 兰␲0 d␪ sin 共␪兲 2 共3 cos2 ␪ − 1兲共1 / ␲兲 = 0. We can similarly show that S = 1
1
18
Introduction to Liquid Crystals, edited by E. B. Priestley, Peter J. Woj-
for a perfectly aligned nematic phase in which P共␪兲 = ␦共cos 共␪兲 − 1兲, where
towicz, and Ping Sheng 共Plenum, New York, 1974兲. In the average over
␦ is the Dirac delta function.
11 angles, uniaxial symmetry in the liquid crystal is assumed, so all values
L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I, 3rd ed.
of ␾ and ␺ are equally probable.
共Pergamon, New York, 1980兲. 19
12
The sample temperature never actually reaches T* because the isotropic- Reference 3 provides a full discussion. The simplest approximation is to
nematic phase transition first occurs at TNI, about 1 ° C above T*. So neglect the anisotropy of the system, treating the internal field as that at
具兩S共q兲兩2典 does not diverge to infinity, but exhibits a strong temperature the center of a spherical cavity in a uniform dielectric, which leads to Eq.
dependence. 共A5兲. The validity of Eq. 共A5兲 is well supported by experimental evi-
13
Small quantities of 5CB 共catalog #32, 851-0兲 are available from Sigma- dence.
20
Aldrich Co. 具www.sigmaaldrich.com典. Bruce J. Berne and Robert Pecora, Dynamic Liquid Scattering with Ap-
14
Cartridge-type resistance heaters 共part no. #E1J40兲 are available from plications to Chemistry, Biology, and Physics 共Krieger, Malabar, FL,
Watlow Inc., St. Louis, MO. 1990兲.

225 Am. J. Phys., Vol. 75, No. 3, March 2007 Moses, Reeves, and Pirondi 225

You might also like