You are on page 1of 6

ARTICLES

PUBLISHED ONLINE: 9 MARCH 2014 | DOI: 10.1038/NMAT3888

Electro-optical switching of graphene oxide liquid


crystals with an extremely large Kerr coefficient
Tian-Zi Shen†, Seung-Ho Hong† and Jang-Kun Song*

The sensitive response of the nematic graphene oxide (GO) phase to external stimuli makes this phase attractive for extending
the applicability of GO and reduced GO to solution processes and electro-optic devices. However, contrary to expectations,
the alignment of nematic GO has been difficult to control through the application of electric fields or surface treatments. Here,
we show that when interflake interactions are sufficiently weak, both the degree of microscopic ordering and the direction of
macroscopic alignment of GO liquid crystals (LCs) can be readily controlled by applying low electric fields. We also show that
the large polarizability anisotropy of GO and Onsager excluded-volume effect cooperatively give rise to Kerr coefficients that
are about three orders of magnitude larger than the maximum value obtained so far in molecular LCs. The extremely large
Kerr coefficient allowed us to fabricate electro-optic devices with macroscopic electrodes, as well as well-aligned, defect-free
GO over wide areas.

G
O-LCs with controllable alignments may provide a new nanotubes to form a nematic phase with a large anisotropy and
way to make use of the many advantages of carbon-based good fluidity.
materials1–5 . In particular, chemical-reduction techniques
can combine the attractive properties of graphene—which has LC phase sequence of GO dispersions
outstanding electrical, thermal and mechanical properties— By analysing the aqueous GO flakes used in our experiments
with the merits of self-assembled GO-LCs (refs 6,7). The wide (Methods and Supplementary Fig. 1), we verified that they were
applicability and commercial success of molecular LCs are largely mostly single-layered and had a mean size of 3.2 µm (refs 2,7,18,
attributed to the easy, fine and dynamic control of their alignment 20,21), which means that the average aspect ratio was about 3,200.
by application of electric fields, which is regarded as an intrinsic Using Onsager’s theory (Supplementary Fig. 2), we determined the
advantage of the fluidic liquid nature of LCs (refs 8,9). However, it theoretical concentrations for phase transitions from the isotropic
was reported that the application of electric fields does not induce phase to the biphase (CIB T = 0.04 vol%), and from the biphase to the
GO-LC switching, but instead causes electrophoretic drift and nematic phase (CBN T = 0.17 vol%). By observing the macroscopic
reduction of the GO, along with electrolysis of water10 . Although birefringence pattern (Supplementary Fig. 3), we also determined
a couple of methods to align GO-LCs using the capillary effect that the experimental values of CIB E and CBN E were approximately
and magnetic fields have been suggested10,11 , these methods have 0.08 and 0.2 vol%, respectively, which were only slightly larger than
clear limitations that prevent the fine and arbitrary control of the theoretical values. This difference may be due to the excessive
GO alignment. The insensitivity to electric fields seems to arise flexibility of the GO flakes22 . By measuring the volume fraction of
from viscous rheology similar to polymeric liquid crystals, and is the high-concentration nematic phase relative to that of the low-
considered as an intrinsic limitation1,10 . Meanwhile, the alignment concentration isotropic phase in a bottle stored for several weeks,
and electrical switching of carbon nanotubes and graphene flakes the heterogeneous volume-fraction diagram was also determined
were demonstrated by using dispersions in a calamitic nematic LC (Fig. 1b)10 . It was found that the heterogeneous biphase (BPH )
host and exploiting its dynamics12–14 . appeared in a wider concentration range.
Onsager’s theory, based on excluded volume, predicts the
transition from an isotropic to a nematic phase passing through Flow- and field-induced birefringence
a biphase (coexistence of isotropic and nematic phases) as the Interestingly, flow-induced optical birefringence was observed
concentration increases in colloidal dispersions of rod-like carbon in the GO dispersions below CIB E , indicating the transition
nanotubes or plate-like GOs. The location of this transition will from the isotropic to an ordered state. As shown in the inset
depend sensitively on the aspect ratio (length/diameter for rods or image in Fig. 1b and Supplementary Fig. 4, the aqueous GO
diameter/thickness for plates)15,16 . Whereas a typical aspect ratio dispersions in a bottle and in a Petri dish looked isotropic initially,
for carbon nanotube LCs is usually less than 50 because of their but bright birefringence patterns were induced under crossed
rather thick diameters and easy entanglement of excessively long polarizers after the bottle was shaken or a magnet bar was rotated
carbon nanotubes1,17 , the aspect ratio of GO-LCs with monoatomic in the Petri dish. A clearer visualization of the flow-induced
thickness can be up to the order of 10,000 (ref. 11). As a result, the birefringence is provided in Supplementary Video 1. The flow-
biphase is theoretically predicted to appear at an order of 1–10 vol% induced birefringence disappeared after a while when the dispersion
for carbon nanotube LCs (ref. 17) and around 0.01 to 0.1 vol% for stopped flowing, and the persistence time of the flow-induced
GO-LCs (refs 10,18). The viscosities of these biphases were reported birefringence in a Petri dish test was recorded as a function of
to be about 500 Pa s for carbon nanotube LCs (ref. 19) and less than concentration. The persistence time increased sharply from the
1 Pa s for GO-LCs (ref. 18). Thus, GO is a better material than carbon concentration of 0.04 vol%, indicating the existence of nematic

School of Information and Communication Engineering, Sungkyunkwan University, Jangan-Gu, Suwon, Gyeonggi-do 440-746, Korea. †These authors
contributed equally to this work. *e-mail: jk.song@skku.edu

394 NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials

© 2014 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3888 ARTICLES
a
Isotropic BPU Nematic
Uniform
phase
sequence

100 µm

Macroscopic CIB_E CBN_E


transition concent 0.08 0.2

Theoretical CIB_T CBN_T


transition concent 0.04 0.17

b
8 100

80
FIB persistence time (min)

N volume fraction (%)


6

60
BPH N
4
40

2
20

0 0
0.0125 0.0250 0.0500 0.1000 0.2000 0.4000 0.8000
Concentration (vol%)

Figure 1 | Phase sequence, nematic volume-fraction diagram and flow-induced birefringence. a, The uniform sequence of phases was determined
experimentally (corresponding phase transition concentrations: CIB E , CBN E ) and theoretically (CIB T , CBN T ; Supplementary Figs 2 and 3 for details). The
polarized optical microscopy images exhibit different birefringence patterns (brushes) depending on the phase. BPU and BPH represent the biphases with
uniform concentration throughout the vessel and with heterogeneous concentrations in a stored bottle, respectively. b, The volume fraction of the nematic
phase as a function of GO concentration (red squares) was obtained from bottles stored for two weeks. The persistence time of the field-induced
birefringence (FIB, inset image) after the shaking stopped is also plotted (blue circles). It rose sharply as the concentration increased above 0.04 vol%.

ordering interactions even in the macroscopically isotropic phase increased as the concentration was increased to 0.1 vol%. However,
below CIB E , as shown in Fig. 1b. it decreased when the concentration was increased above 0.1 vol%,
More interestingly, electric-field-induced birefringence was indicating that the sensitivity to external fields quickly became
observed in the low-concentration nematic phase, the biphase and muted as the mean interflake distance shrank. The nematic GO
the isotropic phase (Fig. 2a). Chemical reduction and electrolysis10 , with a concentration over 0.56 vol% was almost unresponsive to an
which were observed only under the application of a low-frequency applied voltage up to 20 V mm−1 (Fig. 2b).
electric field (below 0.5 kHz) with a sufficiently high amplitude,
were avoided when a 10 kHz electric field was applied, even Extremely large Kerr coefficients
without any insulation layer (Supplementary Fig. 5). We fabricated Finally, to determine accurately the Kerr coefficient, a measure
a 0.1 vol% aqueous GO cell using Cell Type A in Fig. 2a. Clear of optical sensitivity to external fields, a cuvette with parallel
birefringence was induced by the application of a 5 V mm−1 electric electrodes on two opposite walls and a beam-path length of 5 mm
field, which is roughly three orders of magnitude weaker than was used. Because of the appearance of a large initial birefringence
the field for usual molecular LC switching. After the voltage was in high-concentration GO dispersions in the cuvette, Cell Type B
switched off, the field-induced birefringence slowly disappeared, was used instead of the cuvette in the experiments represented
although some weak nematic ordering partially remained (the by the red square data points in Fig. 3a (Supplementary Figs 6
last image in Fig. 2a–i). In contrast, no change was observed in a and 7 for details). The maximum Kerr coefficient was found to
1.1 vol% nematic GO cell up to 20 V mm−1 (Fig. 2a–ii). be approximately 1.8 × 10−5 mV−2 (Fig. 3a). Note that the Kerr
The field-induced birefringence was measured optically using coefficients of nitrobenzene and aqueous 2D gibbsite platelet
Cell Type B, as shown in Fig. 2b, where the vertical axis represents suspensions are of the order of 10−12 and 10−9 mV−2 respectively25 .
the induced birefringence obtained by subtracting the initial Recently, Ashok et al. reported a giant Kerr coefficient of the order
birefringence at 0 V mm−1 from the measured birefringence. of 10−6 mV−2 using thermal rubidium vapour26 . Blue-phase LCs
The plots of field-induced birefringence were without thresholds, have also attracted substantial interest owing to their large Kerr
showed large hysteresis and depended sensitively on the GO coefficient on the order of 10−9 –10−8 mV−2 (refs 9,27,28). Thus, the
concentration23,24 . Because the induced birefringence can be observed Kerr coefficient of GO-LCs on the order of 10−5 mV−2 is
expressed as 1n = 1nSAT × S × Cv (Supplementary Equation (10)), extraordinarily large compared to those of any other Kerr materials.
where 1nSAT is the specific birefringence, S is the order parameter We next considered the microscopic mechanisms behind the
and Cv is the concentration, the magnitude of 1n at 20 V mm−1 observed behaviour of GO-LCs. It has been well documented

NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials 395

© 2014 Macmillan Publishers Limited. All rights reserved


ARTICLES NATURE MATERIALS DOI: 10.1038/NMAT3888

a
0.1 vol%

1 mm

0 V mm–1 2 V mm–1 5 V mm–1 0 V mm–1

1.1 vol%

0 V mm–1 2 V mm–1 5 V mm–1 0 V mm–1

Cell Type A Cell Type B


ITO
Glass ITO

500 μm 300 μm 300 μm 500 μm

b 4.5

4.0
0.01 vol%
3.5
0.03 vol%
3.0
0.06 vol%
Δn (×10–5)

2.5

2.0 0.11 vol%

1.5 0.22 vol%

1.0 0.28 vol%

0.5 0.56 vol%

0
0 5 10 15 20
Electric field (V mm–1)

Figure 2 | Electric-field-induced birefringence. a, Top row: Field-induced birefringence was generated by applying electric fields (10 kHz) to Cell Type A
filled with an aqueous 0.1 vol% GO dispersion. When the field was switched off, the field-induced birefringence almost disappeared, with only slight
nematic aggregation remaining. Bottom row: In the same cell structure with a 1.1 vol% GO LC, no change was detected up to 20 V mm−1 . b, Field-induced
birefringence curves without thresholds were obtained by applying extremely low electric fields using Cell Type B. Part of the hysteresis curve is shown for
the 0.11 vol% sample.

that a GO flake in an aqueous dispersion will be negatively Onsager’s reorientational entropy can be introduced into
charged, with H3 O+ ions forming an electrical double layer on its the field-induced Maxwell–Boltzmann distribution function for
surface29 , which can produce Maxwell–Wagner polarization orders dispersed particles to obtain the Kerr coefficient as32,33
of magnitude larger than that obtained for the other dielectric 1α C
molecular modes. According to the Maxwell–Wagner–O’Konski K ≈ −1nSAT (1)
model30,31 , the anisotropy of the polarizability (1α) of a colloid 15kB T λ 1 − C/CIB
particle depends sensitively on the aspect ratio of the particle, its Here, (1 − C/CIB )−1 comes from Onsager’s excluded volume effect,
volume and the ionic conductivity on its surface. As elucidated and if the concentration is very low and the interparticle interaction
in detail in Supplementary Fig. 8, we calculated the theoretical is negligible, the Kerr coefficient is determined purely by the electric
1α value of a GO flake with an aspect ratio of 3,200 using the potential energy and (1 − C/CIB )−1 ∼1 (Supplementary Equation
known material parameters of GO flakes and H3 O+ ions. The (22) in the discussion of Supplementary Fig. 8). Using the theoretical
theoretical 1α was estimated to be −1.5 × 10−26 F m2 , which is two 1α obtained above, we plotted the Kerr coefficients obtained using
or three orders of magnitude larger than that of any other plate- both the Straley and van der Beek model (equation (1)) and
like or rod-like particles31 . Here, the negative value of 1α indicates the usual Boltzmann distribution without interflake interactions
that the normal direction of the GO plane is likely to be aligned (Supplementary Equation (22)) as a function of concentration in
perpendicular to the electric field. Fig. 3a. The theoretical prediction based on the Onsager–Straley

396 NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials

© 2014 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3888 ARTICLES
a 3.0 a
Exp_cuvette b
Cell Type C
Exp_Cell Type B
2.5
Kerr coefficient(×10–5 m V–2)

Onsager–Straley model
No inter-flake interaction
2.0

1.5

1.0

0.5 c
CIB_E CBN_E Off state
ON state
0.0
0.00 0.04 0.08 0.12 0.16 0.20
Concentration (vol%)
b 4.5
No NaCl
4.0 1.71 × 10–4 M
3.5 1.02 × 10–3 M
5.80 × 10–3 M
3.0
1.71 × 10–2 M
Δn (× 10–5)

2.5
2.0
1.5
1.0
0.5
0.0 Figure 4 | Electro-optic GO-LC device using simple macroscopic wire
0 5 10 15 20
electrodes. a, Substrate with two simple wire electrodes. The average
Electric field (V mm–1)
distance between the two electrodes was 5 mm. b, Photo of the device on a
back-light unit. The electrode was placed on a glass substrate and then
Figure 3 | Extremely large Kerr coefficient and the mechanisms involved. covered by another glass substrate, and the edge of the cell was sealed with
a, Extremely large Kerr coefficients were measured using a cuvette filled PDMS walls of thickness 2 mm. The cell was filled with 0.056 vol% GO-LC
with a GO dispersion as the sample (Supplementary Fig. 6). Below CIB E , and sandwiched between left and right circular polarizers. An electric field
the Onsager–Staley model fitted well with the experimental results, of 20 V at 10 kHz was applied. c, Field-on and field-off states; the contrast
indicating that the nematic ordering interaction partially contributed to the between the on and off states was 15:1.
large Kerr coefficient. It decreased sharply above CIB E . The red squares
represent values calculated from the data in Fig. 2b (Supplementary Fig. 7).
b, The addition of NaCl ions decreased the sensitivity of the electro-optic electrolyte conductivity, and it enhances the interflake interactions
response to external fields, owing to decreases in |1α|. as well, which can be confirmed by the fact that rotational relaxation
slowed down with the addition of NaCl (10−4 to 10−3 M) and by the
observation of GO flocculation with further NaCl addition (0.17 M;
model accords better with the experimental results below CIB E than Supplementary Fig. 9). Although both the decrease in |1α| and
the traditional model (Supplementary Equation (22)). This clearly the increase in interflake interactions may influence the electrical
demonstrates that Onsager’s excluded-volume effect contributed to desensitization, the former seems to contribute more, considering
the large Kerr coefficient near CIB E . In other words, as the flow- that the theoretical |1α| accurately explains the experimental result
induced birefringence persistence time confirmed the existence of and that the increase in kinetic viscosity with NaCl concentration is
nematic ordering interactions below CIB E in Fig. 1b, it can be one order of magnitude smaller than that from the increase of GO
concluded that the nematic ordering interactions encouraged the concentration (Supplementary Fig. 9).
sensitivity to external fields below CIB E .
Optical devices and wide defect-free GO LC
Desensitization of the Kerr effect To demonstrate the significance of the large Kerr coefficient, we
Meanwhile, the Kerr coefficient decreased with further increases made an electro-optic device using two simple wire electrodes
in concentration above CIB E , indicating an increase in interflake separated by 5 mm (Cell Type C in Fig. 4). Note that interdigitated
friction. When the interflake distance became small, van der Waals electrodes are commonly used in liquid-crystal displays (LCDs)
attractions due to deformed ionic clouds around the GO flakes including blue-phase LCDs (ref. 9), in which the distance between
became dominant29 . Such attractive interflake interactions increased two neighbouring electrodes is about 5 µm—one thousandth of that
the frictional effect, which quickly paralysed the extremely sensitive in our sample. Interestingly, our simple optical device worked very
GO dispersion and prevented the Kerr effect. The increase in well under an applied voltage of just 20 V on the electrodes, as shown
interflake friction with concentration can be also confirmed by the in Fig. 4c, although its performance was not comparable to those of
exponentially increasing kinetic viscosity with GO concentration commercial LCDs. Development of a real GO device may require
shown in Supplementary Fig. 1c. an intensive study to synthesize a high-concentration isotropic GO
It was also found that addition of salt dramatically reduced dispersion for highly saturated birefringence, to control the ionic
the sensitivity to external fields beyond a salt concentration influence and precipitation for long-term stability, and to develop
of 1 × 10−5 M (Fig. 3b) and, ultimately, further addition of salt a new driving scheme suitable for electrolyte materials.
completely desensitized the GO-LCs to external fields. The addition In Fig. 2, we demonstrated that order parameter switching was
of salt gives rise to a decrease in |1α|, mainly due to a increase in achievable in a low-concentration GO dispersion, but the director

NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials 397

© 2014 Macmillan Publishers Limited. All rights reserved


ARTICLES NATURE MATERIALS DOI: 10.1038/NMAT3888

a D1 D2 Cell Type D Cell Type E


F C C 2 mm
0.5 mm 1 mm

S C S C S S

300 μm ITO GO droplet


F S C Aluminium

F C C b 0 V mm–1 20 V mm–1
0.056 vol% 0.056 vol%
S C S C S S

1 mm F S C
1 mm

Δn switching Drying about 1 h at room temp.


1.0 1.0 1.0 Director rotating
0.8 0.8 0V mm–1 20 V mm–1
Trans. (a.u.)

Trans. (a.u.)

After drying After drying


0.6 0.6
0.4 0.4
0.2 Δn switching 0.2
Director rotating
0.0 0.0
–2.5 0.0 2.5 5.0 7.5 10 –2.5 0.0 2.5 5.0 7.5 10
Time (s) Time (s)

Figure 5 | Director switching of induced GO-LC alignment and field-induced wide-area uniform nematic GO-LC phase. a, In Cell Type D, filled with
0.056 vol% GO dispersion, the director of the field-induced GO alignment rotated with changes in field direction. S, a.c. signal; C, common signal; F,
floating electrode. 1n switching curves on the bottom left and right represent the field-on (0 V → D1) and field-off (D1 → 0 V) times, respectively.
Director rotating curves correspond to the rising (D2 → D1) and falling (D1 → D2) times, respectively. The dynamics of the director rotating (D1 ↔ D2)
was faster than that of the order parameter or 1n switching (0 V ↔ D1). The rising and falling times between 10% and 90% luminance were
approximately 1.0 and 4.5 s, respectively, for director switching, and 6.5 and 120 s, respectively, for order-parameter switching. b, In Cell Type E, the
isotropic GO phase turned into the LC phase as the water content evaporated. Whereas the GO on the substrate without fields exhibited a deformed LC
director (left), the GO on the substrate with fields exhibited a well-controlled, uniform LC phase (right).

of the spontaneously aligned nematic GO-LCs was not switchable. For the electro-dynamics experiments, the five different types of cells shown
In contrast, we successfully controlled the director of the field- in Fig. 2 (Cell Types A and B), Fig. 4 (Cell Type C) and Fig. 5 (Cell Types D and
induced GO alignment at a low concentration using a 0.056 vol% E) were used. The substrates for these cells were transparent glass (0.5-mm
thickness) and the electrodes in Cell Types A, B and D were indium–tin–oxide
GO cell (Fig. 5a). The director of the field-induced GO was easily (ITO) layers with 100-nm thicknesses. Cell Type E had 0.5-mm-thick aluminium
rotated by changing the direction of the electric fields and optical electrodes. The dimensions are given in each illustration in Figs 2 and 5.
switching behaviour similar to that of a typical nematic LC device Depending on the purpose of the experiment, all the cells were designed to
was obtained. Compared to the order-parameter switching, the produce different field profiles.
switching of the induced director was faster, as shown at the bottom
part of Fig. 5a. Received 26 August 2013; accepted 20 January 2014;
Using Cell Type E with a single substrate, we obtained a published online 9 March 2014
wide-area, electrically aligned, high-concentration nematic GO
dispersion by evaporating some of the water contained in isotropic
aqueous GOs while electric fields were applied (Fig. 5b). An References
1. Zakri, C. et al. Liquid crystals of carbon nanotubes and graphene. Phil. Trans.
isotropic 0.056 vol% aqueous GO dispersion was released between
R. Soc. A 371, 20120499 (2013).
two electrodes on the substrate and left under ambient conditions to 2. Hernandez, Y. et al. High-yield production of graphene by liquid-phase
allow the water content to slowly evaporate. The field-induced GO exfoliation of graphite. Nature Nanotech. 3, 563–568 (2008).
alignment in the isotropic state was sustained after the GO changed 3. Eda, G., Fanchini, G. & Chhowalla, M. Large-area ultrathin films of reduced
to the nematic phase as the dispersion dried, and the uniform graphene oxide as a transparent and flexible electronic material. Nature
GO-LC alignment was not disturbed even after the fields were Nanotech. 3, 270–274 (2008).
4. Loh, K. P., Bao, Q., Eda, G. & Chhowalla, M. Graphene oxide as a chemically
switched off. In contrast, in the cell that dried without an applied
tunable platform for optical applications. Nature Chem. 2, 1015–1024 (2010).
electric field, the GO formed a disordered LC alignment with many 5. Behabtu, N. et al. Spontaneous high-concentration dispersions and liquid
brushes (left Fig. 5b). This clearly shows that GO-LC alignment crystals of graphene. Nature Nanotech. 5, 406–411 (2010).
with arbitrary sizes and shapes can be achieved by designing an 6. Williams, G., Seger, B. & Kamat, P. V. TiO2 -graphene nanocomposites
appropriate electric field. UV-assisted photocatalytic reduction of graphene oxide. ACS Nano 2,
1487–1491 (2008).
7. Dreyer, D. R., Park, S., Bielawski, C. W. & Ruoff, R. S. The chemistry of
Methods graphene oxide. Chem. Soc. Rev. 39, 228–240 (2010).
The preparation of the graphene oxide samples and their characterizations are 8. Demus, D., Goodby, J. W., Gray, G. W. & Spiess, H-W. Handbook of Liquid
described in detail in Supplementary Fig. 1. The quality and dimensions of GO Crystals Vol 1 (Wiley-VCH, 1998).
flakes varied depending on the batch and the sonicating time, so we used a GO 9. Hisakado, Y., Kikuchi, H., Nagamura, T. & Kajiyama, T. Large electro-optic
sample with a mean size of 3.2 µm from a single batch to maintain consistency. Kerr effect in polymer stabilised liquid-crystalline blue phases. Adv. Mater. 17,
The determination of the sequence of uniform phases of the aqueous GO 96–98 (2005).
dispersion, which had a second-order transition between the isotropic and 10. Kim, J. E. et al. Graphene oxide liquid crystals. Angew. Chem. Int. Ed. 50,
nematic phases, is described in Supplementary Fig. 3. 3043–3047 (2011).

398 NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials

© 2014 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3888 ARTICLES
11. Dan, B. et al. Liquid crystals of aqueous, giant graphene oxide flakes. Soft. 27. Haseba, Y., Kikuchi, H., Nagamura, T. & Kajiyama, T. Large electro-optic Kerr
Matter. 7, 11154–11159 (2011). effect in nanostructured chiral liquid-crystal composites over a wide
12. Dierking, I., Scalia, G. & Morales, P. Liquid crystal–carbon nanotube temperature range. Adv. Mater. 17, 2311–2315 (2005).
dispersions. J. Appl. Phys 97, 044309 (2005). 28. Chen, Y., Xu, D., Wu, S-T., Yamamoto, S-I. & Haseba, Y. A low voltage and
13. Dierking, I., Scalia, G., Morales, P. & LeClere, D. Aligning and reorienting submillisecond-response polymer-stabilized blue phase liquid crystal. Appl.
carbon nanotubes with nematic liquid crystals. Adv. Mater. 16, 865–869 (2004). Phys. Lett. 102, 141116 (2013).
14. Tie, W. et al. Dynamic electro-optic response of graphene/graphitic flakes in 29. Dimiev, A. M., Alemany, L. B. & Tour, J. M. Graphene oxide origin of acidity, its
nematic liquid crystals. Opt. Express 21, 19867–19879 (2013). instability in water and a new dynamic structural model. ACS Nano. 7,
15. Onsager, L. The effects of shape on the interaction of colloidal particles. Ann. 576–588 (2013).
NY Acad. Sci. 51, 627–659 (1949). 30. O’Konski, C. T. Electric properties of macromolecules. V. Theory of ionic
16. Bates, M. A. & Frenkel, D. Nematic–isotropic transition in polydisperse polarization in polyelectrolytes. J. Phys. Chem. 64, 605–619 (1960).
systems of infinitely thin hard platelets. J. Chem. Phys. 110, 6553–6559 (1999). 31. Dozov, I. et al. Electric-field-induced perfect anti-nematic order in isotropic
17. Zhang, S. & Kumar, S. Carbon nanotubes as liquid crystals. Small 4, aqueous suspensions of a natural beidellite clay. J. Phys. Chem. B 115,
1270–1283 (2008). 7751–7765 (2011).
18. Xu, Z. & Gao, C. Aqueous liquid crystals of graphene oxide. ACS Nano 5, 32. Straley, J. P. The gas of long rods as a model for lyotropic liquid crystals. Mol.
2908–2915 (2011). Cryst. Liq. Cryst. 22, 333–357 (1973).
19. Davis, V. A. et al. Phase behavior and rheology of SWNTs in superacids. 33. Van der Beek, D. et al. Magnetic-field-induced orientational order in the
Macromolecules 37, 154–160 (2004). isotropic phase of hard colloidal platelets. Phys. Rev. E 73, 041402 (2006).
20. Hummers, W. S. & Offema, R. E. Preparation of graphite oxide. J. Am. Chem.
Soc. 80, 1339 (1958). Acknowledgements
21. Hontoria-Lucas, C., Lopez-Peinado, A. J., Lopez-Gonzalez, J. de D., This work was supported by the National Research Foundation of Korea (NRF) grants
Rojas-Cervantes, M. L. & Martin-Aranda, R. M. Study of oxygen-containing funded by the Korea government (MSIP) (No. 2012R1A1A1012167 and No.
groups in a series of graphite oxides: physical and chemical characterization. 2013R1A1A2057455).
Carbon 33, 1585–1592 (1995).
22. Aboutalebi, S. H., Gudarzi, M. M., Zheng, Q. B. & Kim, J-K. Spontaneous
Author contributions
formation of liquid crystals in ultralarge graphene oxide dispersions. Adv.
J-K.S. planned and supervised the project. T-Z.S. and S-H.H. carried out all the
Funct. Mater. 21, 2978–2988 (2011). experiments. All authors analysed the data and participated in writing the manuscript.
23. Chen, K-M., Gauza, S., Xianyu, H. & Wu, S-T. Hysteresis effects in blue-phase
liquid crystals. J. Disp. Technol. 6, 318–322 (2010).
24. Rao, L. et al. Critical field for a hysteresis-free BPLC device. J. Disp. Technol. 7, Additional information
627–629 (2011). Supplementary information is available in the online version of the paper. Reprints and
25. Jimenez, M. L., Fornasari, L., Mantegazza, F., Mourad, M. C. & Bellini, T. permissions information is available online at www.nature.com/reprints.
Electric birefringence of dispersions of platelets. Langmuir 28, 251–258 (2012). Correspondence and requests for materials should be addressed to J-K.S.
26. Mohapatra, A. K., Bason, M. G., Butscher, B., Weatherill, K. J. & Adams, C. S. A
giant electro-optic effect using polarizable dark states. Nature Phys. 4, Competing financial interests
890–894 (2008). The authors declare no competing financial interests.

NATURE MATERIALS | VOL 13 | APRIL 2014 | www.nature.com/naturematerials 399

© 2014 Macmillan Publishers Limited. All rights reserved

You might also like