Vous êtes sur la page 1sur 12

This article was downloaded by: [Northwestern University]

On: 27 December 2014, At: 05:10


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Liquid Crystals
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/tlct20

New insights into the transitional behaviour of


methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside using
variable temperature FTIR spectroscopy and X-ray
diffraction
a b c c a
A.G. Cook , A. Martinez-Felipe , N.J. Brooks , J.M. Seddon & C.T. Imrie
a
Department of Chemistry, School of Natural and Computing Sciences, University of
Aberdeen, Aberdeen AB24 3UE, Scotland, UK
b
Institute of Materials Technology, Universitat Politècnica de València, 46022 Valencia,
Spain
c
Department of Chemistry, Imperial College of Science, Technology and Medicine, London
SW7 2AY, UK
Published online: 01 Nov 2013.

To cite this article: A.G. Cook, A. Martinez-Felipe, N.J. Brooks, J.M. Seddon & C.T. Imrie (2013) New insights into the
transitional behaviour of methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside using variable temperature FTIR spectroscopy and
X-ray diffraction, Liquid Crystals, 40:12, 1817-1827, DOI: 10.1080/02678292.2013.854556

To link to this article: http://dx.doi.org/10.1080/02678292.2013.854556

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Liquid Crystals, 2013
Vol. 40, No. 12, 1817–1827, http://dx.doi.org/10.1080/02678292.2013.854556

INVITED ARTICLE
New insights into the transitional behaviour of methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside
using variable temperature FTIR spectroscopy and X-ray diffraction
A.G. Cooka, A. Martinez-Felipeb, N.J. Brooksc, J.M. Seddonc and C.T. Imriea*
a
Department of Chemistry, School of Natural and Computing Sciences, University of Aberdeen, Aberdeen AB24 3UE, Scotland,
UK; bInstitute of Materials Technology, Universitat Politècnica de València, 46022 Valencia, Spain; cDepartment of Chemistry,
Imperial College of Science, Technology and Medicine, London SW7 2AY, UK

The liquid crystalline behaviour of methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside, 1, has been characterised


using X-ray diffraction and variable temperature Fourier transform infrared (FTIR spectroscopy). 1 exhibits a
monotropic interdigitated smectic A phase consisting of bilayers in which the alkyl chains are overlapped. The
crystal–isotropic transition is accompanied by a pronounced decrease in the strength of the hydrogen bonding
Downloaded by [Northwestern University] at 05:10 27 December 2014

network involving the sugar groups resulting in a marked change in the environment of the alkyl chains. The
isotropic phase consists of disordered smectic-like domains stabilised via hydrogen bonding between the sugar
groups. At the transition to the smectic A phase, a subtle change in hydrogen bonding is observed which is
manifested by a change in the temperature dependence of the OH stretching peak position in the FTIR spectrum.
On crystallisation, the strong hydrogen bonding network is re-established accompanied by a change in the
conformational distribution of the alkyl chains. A model is proposed in which a combination of hydrogen
bonding (enthalpic effects) and conformational arrangements (entropic effects) promotes initially the formation
of smectic-like domains in the isotropic phase and subsequently stabilises the smectic A phase by inhibiting the
microphase separation leading to the crystal phase.
Keywords: carbohydrate liquid crystal; smectic A phase; variable temperature FTIR; hydrogen bonding

1. Introduction external stimuli such as UV light or mechanical forces.


Carbohydrate-based liquid crystals have attracted [11–15]
increasing attention in recent years and not only because It is clear that the successful design of new liquid
of their important role in a range of biological processes crystal carbohydrates now requires an improved under-
and considerable application potential in emerging bio- standing of the empirical relationships between molecular
technologies [1–4] but also, more fundamentally, as they structure and aggregation behaviour. With that aim,
provide a demanding challenge to our understanding of recently, we studied the transitional behaviour of a series
self-assembly in condensed systems. [2,3,5,6] The mole- of monosaccharides, the methyl-6-O-(n-acyl)-α-D-gluco-
cular architecture of a carbohydrate-based mesogen typi- pyranosides, as a function of the alkyl chain length, n. [16]
cally consists of a hydrophilic (polar) head group At sufficiently high values of n, the compounds exhibit
attached to one or more lipophilic (nonpolar) chains. [2] thermotropic liquid crystallinity, specifically, a monotro-
This amphiphilic character promotes the formation of pic smectic A phase seen on cooling from the isotropic
microphase-separated hydrophilic and hydrophobic phase, and which can be supercooled by ca 30ºC prior to
phases, and consequently these materials have a great crystallising. On increasing the alkyl chain length further,
tendency to self-assemble and exhibit lyotropic liquid liquid crystallinity is extinguished and this has been attrib-
crystalline behaviour. [7] The hydrophilic domains are uted to back folding of the alkyl chains which disrupts the
widely believed to be stabilised via hydrogen bonding packing of the molecules. [16] We now investigate in
between the hydroxyl groups and the hydrophobic greater detail the role of the different intermolecular inter-
domains by van der Waals interactions between the actions in determining the phase behaviour of the dodecyl
alkyl chains. [2] Carbohydrate-based liquid crystals can member of this homologous series, namely, methyl-6-O-
also exhibit thermotropic mesophases. [8–10] Their excit- (n-dodecanoyl)-α-D-glucopyranoside, 1, shown in
ing application potential includes their use as hydrogels or Scheme 1. [16] Specifically, we have characterised the
electrolytes whose physical properties can be tuned by phase structures shown by 1 using X-ray diffraction,

*Corresponding author. Email: c.t.imrie@abdn.ac.uk


Present addresses for Martinez-Felipe: Departamento de Química Orgánica, Facultad de Ciencias, Instituto de Ciencia de Materiales
de Aragón (ICMA), Universidad de Zaragoza-CSIC, Pedro Cerbuna, Zaragoza 50009, Spain and Centro de Tecnologias Físicas,
Unidad Asociada ICMM/CSIC-UPV, Universitat Politècnica de València Av, Los Naranjos s/n, Valencia 46022, Spain.

© 2013 Taylor & Francis


1818 A.G. Cook et al.

(CH2)10CH3
O
H
HO O
H C(MeO)
HO H H
OH C(1) O(1)
H OMe
O(5)
O(2)
C(6) C(2)
C(7)
C(17) C(15) C(13) C(11) C(9) C(3)
O(6) O(3)
C(12) C(10) C(8) C(5)
C(18) C(16) C(14) C(4)

O(7) O(4)

Scheme 1. Chemical structure of 1, methyl-6-O-dodecanoyl-α-D-glucopyranoside, and the numbering scheme used to refer to
the atoms in the molecule.
Downloaded by [Northwestern University] at 05:10 27 December 2014

and attempted to correlate these with the different chemi- to a Continuum FTIR microscope equipped with a
cal environments of the functional groups deduced using Linkam FTIR 600 heating stage and a TMS 93 con-
temperature-dependent Fourier transform infrared trol unit. The sample was sandwiched between a gold-
(FTIR) spectroscopy. [17] It is important to note that coated microscope slide and a 3 mm thick KBr disc,
the monotropic character of 1 [16] allows us to study and a thin film of the sample was obtained by heating
both the crystal-isotropic and isotropic-smectic A transi- it into the isotropic phase. The FTIR spectra were
tions for the same compound, and to probe the differences obtained on subsequent heating and cooling scans in
in the change in molecular interactions occurring at these the 18–93ºC temperature range, with scan rates of
transitions which would be impossible to achieve for an ±3ºC/min. The spectra were collected over 64 scans
enantiotropic material. with a resolution of 4 cm−1, in the 4000 to 400 cm−1
range, and the samples were run in duplicate and the
average was taken. The spectra were recorded in a
2. Materials and experimental trans-reflectance mode, i.e., the IR beam passed
Scheme 1 shows the chemical structure of 1 and defines through the sample was reflected by the gold surface
the numbering scheme used to refer to the atoms within back through the sample, and then detected. [19]
the molecule. The synthesis of methyl-6-O-dodeca-
noyl-α-D-glucopyranoside, 1, and the characterisation
3. Results and discussion
of its thermal behaviour have been described in detail
elsewhere. [16] X-ray diffraction measurements were 3.1. Phase behaviour of 1
carried out using a custom-built beamline based on a Methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside, 1,
Bede Microsource system, with a Photonic Science exhibits a monotropic smectic A phase. The crystal–
CCD detector. The sample, in a 1.5 mm thin-wall isotropic transition temperature (TCrI) is 73ºC while
glass capillary, was mounted in a brass holder with the smectic A–isotropic transition temperature
mylar windows in the X-ray sample chamber and (TSmAI) is 65ºC. Both temperatures were measured
heated into the isotropic phase at 74.7°C. After mea- on heating using differential scanning calorimetry.
suring the diffraction pattern, the sample was cooled in [16] Thus to measure TSmAI, the isotropic phase was
0.5°C steps to 62.8°C, at each temperature waiting 15 cooled to ca 50ºC to obtain the smectic A phase and
seconds, taking a 15-second exposure, waiting a further immediately heated to above 70ºC. On cooling the
15 seconds, then taking a second 15-second exposure. sample in the DSC, the smectic A phase crystallised at
The temperature fluctuation was ± 0.2°C. As the sam- about 30ºC.
ple remained in the smectic A phase after 10 minutes, For the X-ray diffraction measurements, a
the temperature was further lowered in 1°C steps to crystalline sample of 1 was heated through its melting
57.7°C. The images were analysed using AXcess soft- point of 73°C. The isotropic phase at 74.7°C shows a
ware developed in-house, [18] and calibrated with sil- relatively sharp, yet diffuse small-angle peak
ver behenate (d = 58.38 Å). (FWHM = 0.012 Å−1) centred at a spacing of 27.2
The IR spectra as a function of temperature were Å, and a very broad wide-angle peak centred at a
obtained using a Nicolet Nexus FTIR bench attached spacing of 4.65 Å, see Figure 1. The presence of a
Liquid Crystals 1819

1500 4000

3000
1000
Intensity

Intensity
2000

500
1000

0
0
0.0 0.1 0.2 0.3
0.0 0.1 0.2 0.3
–1
s/Å s / Å–1

Figure 1. X-ray pattern of the isotropic phase exhibited by Figure 3. X-ray pattern of the smectic A phase exhibited by
1 at 74.7ºC. 1 at 58.5ºC.
Downloaded by [Northwestern University] at 05:10 27 December 2014

strong and relatively sharp low-angle peak implies that Since the smectic A phase is monotropic, it is meta-
there are residual disordered smectic-like domains sur- stable and can spontaneously transform to the more
viving within the isotropic phase. These are presum- stable crystal phase, via a nucleation and growth
ably stabilised by hydrogen bonding between the sugar mechanism, at any temperature upon cooling below
head groups and we will discuss this issue later. Upon TCrI. For the X-ray sample studied, crystallisation
cooling, a transition to a smectic A phase occurred at occurred at 57.7°C, and was complete within a few
64.7°C, and is marked by an abrupt sharpening of the seconds. The crystal phase exhibits sharp first- and
small-angle peak into a single, first-order Bragg peak second-order Bragg peaks (FWHM = 0.004 Å−1) in
(FWHM = 0.004 Å−1) at a layer spacing of 28.7 Å, the low-angle region, with a layer spacing of 27.8 Å,
see Figure 2. There is the merest hint of a second-order see Figure 4. The wide-angle region contains a number
Bragg peak, see Figure 3, and this fact, together with of sharp peaks, confirming the crystallinity of this
the absence of any higher-order peaks, implies that the phase.
electron density along the layer normal is close to The crystal structure of 1 has been reported pre-
sinusoidal, as is commonly observed with smectic A viously [20] and consists of bilayers composed of
phases. The wide-angle peak remains very diffuse and either the sugar head groups or interdigitated alkyl
centred at the same spacing as for the isotropic phase, chains. The interface between these sub-layers was
proving this to be a fluid mesophase. We can rule out taken to be the plane within the layer containing the
the possibility of a smectic C phase both from the C8 atoms (for the numbering scheme, see Scheme 1).
optical texture, and from the fact that fluid hydrocar- The sugar moieties are disordered within the crystal
bon chains are unable to sustain a cooperative tilt. structure while the ester and alkyl chain fragments are

60

1500

65
1000
T/°C

Intensity

70
500

75 0
0.1 0.2 0.3 0.0 0.1 0.2 0.3
s / Å–1 s / Å–1

Figure 2. (colour online) Temperature cooling scan of the Figure 4. X-ray pattern of the crystal phase exhibited by 1
small-angle diffraction pattern exhibited by 1. at 57.7ºC.
1820 A.G. Cook et al.

ordered with adjacent alkyl planes lying perpendicu- Log (1/R)


CH st.
lar to each other. The c lattice parameter is 28.443 Å 3
C = O st.
which is in good agreement with the layer spacing of OH st.
2.5
27.8 Å reported here for the crystal phase. Hydrogen (a)
bonding involving the hydroxyl groups on the sugar 2
moieties within the hydrophilic sub-layer was consid-
1.5
ered to be the driving force stabilising the crystal
structure. [20] If we now return to the molecular 1
arrangement within the smectic A phase, then we
0.5
have seen that the smectic periodicity is 28.7 Å. The
estimated molecular length of 1 is 22.3 Å such that 0
the ratio of the smectic periodicity to molecular 3600 3200 2800 2400 2000 1600
length, d/l, is 1.3, implying that the smectic phase Wavenumber (cm–1)
must have an interdigitated structure in which the Log (1/R) C–O st. + δ
alkyl chains overlap considerably. Figure 5 shows a 3 (b)
sketch of the molecular organisation within the phase C(6), C(7) Pyranoside Ring
Downloaded by [Northwestern University] at 05:10 27 December 2014

2.5
consistent with this d/l value and is similar to that CH2 δ
CC st.
commonly observed for carbohydrate-based liquid 2 COH δ
crystals. [21,22]
C(5)O(5)C(1)
1.5
CH2 ρ

3.2. FTIR spectrum of 1 in the crystal phase 1

The 3700–1600 cm−1 and 1600–700 cm−1 regions of the 0.5


FTIR spectrum of 1 obtained for a crystalline sample
at 18ºC are shown in Figures 6(a) and 6(b), respec- 0
tively. The broad band between 3600–3000 cm−1 is 1500 1300 1100 900 700
highly characteristic of stretching (st.) vibration Wavenumber (cm–1)
modes of OH bonds (Figure 6(a)). The band is centred
Figure 6. The (a) 3700–1600 cm−1 and (b) 1600–700 cm−1
at lower wavenumbers than that characteristic of free regions of the FTIR spectrum of 1 obtained in the crystal
OH groups (sharp band expected at ν ≥ 3500 cm−1), state (T = 18ºC). The numbering scheme for the atoms is
indicating extensive hydrogen bonding in the crystal shown in Scheme 1.
phase. [23–25] This is consistent with the crystal struc-
ture described earlier, in which the OH groups are
closely packed, and the suggestion that hydrogen The stretching vibration of the ester carbonyl bond,
bonding stabilises the crystal phase. [20] The 3000– C=O st., is associated with at least two strong bands
2700 cm−1 region contains several strong bands asso- located at 1724 and 1735 cm−1, which may be assigned
ciated with C–H stretching. The peaks observed at to carbonyl groups more, or less, strongly involved in
2854 and 2845 cm−1 may be assigned to the C–H hydrogen bonding, respectively.
stretching mode of the carbon atoms neighbouring At lower frequencies, see Figure 6(b), the peaks
ester, C(6), and ether, C(1) and oxygen atoms. [24,26] around 1470 cm−1 are associated with the C–H δ
mode (scissoring vibrations of the alkyl chain C–H)
while the band at 1417 cm−1 may arise from bending
28.7 Å of the C(6)–H bonds, shifting to lower frequencies by
22.3 Å
the proximity of the O(6) atom. The 1400–1200 cm−1
region contains several overlapping bands associated
with the deformation of C–H bonds and C–C stretch-
ing, i.e., dominated by contributions arising from the
aliphatic segment of the molecule. A number of peaks
can be seen between 1320 and 1220 cm−1 and these
are assigned to bending deformations of the C–H
bonds adjacent to OH groups (C(2–4)), oxygen
atoms (C(1), C(5), C(6)) or carbonyl groups (C(8)).
Several strong IR contributions characteristic of
Figure 5. (colour online) Schematic representation of the
local molecular arrangement of 1 found in the smectic A monosaccharides are observed in the complex region
phase. below 1200 cm−1, including C–O and C–C stretching
Liquid Crystals 1821

and deformation modes of the oxymethylene group. 3.3. Temperature dependence of the FTIR spectra of
[26] The two strong bands located at 1196 and 1 on heating
1141 cm−1 may be attributed to stretching vibrations The high- and low-frequency regions of the FTIR
of the C(6)–O and C(7)–O bonds, respectively. spectra of 1 obtained on heating from 18ºC to 93ºC
Between 1120 and 1000 cm−1, we see bands associated are shown in Figures 7(a) and 7(b), respectively.
with mixed stretching and bending vibrations of C–C Several pronounced changes in the spectra are seen
and C–O bonds in the glucopyranoside ring. At lower around the crystal–isotropic phase transition, ca
frequencies, ca 1000–950 cm−1, contributions arising 73ºC. In particular, the OH st. band shifts to
from the oxymethylene group (deformation modes of higher frequencies and decreases in intensity, see
O(1)–C(OMe)–H and stretching vibrations of C Figure 7(a). Thus, in Figure 8, we show the tempera-
(OMe)–O(1)) are seen. [26] The bands present in the ture dependence of the maximum of the OH st. band
950–800 cm−1 region of the spectrum may be asso- peak position, and two distinct regions of behaviour
ciated with the complex stretching vibration of the on heating are seen: below the melting point, the peak
chain of atoms O(6)–C(6)–C(5)–O(5), and their shifts to higher frequency on increasing temperature
appearance is particularly sensitive to configurational while in the isotropic phase appears to remain essen-
and conformational transformations in monosacchar- tially constant. Melting is accompanied by a jump in
Downloaded by [Northwestern University] at 05:10 27 December 2014

ides. [26] The envelope of this region is similar to that the position of the OH st. band of around 70 cm−1 to
observed for glucopyranoside, but contains a new peak higher wavenumbers. These changes indicate a dra-
at 855 cm−1 associated with vibrations involving the matic reduction in the strength of hydrogen bonding
chain of atoms C(8)–C(7)–O(6)–C(6) containing within the system. However, the breadth of the
the ester linkage. Finally the bands seen below OH band and the absence of a sharp peak above
800 cm−1 are related to rocking deformations, ρ, of 3500 cm−1 indicate the persistence of hydrogen bond-
C–H bonds. [27] ing in the isotropic phase, at least to some degree.

Log (1/R)

Heating (a)

3500 3000 2500 1800 1750 1700 1650


Wavenumber (cm–1)

Log (1/R)

(b)
Heating

1500 1300 1100 900 700


Wavenumber (cm–1)

Figure 7. Temperature dependence of the (a) 4000–2600 cm−1 and (b) 1800–700 cm−1 regions of the FTIR spectra of 1
obtained on heating. Large arrows indicate the direction of heating (T = 18, 28, 38, 48, 58, 68, 78 and 88ºC). Dashed arrows
and the grey spectrum indicate the phase transition.
1822 A.G. Cook et al.

ν max (cm–1) this region of the spectra. However, it is clear that


3480 T = 64ºC
WTCI = –0.46 cm–1/K on melting the higher frequency band (ca. 1735 cm−1)
increases in intensity to a much greater extent than
3460
WTCSmA = –1.22 cm–1/K the 1724 cm−1 component (see Figure 7), implying a
3440 reduction in the intermolecular interactions involving
3420 the carbonyl groups. These changes in the carbonyl
3400
stretching region reinforce the view that the strength
of hydrogen bonding is reduced on melting the crystal
3380
WTCCr = 0.79 cm–1/K
phase.
3360 At lower wavenumbers, the 1500–1300 cm−1 region
3340 of the FTIR spectra also shows significant variations
0 20 40 60 80 100 on heating, see Figure 7(b), revealing changes in the
T (°C) hydrophobic regions of the molecule. Increasing
the temperature causes shifts to lower frequencies of
Figure 8. (colour online) Temperature dependence of the
the peaks related to the methylene groups, located
maximum of the OH stretching band on heating from the
principally in the alkyl chain, see, for example, the
Downloaded by [Northwestern University] at 05:10 27 December 2014

crystal phase (red diamonds) and subsequent cooling from


the isotropic phase (blue squares). temperature dependence of the rocking C–H band at
ca. 763 cm−1 in Figure 9. On heating the crystal phase,
the peak position shows a small linear decrease. By
This is consistent with the view that hydrogen bond- comparison, on melting, a sudden shift to lower fre-
ing stabilises disordered smectic-like domains in the quency is seen while on heating the isotropic phase, the
isotropic phase as observed by X-ray diffraction, see peak position shows little change. These observations
Figure 1. reveal, at least in part, changes in the van der Waals
We investigated the dynamic nature of the hydro- forces between the molecules on heating.
gen bonding network by calculating the so-called The region associated with the stretching/bending
wavenumber-temperature-coefficient, WTC, of the vibrations of the pyranose ring also shows
OH st. band which was defined by Wolkers et al. noticeable changes on increasing the temperature.
[23,28] as the gradient of the plot of peak position For example, the peaks related with the methoxy
against temperature and has been interpreted in glasses group (1330–1220 cm−1) remain essentially
in terms of Angell’s strength and fragility concept. unchanged on heating the crystal phase, while at
[29,30] The regions of linear behaviour are also indi- around the melting point rapidly decrease in intensity
cated in Figure 8, and we found a WTCCr value of 0.79 and wavenumber. In the isotropic phase, a rather
cm−1 K−1 in the crystal phase. This is rather higher smooth FTIR profile is observed for this region,
than those typically measured in a sugar-based glass or
melt [28] and presumably reflects an equilibrium shift-
ing towards free OH groups with temperature. ν max (cm–1)
The C=O stretching region is essentially
780
unchanged on heating the crystal phase, but its inten-
sity increases dramatically at the crystal–isotropic 776
transition, see Figure 7(a). This dramatic increase in 772
intensity, which is also seen for other bands, for 768
example, 2930 and 2850 cm−1 (see Figure 7(a)), may 764
be attributed to a loss of surface alignment of the 760
liquid crystal on entering the isotropic phase.
756
Specifically, the incident IR beam is polarised, at
752
least to some extent, on reflection at the gold surface
and so vibrations parallel to the surface will not be 748
detected as strongly as those perpendicular to the 744
surface. [31] Given the strong tendency of carbohy- 740
drate-based liquid crystals to align homeotropically, 0 20 40 60 80 100
the intensity of vibrations perpendicular to the mole- T (°C)
cular long axis, including, for example, the C=O
Figure 9. (colour online) Temperature dependence of the
stretch, will have reduced intensities. [30] The tem- characteristic IR band associated with the rocking mode of
perature dependence of the carbonyl peak positions is C–H bonds on heating (red diamonds) and cooling (blue
difficult to quantify because of the complex nature of squares).
Liquid Crystals 1823

indicating a greater homogeneity of the chemical bonding. A similar behaviour is observed in non-meso-
interactions between the rings. These results presum- genic sugars and the magnitude of the position change
ably reflect the changes of the local structure arising of the OH st. peak appears to be linked to the size of the
from the abrupt change in hydrogen bonding, leading crystallisation enthalpy. [32] In the crystal phase, the
to a higher disordering of the sugar groups than in the spectra recorded on cooling are essentially identical to
crystal phase. those described in the previous section on heating.
The peak frequency of the OH st. band is some 60
cm−1 higher in the smectic A phase than in the crystal
3.4. Temperature dependence of the FTIR spectra of phase, indicative of weaker hydrogen bonding. In
1 on cooling addition, we have seen that the WTC value recorded
Figure 10 shows the FTIR spectra of 1 recorded on on cooling the smectic A phase, 1.22 cm−1/K, is much
cooling from the isotropic phase. The OH st. band larger than that seen on heating the crystal phase,
shifts to lower wavenumbers, with an associated 0.79 cm−1/K, indicating that the hydrogen bond net-
WTCI value of 0.46 cm−1/K, see Figure 8, revealing work is more sensitive to changes in temperature in
changes in the hydrophilic regions of 1. This value is the mesophase. This latter observation presumably
comparable to those measured in the isotropic phases reflects that the increase in the order parameter on
Downloaded by [Northwestern University] at 05:10 27 December 2014

of other conventional low-molar mass sugars. [23] cooling the smectic A phase is considerably larger
There is no step change in either the peak position or than the change in ordering on heating the crystal.
intensity of the band at the isotropic–smectic A transi- This is only the second WTC value reported for a
tion but we do see an increase in the associated carbohydrate-based liquid crystal and is considerably
WTCSmA value to 1.22 cm−1/K, see Figure 8. On crys- larger than that reported previously for a nonsym-
tallisation, the peak maximum jumps to a lower value metric liquid crystal dimer containing a single carbo-
and shows a sudden and pronounced increase in inten- hydrate-based unit. [30] The physical significance of
sity, which indicates a step increase in hydrogen this difference in behaviour is not yet clear.

(a)

Log (1/R)

CT100C_B

3500 3000 2500 1800 1750 1700 1650


–1)
Wavenumber (cm

Log (1/R)

(b)
Cooling

1500 1300 1100 900 700


Wavenumber (cm–1)

Figure 10. Temperature dependence of the (a) 4000–2600 cm−1 and (b) 1800–600 cm−1 regions of the FTIR spectra of 1
obtained on cooling. Large arrows indicate the direction of cooling (T = 88, 78, 68, 58, 48, 38 and 28ºC). Dashed arrows and
the grey spectrum indicate the phase transition.
1824 A.G. Cook et al.

Cooling the isotropic phase is also accompanied unchanged, see Figure 11(a). On melting, we see a
by changes in the IR peaks related to the alkyl groups dramatic decrease in intensity for the highest wave-
of 1 as seen, for example, in Figure 9, which number contribution, a moderate but clear decrease
shows the temperature dependence of the band at in intensity for the 1463 cm−1 band, and an increase
ca. 752 cm−1 associated with the C-H rocking mode. in the intensity of the lower frequency contributions.
The peak position increases slightly and in a linear The changes in this region of the FTIR spectra
fashion on cooling the isotropic and smectic A phases. reflect, at least to some extent, a combination of the
Crystallisation is then accompanied by a sudden complex variations in the intramolecular (conforma-
increase in the peak position suggesting that the tional) and intermolecular factors involving the
alkyl chains are now in an environment of weaker hydrophobic segments of 1 on heating. Specifically,
intermolecular interactions. the 1472 cm−1 band can be assigned to C–H groups
The effect of temperature on the alkyl groups of 1 preferentially found in trans conformations. [33,34]
may be studied in further detail by analysing the The increase in the 1463 cm−1 contribution may
1500–1430 cm−1 region of the FTIR spectra, see then be attributed to a dynamic exchange towards
Figure 11, corresponding to the C–H in-plane (scis- an increased number of gauche configurations of the
soring) vibrations. Despite the complexity of the alkyl chain, indicative of an increase in disorder on
Downloaded by [Northwestern University] at 05:10 27 December 2014

region, at least two high wavenumber bands are evi- melting. In addition, the relative increase of the lower
dent, centred at 1472 and 1463 cm−1, as well as a wavenumber contributions (~1454 cm−1) may be
broad lower wavenumber contribution (~1454 cm−1), understood in terms of a greater number of C–H
presumably consisting of a number of overlapping units being involved in stronger intermolecular inter-
bands and related to methylene groups involved in actions. On cooling, this region shows effectively no
stronger intermolecular interactions. On heating the change in intensity at the smectic A–isotropic transi-
crystal phase, this region of the spectrum is essentially tion, presumably indicating that both the dynamics
and environment of the chains undergo little change
at this transition, see Figure 11(b). This is in agree-
Log (1/R) ment with the behaviour discussed earlier for the C–H
2 T = 88°C rocking bands, see Figure 9. By comparison, on crys-
1.8 (a) tallisation, dramatic changes are seen in this region
T = 53°C
1.6
and, as would be expected, these are opposite in sense
1.4 T = 28°C
to those observed on melting.
1.2
1
0.8
0.6
3.5. Discussion of the phase behaviour of 1
0.4 The changes seen in the FTIR spectra of 1 on heating
0.2 and cooling, described in the preceding sections, pro-
0 vide new insights into the temperature dependence of
1490 1478 1466 1454 1442 1430
the intermolecular environments experienced by the
Wavenumber (cm–1) hydrophobic and hydrophilic segments of 1. In the
reported crystal structure, [35] the pyranose ring is
Log (1/R)
tilted with respect to the long axis of the alkyl chain.
T = 78°C
The alkyl chains are in all-trans conformations facil-
2 T = 48°C
1.8
itating their packing in the hydrophobic domains of
(b)
1.6 T = 28°C the crystal, except for the C(7)–O(6)–C(6)–C(5) bond
1.4 which adopts a gauche conformation allowing the
1.2 sugar head groups to pack efficiently into the hydro-
1 philic layers. This phase separated crystal structure,
0.8
0.6
typical of amphiphilic liquid crystals, is stabilised by
0.4 the strong hydrogen bonds, principally involving the
0.2 OH groups in the sugar moieties.
0 On melting, a fraction of the hydrogen bonds in
1490 1478 1466 1454 1442 1430
the polar regions of the structure is broken (probably
Wavenumber (cm–1)
the strongest, i.e., those associated with O–H st. hav-
Figure 11. (colour online) The 1500–1430 cm−1 region of ing lower frequencies), and this increased molecular
the FTIR spectrum of 1 obtained at three different tempera- freedom allows for conformational rearrangements
tures on (a) heating and (b) cooling. within the system. The geometry of the C(7)–O(6)–C
Liquid Crystals 1825

(6)–C(5) bond changes, resulting in a more linear without the carbonyl link, the methyl-6-O-(n-alkyl)-
molecule, which may assist the formation of the smec- α-D-glucopyranosides, do not exhibit liquid crystal-
tic-like aggregates observed in the isotropic phase, see linity [10], providing further evidence that the carbo-
Figure 5. nyl groups participate in the hydrogen bonding
On supercooling the isotropic phase, the smectic A required for smectic phase stabilisation. From a gen-
phase is formed, and this appears to be facilitated by eral molecular design viewpoint, this observation
the higher free volume available to the hydrophilic highlights the key role of the linker group between
and hydrophobic regions in the melt. The temperature the polar head and the hydrophobic chains in promot-
dependence of the OH st. peak revealed both a subtle ing thermotropic liquid crystallinity in this class of
change in the hydrophilic interactions when cooling sugar-based mesogens.
from the isotropic phase (see Figure 8), and also a
dynamic hydrogen bonding network in the smectic
A phase, reflected in the unusually high WTCSmA 4. Conclusions
(–1.22 cm−1/K) value. It is also interesting to note Methyl-6-O-(n-dodecanoyl)-α-D-glucopyranoside, 1,
that the isotropic–smectic A transition temperature exhibits a monotropic smectic A phase consisting of
seen for 1 is considerably higher than the melting bilayers in which the alkyl chains interdigitate. This
Downloaded by [Northwestern University] at 05:10 27 December 2014

temperatures of either lauric acid (T ~ 40ºC) or study of a monotropic liquid crystal has allowed us to
methyl laureate (T ~ 10ºC), [36] suggesting that the compare the crystal–isotropic and smectic A–isotro-
alkyl chains are expected to be highly mobile in the pic transitions for the same compound.
smectic phase. Thus, it is reasonable to assume that Melting is accompanied by a significant decrease
thermal activation promotes the adoption of a greater in the strength of hydrogen bonding within the hydro-
number of conformations by the hydrophobic frag- philic domains in the crystal phase of 1, although
ments of the molecule above melting. The existence of some hydrogen bonding persists resulting in disor-
this dynamic conformational equilibrium is entropi- dered smectic-like domains in the isotropic phase.
cally favourable and, taken together with additional On cooling the isotropic phase, a subtle change in
kinetic barriers impairing the formation of crystal hydrogen bonding drives the formation of an inter-
nuclei, may offset the unfavourable hydrophilic– digitated smectic phase. Crystallisation is inhibited by
hydrophobic interactions in the interdigitated smectic a combination of enthalpic (persistence of moderate
layers and inhibit crystallisation. hydrogen bonds which stabilise the lamellar order)
Finally, we need to consider the role of the carbo- and entropic factors (the fluid nature of the alkyl
nyl group in determining the phase behaviour of 1. chains and the dynamic nature of the hydrogen
The intermolecular environment experienced by the bond network). Crystallisation of the smectic A
carbonyl group appears little altered at the isotropic– phase is then accompanied by a sudden increase in
smectic transition, but undergoes considerable the strength of hydrogen bonding, leading to a micro-
changes on melting and crystallisation, see Figures 7 phase separated crystal structure. These observations
and 10. The ester group (C(8)–O(7)–O(6)) links the reinforce the view that the melting point of a carbo-
hydrophobic and hydrophilic regions of the molecule, hydrate-based liquid crystal is largely determined by
and its contribution to determining phase behaviour the strength of the hydrogen bonding network within
is twofold. On the one hand, the FTIR spectra show the crystal [37]. By comparison, the van der Waals
that in the crystal phase, the carbonyl group is interactions between the alkyl chains play a rather
involved in the strong hydrogen bonded network secondary role in the formation of the smectic phase
within the hydrophilic layer, thus contributing to the (a view consistent with the absence of change in the
stability of the crystal phase. On the other hand, on FTIR spectra shown, for example, in Figures 9 and
melting, the conformational change of the C(7)–O(6)– 11) but undergo greater change during crystallisation
C(6)–C(5) bond increases the molecular linearity and and melting.
reduces the ability of the carbonyl group to form This model for the phase transition also accounts
strong intramolecular hydrogen bonds. Instead, the for the unusually low values of the entropy change
carbonyl group becomes involved in weaker intermo- observed at the isotropic–smectic A phase transition
lecular hydrogen bonding which helps to stabilise the for 1, [16] which, in fact, is more similar in magnitude
formation of smectic-like aggregates. This view is to that observed for nematic–isotropic transitions.
consistent with both the spectroscopic changes [38–40] This appears to be a rather general observa-
observed in the ν ~1470 cm−1 region, see Figure 11, tion [5,41,42] but has not attracted significant atten-
and with the interdigitated smectic arrangement tion. Thus, the persistence of these hydrogen bonded
found using X-ray diffraction, see Figure 5. It is also aggregates in the isotropic phase reduces the changes
noteworthy that the corresponding materials but in the orientational and translational ordering that
1826 A.G. Cook et al.

would normally be expected at the smectic A–isotro- galactosides in thermotropic liquid crystal self-assem-
pic transition giving a smaller than expected entropy blies. Carbohydr Res. 2011;346(18):2948–2956.
[8] Dorset DL, Rosenbusch JP. Solid-state properties of
change. A similar model has been proposed to
anomeric 1-O-n-octyl-D-glucopyranosides. Chem Phys
account for the low values of the entropy change Lipids. 1981;29(4):299–307.
seen for the smectic A–isotropic transition for ionic [9] Goodby JW. Liquid-crystal phases exhibited by
liquid crystals in which the aggregates are stabilised some monosaccharides. Mol Cryst Liq Cryst.
by ionic interactions. [43] 1984;110(1–4):205–219.
[10] Miethchen R, Holz J, Prade H, Liptak A. Amphiphilic
To the best of our knowledge, this is the first
and mesogenic carbohydrates. 2. Synthesis and char-
detailed study of the temperature dependence of the acterization of mono-O-(n-alkyl)-D-glucose deriva-
FTIR spectrum of a carbohydrate-based liquid crystal tives. Tetrahedron. 1992;48(15):3061–3068.
and has revealed new insights into the phase beha- [11] Martinez-Felipe A. Liquid crystal polymers and iono-
viour of this class of materials. The generality of these mers for membrane applications. Liq Cryst. 2011;38
(11–12):1607–1626.
observations, however, must now be tested for a wide
[12] Kaneko T, Yamaoka K, Gong JP, Osada Y. Liquid-
range of carbohydrate-based liquid crystals. crystalline hydrogels. 1. Enhanced effects of incorpora-
tion of acrylic acid units on the liquid-crystalline order-
ing. Macromolecules. 2000;33(2):412–418.
Downloaded by [Northwestern University] at 05:10 27 December 2014

Acknowledgements [13] Clemente MJ, Fitremann J, Mauzac M, Serrano JL,


Oriol L. Synthesis and characterization of maltose-
The authors would like to thank the Spanish Ministry of based amphiphiles as supramolecular hydrogelators.
Science and Innovation, through the Research Project Langmuir. 2011;27(24):15236–15247.
ENE2011-28735-C02-01, and the financial support of the [14] Clemente MJ, Romero P, Serrano JL, Fitremann J,
Generalitat Valenciana, through the GRISOLIAP/2010/057 Oriol L. Supramolecular hydrogels based on glycoam-
and APOSTD/2013/054 programs. The UPV is also thanked phiphiles: effect of the disaccharide polar head. Chem
for additional support through the PAID-06-11-2037 Mater. 2012;24(20):3847–3858.
program. [15] Clemente MJ, Tejedor RM, Romero P, Fitremann J,
Oriol L. Maltose-based gelators having azobenzene
as light-sensitive unit. Rsc Advances. 2012;2
Funding (30):11419–11431.
[16] Cook AG, Wardell JL, Imrie CT. Carbohydrate liquid
JMS is pleased to acknowledge support from the EPSRC crystals: synthesis and characterisation of the methyl-6-
Platform [grant number EP/G00465X]. O-(n-acyl)-α-D-glucopyranosides. Chem Phys Lipids.
2011;164(2):118–124.
[17] Martinez-Felipe A, Lu Z, Henderson PA, Picken SJ,
References Norder B, Imrie CT, Ribes-Greus A. Synthesis and
characterisation of side chain liquid crystal copolymers
[1] Goodby JW. Liquid-crystalline glycolipids: towards containing sulfonic acid groups. Polymer. 2012;53
understanding the roles of liquid crystals in biological (13):2604–2612.
and life processes. Liq Cryst. 2006;33(11–12): [18] Seddon JM, Squires AM, Conn CE, Ces O, Heron AJ,
1229–1237. Mulet X, Shearman GC, Templer RH. Pressure-jump
[2] Goodby JW, Goertz V, Cowling SJ, Mackenzie G, X-ray studies of liquid crystal transitions in lipids.
Martin P, Plusquellec D, Benvegnu T, Boullanger P, Philosophical Transactions of the Royal Society a-
Lafont D, Queneau Y, Chambert S, Fitremann J. Mathematical Physical and Engineering Sciences.
Thermotropic liquid crystalline glycolipids. Chem Soc 2006;364(1847):2635-2655.
Rev. 2007;36(12):1971–2032. [19] Martinez-Felipe A, Imrie CT, Ribes-Greus A. Study of
[3] Hashim R, Sugimura A, Minamikawa H, Heidelberg structure formation in side-chain liquid crystal copolymers
T. Nature-like synthetic alkyl branched-chain glycoli- by variable temperature fourier transform infrared spectro-
pids: a review on chemical structure and self-assembly scopy. Ind Eng Chem Res. 2013;52(26):8714–8721.
properties. Liq Cryst. 2012;39(1):1–17. [20] Abe Y, Harata K, Fujiwara M, Ohbu K. Molecular
[4] Jewell SA. Living systems and liquid crystals. Liq arrangement and intermolecular hydrogen bonding in
Cryst. 2011;38(11–12):1699–1714. crystals of methyl 6-O-acyl-D-glycopyranosides.
[5] van Doren HA, Smits E, Pestman JM, Engberts J, Langmuir. 1996;12(3):636–640.
Kellogg RM. Mesogenic sugars. From aldoses to liquid [21] Donaldson T, Henderson PA, Achard MF, Imrie CT.
crystals and surfactants. Chem Soc Rev. 2000;29 Chiral liquid crystal tetramers. J Mater Chem. 2011;21
(3):183–199. (29):10935–10941.
[6] Brooks NJ, Hamid HAA, Hashim R, Heidelberg T, [22] Bault P, Gode P, Goethals G, Goodby JW, Haley JA,
Seddon JM, Conn CE, Husseini SMM, Zahid NI, Kelly SM, Mehl GH, Ronco G, Villa P. An homolo-
Hussen RSD. Thermotropic and lyotropic liquid crys- gous series of 6-O-n-alkyl-α-D-galactopyranoses:
talline phases of Guerbet branched-chain ß-D-gluco- synthesis and thermotropic mesomorphic properties.
sides. Liq Cryst. 2011;38(11–12):1725–1734. Liq Cryst. 1998;24(2):283–293.
[7] Hashim R, Mirzadeh SM, Heidelberg T, Minamikawa [23] Wolkers WF, Oliver AE, Tablin F, Crowe JH. A four-
H, Yoshiaki T, Sugimura A. A reevaluation of the ier-transform infrared spectroscopy study of sugar
epimeric and anomeric relationship of glucosides and glasses. Carbohydr Res. 2004;339(6):1077–1085.
Liquid Crystals 1827

[24] Babkov LM, Korolevich MV, Moiseikina EA. [34] Zhu X-Q, Liu J-H, Liu Y-X, Chen E-Q. Molecular
Hydrogen bonding, IR spectrum, and the structure of packing and phase transitions of side-chain liquid crys-
methyl-ß-D-glucopyranoside. J Struct Chem. 2012;53 talline polymethacrylates based on p-methoxyazoben-
(1):55–62. zene. Polymer. 2008;49(13–14):3103–3110.
[25] Vinogradov S, Linnell R. Hydrogen Bonding, 272 p. [35] Abe Y, Harata K, Fujiwara M, Ohbu K. Crystal-
New York (NY): Van Nost Reinhold; 1971. structures of methyl 6-O-n-alkanoyl-ß-D-glucopyrano-
[26] Korolevich FV, Zhbankova MR, Piottukh-Peletskii sides. Carbohydr Res. 1995;269(1):43–51.
VN, Zhbankov RG. Interpretation of the IR spectrum [36] Lide D, editor. CRC handbook of chemistry and phy-
of methyl-ß-D-glucopyranoside based on the theoreti- sics. 86th ed. Boca Raton (FL): CRC Press; 2005.
cal calculation of frequencies and intensities of normal [37] van Doren HA, van Dergeest R, Deruijter CF, Kellogg
vibrations. J Struct Chem. 2007;48(5):821–830. RM, Wynberg H. The scope and limitations of liquid-
[27] Lin-Vien D, Colthup N, Fateley W, Grasselli J, edi- crystalline behavior in monosaccharide amphiphiles –
tors. The handbook of infrared and Raman character- comparison of the thermal-behavior of several homolo-
istic frequencies of organic molecules. Boston (MA): gous series of D-glucose derived compounds with an
Academic Press; 1991. amino-linked alkyl chain. Liq Cryst. 1990;8(1):109–121.
[28] Wolkers WF, Oldenhof H, Alberda M, Hoekstra FA. [38] Imrie CT, Taylor L. The preparation and properties of
A Fourier transform infrared microspectroscopy study low molar mass liquid-crystals possessing lateral alkyl
of sugar glasses: application to anhydrobiotic higher chains. Liq Cryst. 1989;6(1):1–10.
plant cells. Biochim Biophys Acta-General Subjects. [39] Imrie CT, Karasz FE, Attard GS. Comparison of the
Downloaded by [Northwestern University] at 05:10 27 December 2014

1998;1379(1):83–96. mesogenic properties of monomeric, dimeric, and side-


[29] Angell CA, Imrie CT, Ingram MD. From simple elec- chain polymeric liquid-crystals. Macromolecules.
trolyte solutions through polymer electrolytes to super- 1993;26(3):545–550.
ionic rubbers: Some fundamental considerations. [40] Henderson PA, Niemeyer O, Imrie CT. Methylene-
Polym Int. 1998;47(1):9–15. linked liquid crystal dimers. Liq Cryst. 2001;28
[30] Cook AG, Wardell JL, Brooks NJ, Seddon JM, (3):463–472.
Martinez-Felipe A, Imrie CT. Non-symmetric liquid [41] Letellier P, Ewing DF, Goodby JW, Haley J, Kelly
crystal dimer containing a carbohydrate-based moiety. SM, MacKenzie G. The effect of hydrogen bonding,
Carbohydr Res. 2012;360:78–83. molecular shape, dipole moments, and chain length on
[31] Coats AM, Hukins DWL, Imrie CT, Aspden RM. the mesomorphism of some D-glucose and D-glucosa-
Polarization artefacts of an FTIR microscope and the mine derivatives. Liq Cryst. 1997;22(5):609–620.
consequences for intensity measurements on anisotro- [42] Ewing DF, Glew M, Goodby JW, Haley JA, Kelly
pic materials. J Microscopy-Oxford. 2003;211:63–66. SM, Komanschek BU, Letellier P, Mackenzie G,
[32] Ottenhof MA, MacNaughtan W, Farhat IA. FTIR Mehl GH. N-acyl-ß-D-glycopyranosylamines contain-
study of state and phase transitions of low moisture ing 1,4-disubstituted cyclohexyl and phenyl rings:
sucrose and lactose. Carbohydr Res. 2003;338 mesomorphism and molecular structure relationships.
(21):2195–2202. J Mater Chem. 1998;8(4):871–880.
[33] Cook AG, Inkster RT, Martinez-Felipe A, Ribes- [43] De Roche J, Gordon CM, Imrie CT, Ingram MD,
Greus A, Hamley IW, Imrie CT. Synthesis and phase Kennedy AR, Lo Celso F, Triolo A. Application of
behaviour of a homologous series of polymethacrylate- complementary experimental techniques to characteri-
based side-chain liquid crystal polymers. Eur Polym J. zation of the phase behavior of C(16)mim PF6 and C
2012;48(4):821–829. (14)mim PF6. Chem Mater. 2003;15(16):3089–3097.

Vous aimerez peut-être aussi