Sie sind auf Seite 1von 4

LABORATORY EXPERIMENT

pubs.acs.org/jchemeduc

Synthesis and Characterization of Self-Assembled Liquid


Crystals: p-Alkoxybenzoic Acids
Jana Jensen, Stephan C. Grundy, Stacey Lowery Bretz, and C. Scott Hartley*
Department of Chemistry and Biochemistry, Miami University, Oxford, Ohio 45056, United States

bS Supporting Information
ABSTRACT: Thermotropic liquid crystal phases are ordered uids found, for
some molecules, at intermediate temperatures between the crystal and liquid
states. Although technologically important, these materials typically receive little
attention in the undergraduate curriculum. Here, we describe a laboratory
activity for introductory organic chemistry students on the synthesis and
characterization of the p-alkoxybenzoic acids. These compounds, through the
formation of carboxylic acid dimers, exhibit liquid crystal phases common in rodlike (calamitic) molecules. The students are assigned dierent alkoxy chain
lengths and synthesize the compounds through microwave-assisted nucleophilic
substitution. Characterization of the phase behavior is then carried out by
standard melting point techniques, dierential scanning calorimetry, or polarized optical microscopy. The results for the class are pooled to allow the students
to consider structure property eects for the series. This activity allows
students to explore small-molecule synthesis applied to materials chemistry and concepts of self-assembly: the benzoic acids
associate through hydrogen bonding, and the resulting rod-like dimers further organize into the liquid crystal phases.
KEYWORDS: Second-Year Undergraduate, Laboratory Instruction, Organic Chemistry, Hands-On Learning/Manipulatives,
Calorimetry/Thermochemistry, IR Spectroscopy, Materials Science, NMR Spectroscopy, Nucleophilic Substitution, Phases/Phase
Transitions/Diagrams

hermotropic liquid crystal phases are condensed phases


exhibited by some compounds at intermediate temperatures
between the liquid and crystalline solid.1,2 Similar to conventional isotropic liquids, the molecules in a liquid crystal phase
are mobile (i.e., diuse). However, similar to crystals, liquid
crystalline materials exhibit properties arising from long-range
molecular order. As a consequence of this combination of
molecular mobility and crystal-like properties, liquid crystals
are used in a number of technological applications, most
prominently as the active components of the at panel video
displays of televisions, computers, and other devices. Liquid
crystals are also used in other interesting applications, for
example, as ultrahigh-strength materials including Kevlar and
natural silk.1,3 Despite their signicance to both current technologies and fundamental science, liquid crystals are typically
omitted from the undergraduate curriculum.4,5 However, they
provide a straightforward introduction to organic materials
chemistry and self-assembly and are readily introduced as an
activity at the introductory organic chemistry level. Here, we
present a laboratory activity based on the synthesis and characterization of a series of p-alkoxybenzoic acids. The goal of this
activity, which can be completed in a single laboratory session, is
to introduce the students to both liquid crystals as a fundamental
state of matter and the concepts of self-assembly in materials
science.
The oldest class of liquid crystal-forming molecules, the
calamitics, are dened by their rod-like shape.6 The simplest
Copyright r 2011 American Chemical Society and
Division of Chemical Education, Inc.

liquid crystal phase exhibited by the calamitics is the nematic


phase, which is characterized by long-range orientational order:
on the time-average, the molecular long axes are aligned in a
common direction dened by the director (^n), as shown in
Figure 1. In the smectic liquid crystal phases, the molecules are
further oriented into layers. Shown in Figure 2, the simplest
smectic phases are the smectic A, in which the molecules are
oriented along the layer normal (^z), and the smectic C, in which
they are tilted from the layer normal. The p-alkoxybenzoic acids
synthesized in this experiment exhibit both the nematic and
smectic C phases, depending on their structure.
Liquid crystals have been used in physics and physical
chemistry laboratory courses to demonstrate physical properties,
such as order parameters, electrooptic eects, and optical properties (e.g., birefringence, refractive index anisotropy, selective
reection).5,7 12 They have also been used to study other
applications such as the orientation of dye molecules in liquid
crystalline solvents7,13,14 and thermochromic thermal mapping
(i.e., mood rings).15,16 Laboratory experiments have also
featured lyotropic (solvent-dependent) liquid crystals.17 The
synthesis of liquid crystals provides a fundamental opportunity
for students to gain hands-on experience with organic materials
chemistry. However, there are relatively few published experiments that allow students to explore the synthesis of liquid
Published: June 03, 2011
1133

dx.doi.org/10.1021/ed101090t | J. Chem. Educ. 2011, 88, 11331136

Journal of Chemical Education

LABORATORY EXPERIMENT

Scheme 1. Synthesis of Liquid Crystals from p-Alkoxybenzoic Acid

Figure 1. The nematic phase of calamitic liquid crystals: (left) a uid in


which the molecules point in the same direction (^n); (right) in the
context of this experiment, the rod-like molecules are hydrogenbonded p-alkoxybenzoic acid dimers.

Figure 2. The (A) smectic A and (B) smectic C phases of a calamitic


liquid crystal.

crystalline materials. Previous reports have included the preparation of cholesteryl benzoates, the rst liquid crystals
discovered,6,18,19 and N-(p-methoxybenzylidene)-p-n-butylaniline (MBBA), the rst liquid crystal used in displays.4 Although
both of these activities employ synthesis of a single liquid
crystalline compound, our experiment is unique in two ways:
(i) the students synthesize a series of compounds, allowing them
to pool data and explore structure property relationships, and
(ii) it emphasizes the concept of self-assembly in organic
materials.
Jones rst discovered liquid crystal phases in the p-alkoxybenzoic acids in 1929.20 23 In developing a teaching activity, we were
attracted to these molecules for three reasons. First, they can be
readily synthesized in a single step from inexpensive, commercially available starting materials using straightforward chemistry
(nucleophilic substitution), as shown in Scheme 1. Consequently, this experiment could be used early in the organic
laboratory curriculum, as both SN2 substitution and intermolecular forces have usually been discussed at the midpoint of the
rst semester. Second, these molecules provide an ideal introduction to liquid crystals as a form of self-assembly.24 The liquid
crystal phases formed by p-alkoxybenzoic acids incorporate two
levels of hierarchical self-assembly. The benzoic acids dimerize
through hydrogen bonding to give aggregates with a more
pronounced rod-like shape; these rod-like dimers then further
assemble into the liquid crystal phases (Figure 1). Third, the
students can synthesize targets with dierent alkoxy chain
lengths (1 6). Variation of side-chain lengths on the same core
structure is very common in liquid crystals science as a means to
optimize phase behavior. The class results can then be pooled to
explore the eect of molecular structure on bulk properties, such
as phase transitions.

EXPERIMENTAL DETAILS
The syntheses were conducted in the organic chemistry
laboratory with second-year chemistry majors. Eleven students
completed the experiment successfully. We adapted the synthesis
to make use of a microwave reactor to shorten the duration of the
lab; if a microwave is not available, the reaction can be performed
at reux as in the original synthesis.21 Students worked in groups
of two or three to save time using the microwave. In our class,
each group was assigned either pentyl, heptyl, or nonyl bromide.
In a typical procedure, p-hydroxybenzoic acid (10 mmol) was
dissolved in 1 M solution of KOH in methanol (21 mL). The
bromoalkane (11 mmol) was added using a syringe and the
mixture was swirled until well mixed. After microwave irradiation
(20 min at 50% power using a 630 W CEM MDS 2000
Microwave Digestion System), the product was transferred to
a ask using a small volume of methanol and acidied with 1 M
HCl(aq) (100 mL). The resulting precipitate was isolated by
vacuum ltration, washing with methanol (100 mL); if a signicant quantity of additional product precipitated in the ltrate,
it was isolated as well. To remove unreacted starting material, the
total product was dissolved in dichloromethane, ltered, and
concentrated under reduced pressure. The crude products were
then recrystallized from minimum volumes of hexanes. Student
yields, after recrystallization, ranged from 13% to 34%. The
products were characterized using NMR and IR spectroscopy,
aording spectra typical of these compounds (see the Supporting
Information). It is noteworthy that IR spectroscopy shows a
broad band in the neighborhood of 3000 2500 cm 1, a
characteristic signature of the hydrogen-bonded carboxylic acid
dimer.25
Characterization of the liquid crystal phases can be done with a
variety of techniques, ranging from simple observation with a
standard melting point apparatus to dierential scanning calorimetry (DSC) and polarized optical microscopy. This experiment does not require sophisticated equipment, as the
transitions to and from the liquid crystal phases are readily
discernible using a standard capillary melting point apparatus:21
they appear as opaque, opalescent uids. On heating, it was
possible to observe the more subtle smectic C to nematic
transition, but this can be challenging without prompting of
the expected transition temperatures. Students using the capillary
melting point apparatus usually reported phase transitions below
those reported by Gray and Jones.23 Our class also used DSC to
measure phase-transition temperatures and enthalpies. The
smectic C to nematic phase transitions were readily discernible
by DSC. In general, the phase transitions of the student products
were within 4 C of the literature values. As is common for liquid
crystals, the smectic C to nematic and nematic to isotropic liquid
transitions were characterized by considerably smaller enthalpies
of transition (ca. 5 J/g) than the crystal to liquid-crystal transitions (ca. 50 J/g), reecting the structural similarity of the liquid
crystal and isotropic liquid phases. A complete list of phase
1134

dx.doi.org/10.1021/ed101090t |J. Chem. Educ. 2011, 88, 11331136

Journal of Chemical Education

LABORATORY EXPERIMENT

Figure 3. Polarized optical micrographs of the schlieren textures of (A) the nematic and (B) the smectic C phases of student-synthesized pheptyloxybenzoic acid. Also shown is the crystal phase (C). Note that the spots interrupting the texture likely represent small quantities of impurities in
the sample, possibly 4-hydroxybenzoic acid.

transitions and enthalpies is included in the Supporting Information. As groups of students had been assigned dierent alkyl
chain lengths, the results were pooled to illustrate structure
property eects in this system. In very broad terms, increasing
the alkyl chain length decreases the clearing point (nematic to
isotropic transition temperature), reecting the less ecient
nematic packing as the conformational exibility of the chains
increases. The other obvious trend was promotion of the smectic
C phase at long chain lengths, which may be ascribed to increased
microphase segregation between the rigid benzoic acid dimer
cores and the exible side chains.
One of the most attractive aspects of liquid crystals research is
that, similar to crystals, liquid crystals are birefringent (i.e., liquid
crystals, as anisotropic materials, have dierent refractive indices
depending on the polarization of incident light). Consequently,
thin lms of liquid crystals give highly colored textures when
viewed through crossed polarizers, whereas isotropic phases
appear black. Although not required for this experiment, the
construction of polarizing microscopes out of easily accessible
items has been previously described: regular microscopes can be
used along with polarizing lm,4 a hot air gun can be used to
control temperature of the materials, or a convenient variable
temperature stage can be constructed as described by Verbit and
Halbert.26 In the case of the p-alkoxybenzoic acids, the nematic
and smectic C phases displayed characteristic so-called schlieren
textures, as shown in Figure 3 (also shown, for comparison, is the
crystal phase). Although a detailed discussion of the molecular
origins of liquid crystal textures is beyond the scope of this
activity, we point out that the wavy black lines of the schlieren
texture correspond to regions in which the long axes of the
molecules are lined up parallel or perpendicular to the polarizers
of the microscope. Thus, the textures are a direct reection of the
bulk molecular order of a liquid crystalline material.

HAZARDS
The 1-bromoalkanes are considered irritants and may be
harmful if ingested or inhaled. p-Hydroxybenzoic acid is also
an irritant. Sodium hydroxide is caustic and hydrochloric acid is
corrosive. Methanol is highly ammable and toxic. Hexane is
dangerous for the environment, harmful, and highly ammable.
All chemicals in the experiment should be handled with goggles
and protective gloves. The synthesis is best done in a fume hood,
and waste should be disposed of according to standard procedures.
STUDENT DISCUSSIONS
Five discussion questions and answers are provided in the
instructors manual included in the Supporting Information. The
students answers indicated comprehension of most of the liquid

crystal topics introduced. They were able to note that liquid


crystal to isotropic transition enthalpies are typically smaller than
crystal to liquid crystal enthalpies, including a discussion of the
relationship between transition temperatures and structures. The
students were also able to distinguish the dierent liquid crystal
phase textures produced by a polarizing microscope. Some
students demonstrated a thoughtful understanding of the alkylation reaction, explaining that the reaction occurs with high
chemoselectivity due to the higher nucleophilicity of the more
basic phenoxide group; one such student explained that the
reaction only happens at hydroxy [sic] group because a deprotonated hydroxy is much more reactive than COOH due to the
electron withdrawing properties of the carbonyl group. A few
students misunderstood the use of hydrochloric acid and thought
it was used to remove excess starting material. Also, whereas most
students grasped the relationship between structure and trends in
phase transition temperatures, a few students misinterpreted how
hydrogen bonding aected these trends. For example, students
successfully noted that the initial melting points are higher for 1
and 2. However, some students incorrectly reasoned that this
occurred because shorter chains are more apt to form H-bonds
and at 7 8, the steric factors make hydrogen bonds less eective.

CONCLUSION
The microwave-assisted synthesis and characterization of palkoxybenzoic acids is well suited for the introduction of organic
materials chemistry in the undergraduate laboratory. The target
compounds are readily prepared by alkylation of p-hydroxybenzoic acid with microwave irradiation. Observation with a standard melting point apparatus, as well as dierential scanning
calorimetry and polarized optical microscopy, can be used to
study the phase behavior. Collaborative eorts by student groups
permits pooling of data to observe structure property relationships: the eect of varying the molecular structure on the phase
behavior of a series of compounds.
ASSOCIATED CONTENT

bS

Supporting Information
Student handouts; instructor notes, including phase transitions and enthalpies; CAS registry numbers of chemicals;
manufacturers of equipment; NMR and IR spectra; and polarizing microscope photos. This material is available via the Internet
at http://pubs.acs.org.

AUTHOR INFORMATION
Corresponding Author

*E-mail: scott.hartley@muohio.edu.
1135

dx.doi.org/10.1021/ed101090t |J. Chem. Educ. 2011, 88, 11331136

Journal of Chemical Education

LABORATORY EXPERIMENT

ACKNOWLEDGMENT
The authors wish to thank the National Science Foundation
(award #0733642) and Miami University for nancial support.
We are grateful to the students and Ben Gung. We also thank
Jason Crase for aid in optimizing the procedures used.
REFERENCES
(1) Brown, G. H. J. Chem. Educ. 1983, 60, 900905.
(2) Collings, P. J.; Hird, M. Introduction to Liquid Crystals; Taylor &
Francis Ltd.: London, 1997.
(3) Kerkam, K.; Viney, C.; Kaplan, D.; Lombardi, S. Nature 1991,
349, 596598.
(4) Liberko, C. A.; Shearer, J. J. Chem. Educ. 2000, 77, 12041205.
(5) Waclawik, E. R.; Ford, M. J.; Hale, P. S.; Shapter, J. G.; Voelcker,
N. H. J. Chem. Educ. 2004, 81, 854858.
(6) Reinitzer, F. Monatsh. Chem. 1888, 9, 421441.
(7) Demirbas, E.; Devonshire, R. J. Chem. Educ. 1996, 73, 586589.
(8) DuPre, D. B.; Chapoy, L. L. J. Chem. Educ. 1979, 56, 759761.
(9) Lalanne, J. R.; Hare, F. J. Chem. Educ. 1976, 53, 793795.
(10) Jeppesen, M. A.; Hughes, W. T. Am. J. Phys. 1970, 38, 199201.
(11) Olah, A.; Doane, J. W. Am. J. Phys. 1977, 45, 485488.
(12) Van Hecke, G. R.; Karukstis, K. K.; Li, H. H.; Hendargo, H. C.;
Cosand, A. J.; Fox, M. M. J. Chem. Educ. 2005, 82, 13491354.
(13) Sadlej-Sosnowska, N. J. Chem. Educ. 1980, 57, 223224.
(14) Wilson, R. M.; Gardner, E. J.; Squire, R. H. J. Chem. Educ. 1973,
50, 9498.
(15) Fergason, J. L. Am. J. Phys. 1970, 38, 425428.
(16) Lisensky, G. C.; Horoszewski, D.; Gentry, K.; Zenner, G. M.;
Crone, W. C. Sci. Teach. 2006, 73, 3035.
(17) Friberg, S. E.; Bendiksen, B. J. Chem. Educ. 1979, 56, 553555.
(18) Patch, G.; Hope, G. A. J. Chem. Educ. 1985, 62, 454455.
(19) Verbit, L. J. Chem. Educ. 1972, 49, 3639.
(20) Bradeld, A. E.; Jones, B. J. Chem. Soc. 1929, 26602661.
(21) Jones, B. J. Chem. Soc. 1935, 1874.
(22) Bennett, G. M.; Jones, B. J. Chem. Soc. 1939, 420425.
(23) Gray, G. W.; Jones, B. J. Chem. Soc. 1953, 41794180.
(24) Kato, T.; Mizoshita, N.; Kishimoto, K. Angew. Chem., Int. Ed.
2006, 45, 3868.
(25) Silverstein, R. M.; Webster, F. X.; Kiemle, D. J. Spectrometric
Identication of Organic Compounds, 7th ed.; John Wiley & Sons, Inc.:
Hoboken, NJ, 2005.
(26) Verbit, L.; Halbert, T. R. J. Chem. Educ. 1971, 48, 773774.

1136

dx.doi.org/10.1021/ed101090t |J. Chem. Educ. 2011, 88, 11331136

Das könnte Ihnen auch gefallen