Sunteți pe pagina 1din 4

Society for Experimental Mechanics, 2002 SEM Annual Conference Proceedings, Milwaukee, WI, 2002.

Mechanical Properties of a pH Sensitive Hydrogel

æ § ¥ § æ æ
B. Johnson , D.J. Niedermaier , W.C. Crone , J. Moorthy and D.J. Beebe
æ ¥ §
Department of Biomedical Engineering, Department of Engineering Physics, Engineering Mechanics Program
University of Wisconsin-Madison, Madison, WI 53706, USA

ABSTRACT sensitive to the pH of their aqueous environment; expanding


at high pH and shrinking at low pH. The structures created
An experimental protocol was developed to determine the from this hydrogel for use in microfluidic devices must
basic mechanical properties of a poly(2-hydroxyethyl endure forces imposed by flow of an aqueous solution, the
methacrylate (HEMA)) hydrogel at various levels of pH constraint of device walls during swelling, and the restoring
equilibrium. An Instron micromechanical testing machine force of elastic membranes. Optimizing the design of these
was modified such that tensile tests could be carried out in devices demands an understanding of how the mechanical
temperature controlled pH solution. Values of Young’s properties of the gel change with its surroundings.
modulus were measured for a range of pH values from 2 to
12 and the experimental data was compared with theoretical To assist device designers, material properties must be
predictions of rubber elasticity. Although theory known in order to ensure functional and reliable structures.
overpredicted the decrease in modulus by 0.05 MPa, the Additionally, there is a strong desire to better understand the
global decrease in modulus with increased swelling was underlying mechanisms for swelling in hydrogels so that
experimentally confirmed. more predictive models can be developed to guide the
design process [10]. Although our research focuses on
1 INTRODUCTION HEMA based gels, the methods outlined are applicable to a
variety of polymeric gels and biomaterials.
There is a growing interest in developing engineered
actuation systems that have properties more in common with 2 BACKGROUND INFORMATION
soft biological materials, such as muscles and tendons, than
with traditional engineering materials. In an aqueous Theories describing the behavior of cross-linked polymer
environment, hydrogels will undergo a reversible phase networks were originally developed to understand the
transformation that results in dramatic volumetric swelling elasticity of vulcanized rubber. While early molecular
and shrinking upon exposure and removal of a stimulus. theories were restricted to uniaxial tension, statistical
Thus, these materials can be used as muscle-like actuators formulations successfully dealt with alternate states of strain
[1], fluid pumps [2], and valves [3]. Interest in hydrogels has (e.g. simple shear or uniaxial compression). Kuhn [11] was
gained momentum recently because these materials can be the first to relate the elastic modulus and the molecular
actuated by a variety of stimuli such as pH [4], salinity [5], weight of the chains of a polymer network, however his
electrical current [6], temperature [7], and antigens [8]. derivation did not account for large strains. Wall [12]
extended Kuhn’s work with a formulation for the elastic
Since the rate of swelling and shrinking in a hydrogel is modulus at any elongation:
diffusion-limited, the temporal response of macroscale
hydrogel structures to a given stimulus is typically measured  RTρ  2
on time scales of hours or days. Thus, by reducing the size ε =  1 + 3  , (1)
of hydrogel structures to the microscale, volumetric  M c  α 
transitions occur within minutes or even seconds. The
favorable scaling of hydrogel dynamics has been the where R is the universal gas constant, T is absolute
essential element in the development of micro-fluidic devices temperature, ρ is the density of the material, M c is the
that employ hydrogel valves for flow control [9]. One benefit
molecular weight of the polymer chain, and α is the ratio of
of these devices is that they are completely autonomous,
the deformed length to the initial length in the direction of the
and therefore require no external power source.
applied strain.
Our research considers a poly(2-hydroxyethyl methacrylate
Flory and Rehner [13, 14] recognized the swelling
(HEMA)) gel applied to the autonomous control of flow in
phenomenon exhibited by cross-linked polymers when
microfluidic devices [9]. HEMA based hydrogels are
exposed to certain solvents and derived an expression for with an optical microscope and image analysis software
the elastic modulus of a swollen polymer network (Metamorph 4.6r6, Universal Imaging Corp.) after each test
was completed. All strain measurements were calculated
from the initial, stress-free state the gel assumes after
 RTρν 21 3 
ε ' =  1 + 23  , (2) reaching equilibrium with its respective pH solution.
M  α  Therefore, only strains due to mechanical deformation, and
 c  not swelling, were considered in the determination of
mechanical properties.
where ν2 is the volume ratio of unswollen gel to swollen gel.
Thus, the modulus of elasticity for a swollen cross-linked The experimental data was divided into swollen and
polymer network can be related to an unswollen network by unswollen regions for the purpose of comparison to the
theoretical predictions. The transition pH of 5.0 was based
1 3. (3)
ε ' = εν 2 on swelling experiments in which the diameters of cylindrical
hydrogel structures were measured after reaching
This relation is compared below to experimental results equilibrium with their respective pH environment [10]. An
found for a HEMA gel. average value of 0.29 MPa was calculated from the first four
data points of Figure 1 for tension samples tested in the
3 EXPERIMENTS unswollen state. The modulus value at pH 5.5 was excluded
since it was within the region of phase transformation. The
average of the remaining data points was 0.21 MPa.
Tension samples were fabricated from a prepolymer solution
of HEMA gel in the liquid state. The liquid prepolymer was
injected into a polydimethylsiloxane (PDMS) mold that took
the form of dumbbell shaped tension samples scaled down
from the Type IV geometry outlined in ASTM D 638-99.
Ultraviolet (UV) light from a Novacure UV source was
2
delivered at an intensity of 15 mW/cm for two minutes to
achieve proper polymerization within the PDMS mold. The
resulting samples had a 5.5 mm gage length, 1 mm neck
width, and an approximate thickness of 200 µ m. The
samples were washed with methanol and deionzed water
and were stored in the appropriate pH solution for at least 24
hours prior to testing. The 0.2 M ionic strength pH solutions
used in this study were made from NaH2PO4, Na2HPO4, HCl,
NaOH, and NaCl.

An Instron Model 5548 MicroTester, equipped with a 10 N


load cell, was used for measuring the mechanical properties
of the hydrogel. The position of the screw-driven actuator
could be determined to within ±6.0 µm over 100 mm of
travel, while loads could be measured to within ±0.25% of
Figure 1: Young’s modulus as a function of pH. The solid
the indicated force. A temperature controlled environmental
horizontal lines indicate the averaged modulus values for the
chamber added to the MicroTester allowed the samples to
unswollen and swollen states. The shaded area signifies the
be tested in solution. The buoyant force exerted by the test
region of phase transformation for which data points were
solution on the grip fixtures was accounted for by conducting
discarded. The dashed lines indicate values calculated from
a test in which the sample was absent from the experimental
equation (3) based on an initial unswollen modulus of 0.29
setup. The resulting force-displacement curve was
MPa.
subtracted from the tests in which the sample was present.

4 RESULTS The volume of the gage section of the tensile samples were
measured in the absence of strain and the values were
Young’s modulus was determined from the initial portion of averaged amongst the swollen and unswollen groupings for
the stress/strain curve in tensile tests carried out to failure at the purpose of calculating ν 2 . Similarly, cylinder volumes
a constant crosshead displacement rate of 5 mm/min. A set were averaged from data provided by the aforementioned
of five samples was tested for each pH value investigated. A swelling experiments. The swelling that takes place in the
mean value of Young’s modulus was calculated from this swelling experiments is considered to be constrained in one
data set and is indicated with a data point in Figure 1, while dimension due to rigid upper and lower channel walls.
error bars denote the standard deviation within each data Therefore, a fixed channel height of 0.180 mm was assumed
set. in the volume calculation. The impact of this constraint on
the final swelling volume, and thus ν2, is outlined in Table 1.
Young’s modulus was calculated from the slope of the
stress-strain curve within the region of 10% strain, where Taking the average unswollen elastic modulus as 0.29 MPa,
axial strains were determined from crosshead position.
and ν2 as 0.19 for this particular gel chemistry, equation (3)
Cross-sectional areas for stress calculations were measured
yields a value of 0.16 MPa for the elastic modulus of the gel than what is revealed in the averaged experimental data
in its swollen state using the unconstrained volume data. (Figure 1). Therefore, the theory discussed above may not
Using values from the swelling experiments collected under adequately describe the influence of swelling on the elastic
constrained conditions yields a value of 0.18 MPa for the modulus.
elastic modulus.
Hydrogel swelling experiments show that the phase
transformation for this gel chemistry occurs around a pH
Table 1: Volume data for constrained and free swelling value of 5.0 [10]. They also reveal that cylindrical hydrogel
hydrogel, and the resulting volume ratios and moduli. structures tend to expand to approximately twice their
Avg. Unswollen Avg. Swollen original diameter (ν2 = 0.25) whereas samples used in the
Volume (mm )
2
Volume (mm )
2 ν2 ε'
tension tests swelled to a greater degree (ν 2 = 0.19). This
Constrained 1.45 × 10 −2
5.82 × 10 −2
0.25 0.18
may be attributed to the fact that the swelling experiments
Unconstrained 3.50 18.8 0.19 0.16
were conducted within a micro-channel where upper and
lower walls constrained expansion of the hydrogel structure
in one dimension. Conversely, the tension samples used in
An additional study was undertaken to investigate the affect this study were allowed to expand freely in bulk solution prior
of polymerization time on the elastic modulus. to testing. The dependence of swelling on an applied stress
Representative data is shown in Figure 2 for samples that or strain was confirmed experimentally by Treloar [15]. His
were exposed to UV light for durations varying from one to experiments on natural rubber showed that an applied
ten minutes. For simplicity, tests were performed in uniaxial tensile stress produced an increase in swelling,
deionized water and sample cross sections were slightly while a uniaxial compressive stress caused a reduction in
larger (1.5 mm by 0.5 mm) than those used in the pH study. swelling.

In addition to the above discussion, the results shown in


Figure 1 provide essential information for the efficient design
of hydrogel structures in microfluidic devices. Young’s
modulus data has been implemented in mathematical
models that are able to predict the swelling characteristics of
the gel [10]. Specifically, radial displacements in a
cylindrical gel can be related to material constants (Young’s
modulus and Poisson’s ratio) and the osmotic pressure that
arises from the inflow of ions that maintain charge neutrality
in the presence of basic buffer solutions.

As demonstrated by the results presented in Figure 2, the


modulus of a hydrogel is also affected by the polymerization
process. Photo Differential Scanning Calorimetry (Photo-
DSC) experiments on a HEMA gel of similar composition
showed that there is a peak time during photopolymerization
when the reaction rate is maximum and then stalls to zero
upon further UV exposure [16]. As the reaction rate
Figure 2: Young’s modulus as a function of polymerization increases, an increase in the conversion of monomer to
time. Data points indicate the average modulus of four polymer is expected. This would also imply an increase in
samples while error bars denote the standard deviation cross-linking and thus modulus. Such a trend is represented
within each sample set (the data set corresponding to a 6 in Figure 2 for values of Young’s modulus corresponding to
minute polymerization time contains three samples). polymerization times of one, two, and four minutes. The
modulus appears to reach a maximum at a polymerization
5 DISCUSSION time of four minutes, after which it begins to decay. Photo-
DSC experiments reported a peak in the reaction rate
between two and four minutes of UV exposure [16]. The
The effectiveness of the theory to predict changes in
location of this peak may be influenced by a number of
modulus with swelling, and the implications of mechanical
parameters including the UV source, type and concentration
constraint on swelling will be discussed. The impact of
of photo-initiator, and presence of other monomers.
polymerization conditions on Young’s modulus will also be
addressed.
After peak cross-linking is achieved, an apparent decay in
Young’s modulus is observed for polymerization times of six
Although Figure 1 does not display a step wise change in
and ten minutes. Because the conversion of monomer to
Young’s modulus at the pH where transformation occurs, the
polymer should cease once the reaction stalls, the modulus
data above and below phase transition was averaged such
would be expected to maintain its peak value upon
that a constant value of modulus was assumed for each of
prolonged exposure to UV. However, in the presence of UV,
the swollen and unswollen states. This approach was
oxygen forms ozone that can subsequently form oxygen
adopted primarily for the purpose of comparing experimental
radicals. The oxygen radicals have the ability to break
results with theoretical predictions. The theoretical
bonds, thereby causing a decrease in the modulus.
formulation tends to predict a greater decrease in modulus
9. Beebe, D.J., et al., Functional hydrogel structures for
6 CONCLUSIONS autonomous flow control inside microfluidic channels.
Nature, 404(6778), pp. 588-590, 2000.
Experiments were conducted on a pH sensitive hydrogel to 10. De, S.K., et al., Equilibrium Swelling and Kinetics of pH-
determine the dependence of Young’s modulus on two responsive Hydrogels: Models, Experiments, and
parameters: the environmental pH and the time of exposure Simulations. submitted for publication to the Journal of
to UV light during polymerization. Theoretical predictions Microelectromechanical Systems.
were in agreement with experiments in reference to the 11. Kuhn, W., Kolloid-Zeitschrift, 76, pp. 258-271, 1936.
global decrease in modulus with increased swelling. 12. Wall, F.T., Statistical Thermodynamics of Rubber. II.
However the magnitude of this decrease was slightly The Journal of Chemical Physics, 10, pp. 485-488,
overpredicted. 1942.
13. Flory, P.J. and J. Rehner Jr., Statistical Mechanics of
A brief investigation of the impact of mechanical constraint Cross-Linked Polymer Networks: I. Rubberlike Elasticity.
on the ratio of unswollen to swollen volume (ν 2) was also The Journal of Chemical Physics, 11(11), pp. 512-520,
carried out. This study shows the need for further 1943.
investigation with experiments designed to test the change in 14. Flory, P.J. and J. Rehner Jr., Statistical Mechanics of
ν2 with varying levels of applied strain. This information may Cross-Linked Polymer Networks: II. Swelling. The
be particularly useful in the design of improved microfluidic Journal of Chemical Physics, 11(11), pp. 521-526, 1943.
devices that utilize the relationship between swelling and 15. Treloar, L.R.G., The Swelling of Cross-Linked
strain. Amorphous Polymers Under Strain. Transactions of the
Faraday Society, 46, pp. 783-789, 1950.
Additionally, it was experimentally demonstrated that the 16. Lai, Y.C. and E.T. Quinn, The effects of initiator and
duration of UV exposure during photopolymerization diluent on the photopolymerization of 2-hydroxyethyl
influences the mechanical properties of this hydrogel. methacrylate and on properties of hydrogels obtained.
Accounting for differences in polymerization conditions and Photopolymerization, 673, pp. 35-50, 1997.
chemical composition, the peak in the modulus coincides
reasonably well with reaction rate peaks measured with
photo-DSC experiments.

ACKNOWLEDGEMENTS

This research is sponsored by DARPA, AFRL, Air Force


Command, and USAF under agreement F30602-00-1-0570.
The authors would also like to thank Narayana Aluru, Joseph
Bauer, David Eddington, and Glennys Mensing for helpful
discussions and technical assistance.

REFERENCES

1. Shahinpoor, M., Microelectromechanics of Ionic


Polymeric Gels as Electrically Controllable Artificial
Muscles. Journal of Intelligent Material Systems and
Structures, 6(3), pp. 307-314, 1995.
2. Seigel, R.A., Implantable, Self-Regulating
Mechanochemical Insulin Pump. 1991, The Regents of
the University of California, Berkeley, CA: USA.
3. Osada, Y. and S.B. Rossmurphy, Intelligent Gels.
Scientific American, 268(5), pp. 82-87, 1993.
4. Tanaka, T., et al., Phase Transitions in Ionic Gels.
Physical Review Letters, 45(20), pp. 1636-1639, 1980.
5. Grodzinsky, A.J. and P.E. Grimshaw, Electrically and
Chemically controlled Hydrogels for Drug delivery.
Pulsed and Self-Regulated Drug Delivery, pp. 47-64,
1990.
6. Tanaka, T., et al., Collapse of Gels in an Electric-Field.
Science, 218(4571), pp. 467-469, 1982.
7. Hu, Z.B., X.M. Zhang, and Y. Li, Synthesis and
Application of Modulated Polymer Gels. Science,
269(5223), pp. 525-527, 1995.
8. Miyata, T., N. Asami, and T. Uragami, A reversibly
antigen-responsive hydrogel. Nature, 399(6738), pp.
766-769, 1999.

S-ar putea să vă placă și