PLLA: PCL Woodard2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Macromolecular

Communication Rapid Communications

PCL–PLLA Semi-IPN Shape Memory Polymers


(SMPs): Degradation and Mechanical
Properties
Lindsay N. Woodard, Vanessa M. Page, Kevin T. Kmetz,
Melissa A. Grunlan*

Thermoresponsive shape memory polymers (SMPs) based on poly(ε-caprolactone) (PCL)


whose shape may be actuated by a transition temperature (Ttrans) have shown utility for a
variety of biomedical applications. Important to their utility
is the ability to modulate mechanical and degradation proper-
ties. Thus, in this work, SMPs are formed as semi-interpene-
trating networks (semi-IPNs) comprised of a cross-linked PCL
diacrylate (PCL-DA) network and thermoplastic poly(l-lactic
acid) (PLLA). The semi-IPN uniquely allows for requisite
crystallization of both PCL and PLLA. The influence of PLLA
(PCL:PLLA wt% ratio) and PCL-DA molecular weight (n) on
film properties are investigated. PCL–PLLA semi-IPNs are able
to achieve enhanced mechanical properties and accelerated
rates of degradation.

1. Introduction well as able to recover their original shape upon heating


(T > Ttrans). SMPs have been investigated for numerous bio-
Thermoresponsive shape memory polymers (SMPs) encom- medical applications, including minimally-invasive cardio-
pass a class of materials capable of shape change via vascular devices and custom-fitted orthopedic implants,
regions within the polymer (i.e., amorphous or crystalline with SMPs having a sharp shape transition (melt transition
domains), referred to as “switching segments,” actuated temperature, Tm = Ttrans), robust mechanical properties and
by a transition temperature (Ttrans). Switching segments the capability of biodegradation of particular interest.[2,3]
are responsible for molecular movement (T > Ttrans) or fixa- Poly(ε-caprolactone) (PCL) has been commonly investi-
tion (T < Ttrans), while “netpoints” (i.e., chemical or physical gated for SMPs due to its biodegradability and a Tm (Ttrans)
cross-links) define the permanent shape.[1] Thus, SMPs are which is tunable (43–60 °C, based on Mn) and also toler-
capable of holding a temporary shape following sequential ated by tissues.[4–6] Chemically cross-linked SMP networks
heating (T > Ttrans), shaping and cooling (T < Ttrans), as formed from PCL dimethacrylates have been previously
reported,[5] and our group recently reported solvent-casted
L. N. Woodard, V. M. Page, K. T. Kmetz, Prof. M. A. Grunlan PCL diacrylate (PCL-DA) SMPs.[7,8] While this PCL-DA SMP
Department of Biomedical Engineering
displayed excellent shape memory, modification with a
Texas A&M University
second polymer component would favorably broaden its
College Station, TX 77843, USA
E-mail: mgrunlan@tamu.edu
mechanical and degradation properties. PCL-based SMPs
Prof. M. A. Grunlan have been modified with other copolymer segments,
Department of Materials Science and Engineering including polyurethanes,[9] poly(ethylene glycol),[10] and
Texas A&M University poly(n-butyl acrylate).[11] However, poly(l-lactic acid)
College Station, TX 77843, USA (PLLA) represents a potentially useful option due to its

Macromol. Rapid Commun. 2016, DOI: 10.1002/marc.201600414


© 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim wileyonlinelibrary.com DOI: 10.1002/marc.201600414 1

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Macromolecular
Rapid Communications L. N. Woodard et al.

www.mrc-journal.de

biodegradability as well as high modulus (E ∼ 2.7 GPa) 2.3. Fabrication


stemming from its semi-crystallinity and relatively high Solutions (25 wt% CH2Cl2) of PCL:PLLA (100:0 [PCL-DA control],
Tg (≈60 °C).[12] While SMPs comprised of PCL and PLLA 90:10, 75:25, 60:40 wt% ratio) were prepared with 15 vol% of a
have been reported, they have often been formed as copo­ photoinitiator solution (10 wt% DMP in NVP). The solution was
lymers, including random copolymers (such that the Ttrans transferred to a circular silicone mold (45 mm x 2 mm; McMaster-
is the PLLA Tg rather than the PCL Tm) or as cross-linked Carr) secured between glass slides and exposed to UV light (UV-
PCL–PLLA copolymer single networks.[13,14] In contrast, we Transilluminator, 6 mW cm−2, 365 nm) for 3 min per side. The
report herein PCL–PLLA SMPs formed as semi-interpene- solvent-swollen disk was sequentially removed from the mold,
trating networks (semi-IPNs) based on the combination of air dried (RT, 12 h), dried in vacuo (36 in.Hg, RT, 4 h), soaked
a PCL-DA network with thermoplastic PLLA. Importantly, in ethanol (3 h) and air dried. The disk was finally annealed
(36 in.Hg, 85 °C, 1 h) and allowed to set at RT (48 h) before testing.
since cross-linking is known to reduce chain mobility, and
hence crystallization,[15] incorporation of non-crosslinked
PLLA into the PCL-DA network was expected to yield the 2.4. SMP Characterization
desired semi-crystallinity needed to maintain shape
2.4.1. Semi-IPN Formation
memory capability while improving mechanical prop-
erties. It was further anticipated that combination as a See Supporting Information.
semi-IPN would provide the ability to tune the degrada-
tion rate of a PCL-based SMPs. 2.4.2. Thermal Properties
Herein, PCL–PLLA SMP semi-IPNs were prepared with
Differential scanning calorimetry (DSC, TA Instruments Q100) of
varying PCL:PLLA wt% ratios (100:0 [PCL-DA control],
specimens (≈12 mg; N = 3) in hermetic pans was ran from RT to
90:10, 75:25, 60:40) and with varied PCL-DA molecular
200 °C at a heating rate of 5 °C min−1. From the endothermic PCL
weight (n = 25, 45). Following UV cross-linking, the and PLLA melting peaks, melting temperature (Tm) and enthalpy
effects on thermal properties (i.e., crystallinity and change (ΔHm) were measured. Percent crystallinity (% χc) was cal-
Tm), shape memory behavior, mechanical properties culated via Equation (1)
(i.e., stiffness and strength), and degradation rate were
∆H m
studied. % χc = × 100, (1)
∆H m0

where ΔHm was calculated by the area of the melting peak,
and ∆H m0 is the enthalpy of fusion of 100% crystalline PCL
2. Experimental Section
(139.5 J g−1)[16] or PLLA (93.0 J g−1).[17]

2.1. Materials
2.4.3. Shape Memory Behavior
Polycaprolactone diol (PCL90-diol; Mn ∼ 10 000 g mol−1),
ε-caprolactone, l-lactide, stannous 2-ethylhexanoate, triethyl- Rectangular specimens (≈20 mm × ≈3.3 mm × ≈1.1 mm) were
amine (Et3N), acryloyl chloride, 4-dimethylaminopyridine subjected to the following sequence: (1) after exposure to a
(DMAP), 2,2-dimethoxy-2-phenylacetophenone (DMP), water bath at 80 °C (Thigh) for 1 min, deform the film strip into
1-vinyl-2-pyrrolidinone (NVP), potassium carbonate (K2CO3), a coiled shape with a metal mandrel, (2) place the film in an ice
sodium hydroxide (NaOH), ethylene glycol, and solvents were bath (≈0 °C, Tlow) for 1 min to fix this temporary shape, (3) place
obtained from Sigma-Aldrich. Anhydrous magnesium sulfate the fixed coil in the 80 °C water bath and observe recovery at
(MgSO4) was obtained from Fisher. Reagent-grade CH2Cl2 and t = 0 and 8 s.
NMR-grade CDCl3 were dried over 4 Å molecular sieves prior
to use. 2.4.4. Mechanical Properties
Tensile properties were evaluated at RT with a ten-
2.2. Material Synthesis sile tester (Instron 3340) by subjecting rectangular strips
(≈15 mm × ≈3.3 mm × ≈1.1 mm; N = 5–10) with a gauge length
PCL2n-diol and PLLA2m-diol (m = 90) were synthesized by the of ≈3 mm to a constant strain rate (50 mm min−1) in tension until
ring-opening polymerization of ε-caprolactone or l-lactide, break. From the resulting stress–strain curves, modulus (E), ten-
respectively, in the presence of ethylene glycol initiator and sile strength (TS), and strain at break (% ε) were determined.
stannous 2-ethylhexanoate catalyst. The degree of poly-
merization (n, m) was controlled by the ratio of monomer
2.4.5. Degradation
to ethylene glycol. The resulting terminal hydroxyl groups
of PCL2n-diol were subsequently converted to photosensi- Disks (≈7 mm × ≈1.1 mm; N = 3) were immersed in 20 mL of
tive acrylate (OAc) groups by reacting with acryloyl chloride. 1 m NaOH in a centrifuge tube maintained at 37 °C via a water
The number average molecular weights (Mn) and degree bath. Every 8 h for 72 h, the samples were taken from solution,
of acrylation were determined by 1H NMR (see Supporting thoroughly rinsed with DI water, blotted, and dried in vacuo
Information). (36 in.Hg, RT, ≈16 h). The mass and thickness of the samples were

Macromol. Rapid Commun. 2016,  DOI: 10.1002/marc.201600414


2 © 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim www.MaterialsViews.com

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Macromolecular
PCL–PLLA Semi-IPN Shape Memory Polymers (SMPs): Degradation and Mechanical Properties Rapid Communications
www.mrc-journal.de

Table 1.  Thermal and mechanical properties of SMP networks.

PCLn:PLLA ratio PCL Tm Crystallinity PLLA Tm Crystallinity Tensile mod- Tensile Strain at
[°C] [%] [°C] [%] ulus, E [MPa] strength, TS break, ε
[MPa] [%]
PCL25-DA control 48.8 ± 0.3 34.6 ± 1.1 N/A N/A 75.8 ± 7.7 18.4 ± 1.9 707 ± 150
9025:10 49.2 ± 0.3 30.9 ± 2.2 146.0 ± 0.8 5.4 ± 1.2 83.9 ± 11.6 19.0 ± 2.2 752 ± 140
7525:25 49.1 ± 0.5 25.4 ± 1.5 157.9 ± 0.7 15.4 ± 0.9 93.3 ± 9.3 18.8 ± 1.1 593 ± 190
6025:40 49.7 ± 0.3 21.2 ± 0.5 157.6 ± 0.6 28.9 ± 0.6 N/A a)
N/A a)
N/Aa)
PCL45-DA control 54.5 ± 0.2 43.5 ± 0.4 N/A N/A 67.3 ± 6.4 23.4 ± 2.1 1350 ± 99
9045:10 53.2 ± 0.3 39.8 ± 1.7 146.7 ± 0.5 5.3 ± 1.6 71.9 ± 9.8 24.5 ± 0.9 1350 ± 240
7545:25 52.8 ± 0.1 31.0 ± 0.3 152.7 ± 0.3 16.7 ± 0.5 77.3 ± 4.7 23.4 ± 1.9 1230 ± 160
6045:40 52.7 ± 0.3 26.0 ± 0.8 152.7 ± 0.3 25.8 ± 0.6 63.07 ± 8.3 16.5 ± 1.2 167 ± 200
a)Sample was too brittle for testing.

determined prior to returning to fresh NaOH solution. At 48 h, a


sample was imaged via scanning electron microscopy (SEM). Sur-
faces and cross sections were subjected to Au–Pt coating (≈4 nm).
Images were obtained using a JEOL 6400 SEM with an acceler-
ating voltage of 10 kV.

2.4.6. Statistical Analysis


Data was reported as the mean ± standard deviation. Values
were compared using a Student’s t-test to determine p-values.

3. Results and Discussion

To verify the semi-IPN PCL:PLLA wt% ratio of PCL to PLLA,


thermogravimetric analysis was utilized (see Figure S4,
Supporting Information). Because PLLA underwent deg-
radation at a distinguishably lower temperature than
the PCL-DA network, the wt% of each of PCL and PLLA in
the semi-IPN could be determined and was confirmed
to be similar to the solution wt% ratios used in semi-IPN
preparation. Moreover, sol content confirmed that the
PCL-DA network was effectively cross-linked, as only ther-
moplastic, non-crosslinked PLLA could be extracted (see
Table S1, Supporting Information). Hence, due to the high
degree of acrylation (≈98%–100%; see Supporting Infor-
mation), PCL-DA was able to successfully cross-link in the
presence of PLLA.
Tm and % crystallinity for both PCL and PLLA com-
ponents were determined for all networks (Table 1;
Figure 1a). As observed previously,[8] the PCL-DA control
networks exhibited the crystallinity required to serve as
shape memory switching segments. Crystallinity was
higher for networks prepared with PCL-DA n = 45 versus Figure 1.  a) PCL and PLLA % crystallinity values for PCL–PLLA semi-
n = 25, with longer PCL chains between cross-links more IPNs and PCL-DA controls (**p < 0.01 versus corresponding PCL-DA
network control; +p < 0.05, ++p < 0.01 versus corresponding n = 25
able to organize into lamellae. PCL–PLLA semi-IPNs network). b) Photo time-series of n = 25 SMPs (left) and n = 45
permitted PCL crystallization, although this decreased SMPs (right) during shape recovery upon submersion in 80 °C
somewhat with increased PLLA content. We attribute water at t = 0 and 8 s.

Macromol. Rapid Commun. 2016,  DOI: 10.1002/marc.201600414


www.MaterialsViews.com © 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim 3
Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Macromolecular
Rapid Communications L. N. Woodard et al.

www.mrc-journal.de

diminished crystallinity to the reduction of PCL chain (PCL:PLLA) prepared with n = 25 versus n = 45 PCL-DA.
mobility and overall decreased proximity among PCL In the case of TS, values of PCL-DA network controls and
chains when surrounded by an increasing concentration PCL–PLLA semi-IPNs were higher by ≈27% when prepared
of thermoplastic PLLA chains. In addition, as molecularly with n = 45 rather than n = 25 PCL-DA. Due to a reduced
mobile un-crosslinked chains within the PCL-DA network, cross-link density, these networks were able to undergo
PLLA also exhibited the ability to crystallize, with crystal- more strain, permitting them to reach a higher maximum
linity increasing with PLLA content. stress. Semi-IPNs based on 90:10 and 75:25 wt% PCL:PLLA
Upon placement in 80 °C (Thigh) water, the extent of exhibited statistically similar TS values versus the corre-
shape recovery from the “fixed,” temporary shape (coil) to sponding PCL-DA network control.
the permanent shape (flat strip) was observed (Figure 1b). Degradation behavior was assessed under accelerated
By 8 s, versus the PCL-DA network controls, PCL–PLLA conditions (1 m NaOH, 37 °C) in terms of mass loss as well
semi-IPNs had not quite achieved the full recovery seen as observed physical erosion. Mass loss occurred at a sub-
for PCL-DA controls. As the switching segment for shape stantially higher rate for PCL–PLLA semi-IPNs versus the
memory behavior, PCL crystalline domains maintain the corresponding PCL-DA network controls, which showed
temporary shape and, upon melting, dictate recovery.[5] effectively negligible weight loss (Figure 3a). As has been
For the semi-IPNs, at 80 °C, although the PCL-DA network observed for PCL and PLLA blends,[18] we hypothesize that
is in the melted state (T > Tm), PLLA remains semi-crys- phase separation as well as decreased PCL crystallinity led
talline (Tm ∼ 175 °C).[12] Thus, we hypothesize that semi- to enhanced solution diffusion and hydrolysis rates for
crystalline PLLA regions reduce the volume fraction of PCL the PCL–PLLA semi-IPNs versus PCL-DA network controls.
crystalline domains, restricting switching phase mobility For semi-IPNs based on PCL-DA n = 25, an increased PLLA
and slightly reducing the rate of recovery. content produced a systematic increase in weight loss. For
The tensile modulus (E), TS, and strain at break those based on PCL-DA n = 45, differences in weight loss
(% ε) were determined for all SMP networks (Table 1; with PLLA content were less pronounced, perhaps due to
Figure 2). For PCL–PLLA semi-IPNs, when the PLLA content the relatively lower cross-link density that enhanced deg-
was increased to 75:25 wt% (PCL:PLLA), ≈23% and ≈15% radation overall. We confirmed that degradation rates were
increases in E were observed versus the corresponding not due to the dissolution and extraction of PLLA, as sam-
PCL-DA network controls for n = 25 and n = 45 PCL- ples under non-accelerated conditions (pH ∼ 7, 37 °C) did
DA, respectively. A continued increase in PLLA content not show detectable mass loss from a 60:40 wt% (PCL:PLLA)
(60:40 wt% [PCL:PLLA]) yielded reduced E values, with the semi-IPN (see Figure S7, Supporting Information).
n = 25 semi-IPN fracturing when secured in clamps. This While PCL and PLLA are known to degrade through bulk
brittle behavior is attributed to a high total crystallinity erosion,[4] surface erosion was observed to occur for the
from both PCL and PLLA components. Due to an increased semi-IPNs such that specimen thickness decreased over
cross-link density, E was higher for the 75:25 wt% time (Figure 3b, see Figures S5 and S6, Supporting Informa-

Figure 2.  a) Tensile modulus (E) values of SMP networks (**p < 0.01 vs corresponding PCL-DA network control; +p < 0.05, ++p < 0.01 vs cor-
responding n = 25 network). b) Tensile strength (TS) values of SMP networks (**p < 0.01 vs corresponding PCL-DA network control; ++p < 0.01
vs corresponding n = 25 network).

Macromol. Rapid Commun. 2016,  DOI: 10.1002/marc.201600414


4 © 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim www.MaterialsViews.com

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Macromolecular
PCL–PLLA Semi-IPN Shape Memory Polymers (SMPs): Degradation and Mechanical Properties Rapid Communications
www.mrc-journal.de

Figure 3.  a) Film mass loss under accelerated conditions (1 m NaOH, 37 °C) for n = 25 (top) and n = 45 (bottom) compositions. b) SEM images
of film cross-sections after 48 h of degradation [scale bars = 200 μm].

tion). The interesting erosion behavior, however, is thought Supporting Information


to be attributed to the high pH conditions.[19,20]
Supporting Information is available from the Wiley Online
Library or from the author.
4. Conclusions
Acknowledgements: Funding from the Texas A&M Engineering
In this work, PCL–PLLA SMP semi-IPNs were prepared via Experiment Station (TEES) and the NSF Graduate Research
Fellowship (NSF GRFP) (L.W.) is gratefully acknowledged.
the photochemical cure of PCL-DA and PLLA. The PCL:PLLA
wt% ratios were varied (100:0 [PCL-DA control], 90:10, 75:25,
Received: June 30, 2016; Revised: September 20, 2016;
60:40) as was the PCL-DA molecular weight (n = 25, 45). Published online: ; DOI: 10.1002/marc.201600414
The PCL-DA network provided the switching segments
(i.e., Tm = Ttrans) and netpoints (i.e., chemical cross-links) Keywords: degradation; poly(ε-caprolactone); poly(l-lactic acid);
required for shape memory behavior, while un-crosslinked semi-interpenetrating networks; shape memory polymers
PLLA was incorporated to broaden and tailor proper-
ties. Due to semi-IPN design, PCL underwent sufficient
crystallization (21%–40% crystallinity) for shape memory, [1] A. Lendlein, M. Behl, B. Hiebl, C. Wischke, Expert Rev. Med.
while PLLA also exhibited the desired semi-crystallinity. Devices 2010, 7, 357.
Whereas 75:25 wt% (PCL:PLLA) semi-IPNs exhibited an [2] S. Witold, M. Annick, H. Shunichi, Y. L’Hocine, R. Jean,
increased E (n = 25:≈93 MPa; n = 45:≈77 MPa) over cor- Biomed. Mater. 2007, 2, S23.
responding PCL-DA network controls (n = 25:≈76 MPa; [3] G. J. Berg, M. K. McBride, C. Wang, C. N. Bowman, Polymer
2014, 55, 5849.
n = 45:≈67 MPa), TS was higher for controls and semi-IPNs [4] J. C. Middleton, A. J. Tipton, Biomaterials 2000, 21, 2335.
prepared with n = 45 PCL-DA (≈23–25 MPa) rather than [5] A. Lendlein, A. M. Schmidt, M. Schroeter, R. Langer, J. Polym.
n = 25 (≈18–19 MPa). Under accelerated conditions, while Sci., Part A: Polym. Chem. 2005, 43, 1369.
PCL-DA network controls exhibited minimal degradation [6] E. D. Rodriguez, X. Luo, P. T. Mather, ACS Appl. Mater.
(≈5% loss) after 72 h (1 m NaOH, 37 °C), PCL–PLLA semi-IPNs Interfaces 2011, 3, 152.
[7] D. Zhang, M. L. Giese, S. L. Prukop, M. A. Grunlan, J. Polym.
degraded rapidly (≈80%–100% loss). With their broadened Sci. Pol. Chem. 2011, 49, 754.
properties, these semi-IPNs may enhance the utility of bio- [8] C. A. Schoener, C. B. Weyand, R. Murthy, M. A. Grunlan,
degradable SMPs. J. Mater. Chem. 2010, 20, 1787.

Macromol. Rapid Commun. 2016,  DOI: 10.1002/marc.201600414


www.MaterialsViews.com © 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim 5
Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Macromolecular
Rapid Communications L. N. Woodard et al.

www.mrc-journal.de

[9] L. Peponi, I. Navarro-Baena, A. Sonseca, E. Gimenez, [15] H. Kweon, M. K. Yoo, I. K. Park, T. H. Kim, H. C. Lee, H.-S. Lee,
A. Marcos-Fernandez, J. M. Kenny, Eur. Polym. J. 2013, 49, J.-S. Oh, T. Akaike, C.-S. Cho, Biomaterials 2003, 24, 801.
893. [16] C. G. Pitt, F. Chasalow, Y. Hibionada, D. Klimas, A. Schindler,
[10] M. Nagata, I. Kitazima, Colloid Polym. Sci. 2006, 284, 380. J. Appl. Polym. Sci. 1981, 26, 3779.
[11] M. Saatchi, M. Behl, U. Nöchel, A. Lendlein, Macromol. Rapid [17] G. L. Siparsky, K. J. Voorhees, J. R. Dorgan, K. Schilling,
Commun. 2015, 36, 880. J. Environ. Polym. Degrad. 1997, 5, 125.
[12] K. M. Nampoothiri, N. R. Nair, R. P. John, Bioresour. Technol. [18] H. Tsuji, Y. Ikada, J. Appl. Polym. Sci. 1998, 67, 405.
2010, 101, 8493. [19] F. von Burkersroda, L. Schedl, A. Göpferich, Biomaterials
[13] X. Lu, W. Cai, Z. Gao, J. Appl. Polym. Sci. 2008, 108, 1109. 2002, 23, 4221.
[14] M. Nagata, Y. Yamamoto, Macromol. Chem. Phys. 2010, 211, [20] H. Tsuji, Y. Ikada, J. Polym. Sci., Part A: Polym. Chem. 1998, 36,
1826. 59.

Macromol. Rapid Commun. 2016,  DOI: 10.1002/marc.201600414


6 © 2016  WILEY-VCH Verlag GmbH &  Co.  KGaA, Weinheim www.MaterialsViews.com

Early View Publication; these are NOT the final page numbers, use DOI for citation !!

You might also like