improving the oxidative stability of shape memory

10
Improving the Oxidative Stability of Shape Memory Polyurethanes Containing Tertiary Amines by the Presence of Isocyanurate Triols A. C. Weems,* ,J. K. Carrow, A. K. Gaharwar, ,,§ and D. J. Maitland Department of Biomedical Engineering, Department of Material Sciences, and § Center for Remote Health Technologies and Systems, Texas A&M University, College Station, Texas 77843, United States * S Supporting Information ABSTRACT: Shape memory polymers (SMPs) have been proposed for a wide variety of biomedical applications, such as cerebrovascular aneurysm occlusion; however, risks associated with degradation byproduct formation have raised the need for more biostable formulations. In this study, the use of cyclized isocyanates, in the form of isocyanurate-containing alcohols, was examined in aliphatic SMPs as a method of improving biostability. The materials were investigated for thermomechanical behavior, shape memory response, biostability, and cytocom- patibility. In vitro experiments indicated that the isocyanurate-containing porous SMPs possess similar mechanical properties, albeit with improved toughness and no impact to strain recovery shape memory behavior, although atmospheric water uptake increased. Importantly, SMP life span was greatly increased, as control materials were demonstrated to fully degrade in accelerated testing by 25 days, while the modied materials retain more than 90% of their original mass. These modications also resulted in increased cytocompatibility, indicating that this method may be used for translating biomaterials toward clinical applications. INTRODUCTION Polyurethane (PUs) materials have been a gold standard for blood contacting medical devices, particularly for cardiovas- cular and similar long-term implants. 1 While there are many studies focusing on the development of polyurethanes with tunable degradation rates, production of more biostable PUs is not as widely studied as degradable PU formulations, despite failure resulting in billions of dollars in damages for pacemaker lead coatings alone. 17 For shape memory polymer (SMP) applications, such as cerebrovascular aneurysm occlusion, reducing degradation product accumulation may reduce risks associated with the long-term implantation of the device, as product accumulation close to the brain may result in a regulatory challenge for translation out of the lab and into the clinic. 8 Greater biostability of PUs may result in fewer failures and reduced patient risk due to decreased degradation product exposure, both of which ultimately lead to improved clinical outcomes. Previous work with aliphatic, highly porous thermoset SMP PUs revealed superior hydrolytic stability compared with aromatic PUs of similar morphology, with 1% of mass lost after 250 days in 0.1 N NaOH solution. These studies are similar to those examining thermoplastic PUs with ether soft segments; the authors approximated 50% mass loss within 80 years in vivo. 4 Unfortunately, amine-containing SMPs are susceptible to oxidation, completely degrading within 18 months in vitro, due to the oxidation of tertiary amines. 811 The formation of N-oxides from tertiary amines was studied using hydrogen peroxide, where the N-oxides may fragment through oxidationreduction reactions to form aldehydes and lower amines and are in agreement with previous studies of the SMPs. 811 These aldehydes may then undergo subsequent reactions, such as with nearby nitrogen atoms, or may further oxidize to carboxylic acids. Our in vitro analysis indicates that while the starting materials are cytocompatible, degradation may result in potentially toxic byproducts and surfaces with reduced cellular adhesion. 8 Ultimately, the reactivity of aldehydes in addition to the toxicity of the degrading tetrafunctional alcohol, which may form the renal toxin oxalic acid, indicates that an alternative monomer structure is needed to ensure long-term cytocompatibility. While toxic risk from these materials is expected to be minimal due to their low density, further risk reduction could be accomplished by increasing oxidative stability (reducing the rate of mass loss). Cyclization of isocyanates, which may be achieved in the presence of appropriate catalysts, produces six- membered rings (isocynaurate) that have greater thermal stability and imbues the bulk material with more robust mechanical properties. 1214 While these isocyanurates have been investigated for coatings and thermally stable foams, very little work has been pursued in biomedical applications. 15 More recently, these structures have been used as cross-linkers in self-healing, shape memory, and other functional or advanced materials. We hypothesize that this structure could be used to develop more biostable SMP formulations. 1619 More importantly, isocyanurates present a more similar Published: November 5, 2018 Article pubs.acs.org/Macromolecules Cite This: Macromolecules 2018, 51, 9078-9087 © 2018 American Chemical Society 9078 DOI: 10.1021/acs.macromol.8b01925 Macromolecules 2018, 51, 90789087 Downloaded via TEXAS A&M UNIV COLG STATION on April 27, 2020 at 01:46:37 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Upload: others

Post on 22-Apr-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Improving the Oxidative Stability of Shape Memory

Improving the Oxidative Stability of Shape Memory PolyurethanesContaining Tertiary Amines by the Presence of Isocyanurate TriolsA. C. Weems,*,† J. K. Carrow,† A. K. Gaharwar,†,‡,§ and D. J. Maitland†

†Department of Biomedical Engineering, ‡Department of Material Sciences, and §Center for Remote Health Technologies andSystems, Texas A&M University, College Station, Texas 77843, United States

*S Supporting Information

ABSTRACT: Shape memory polymers (SMPs) have been proposed for a widevariety of biomedical applications, such as cerebrovascular aneurysm occlusion;however, risks associated with degradation byproduct formation have raised theneed for more biostable formulations. In this study, the use of cyclized isocyanates,in the form of isocyanurate-containing alcohols, was examined in aliphatic SMPsas a method of improving biostability. The materials were investigated forthermomechanical behavior, shape memory response, biostability, and cytocom-patibility. In vitro experiments indicated that the isocyanurate-containing porousSMPs possess similar mechanical properties, albeit with improved toughness andno impact to strain recovery shape memory behavior, although atmospheric wateruptake increased. Importantly, SMP life span was greatly increased, as controlmaterials were demonstrated to fully degrade in accelerated testing by 25 days,while the modified materials retain more than 90% of their original mass. Thesemodifications also resulted in increased cytocompatibility, indicating that thismethod may be used for translating biomaterials toward clinical applications.

■ INTRODUCTION

Polyurethane (PUs) materials have been a gold standard forblood contacting medical devices, particularly for cardiovas-cular and similar long-term implants.1 While there are manystudies focusing on the development of polyurethanes withtunable degradation rates, production of more biostable PUs isnot as widely studied as degradable PU formulations, despitefailure resulting in billions of dollars in damages for pacemakerlead coatings alone.1−7 For shape memory polymer (SMP)applications, such as cerebrovascular aneurysm occlusion,reducing degradation product accumulation may reduce risksassociated with the long-term implantation of the device, asproduct accumulation close to the brain may result in aregulatory challenge for translation out of the lab and into theclinic.8 Greater biostability of PUs may result in fewer failuresand reduced patient risk due to decreased degradation productexposure, both of which ultimately lead to improved clinicaloutcomes.Previous work with aliphatic, highly porous thermoset

SMP PUs revealed superior hydrolytic stability comparedwith aromatic PUs of similar morphology, with ∼1% of masslost after 250 days in 0.1 N NaOH solution. These studies aresimilar to those examining thermoplastic PUs with ether softsegments; the authors approximated 50% mass loss within 80years in vivo.4 Unfortunately, amine-containing SMPs aresusceptible to oxidation, completely degrading within 18months in vitro, due to the oxidation of tertiary amines.8−11

The formation of N-oxides from tertiary amines was studiedusing hydrogen peroxide, where the N-oxides may fragment

through oxidation−reduction reactions to form aldehydes andlower amines and are in agreement with previous studies of theSMPs.8−11 These aldehydes may then undergo subsequentreactions, such as with nearby nitrogen atoms, or may furtheroxidize to carboxylic acids. Our in vitro analysis indicates thatwhile the starting materials are cytocompatible, degradationmay result in potentially toxic byproducts and surfaces withreduced cellular adhesion.8 Ultimately, the reactivity ofaldehydes in addition to the toxicity of the degradingtetrafunctional alcohol, which may form the renal toxin oxalicacid, indicates that an alternative monomer structure is neededto ensure long-term cytocompatibility.While toxic risk from these materials is expected to be

minimal due to their low density, further risk reduction couldbe accomplished by increasing oxidative stability (reducing therate of mass loss). Cyclization of isocyanates, which may beachieved in the presence of appropriate catalysts, produces six-membered rings (isocynaurate) that have greater thermalstability and imbues the bulk material with more robustmechanical properties.12−14 While these isocyanurates havebeen investigated for coatings and thermally stable foams, verylittle work has been pursued in biomedical applications.15

More recently, these structures have been used as cross-linkersin self-healing, shape memory, and other functional oradvanced materials. We hypothesize that this structure couldbe used to develop more biostable SMP formulations.16−19

More importantly, isocyanurates present a more similar

Published: November 5, 2018

Article

pubs.acs.org/MacromoleculesCite This: Macromolecules 2018, 51, 9078−9087

© 2018 American Chemical Society 9078 DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

Dow

nloa

ded

via

TE

XA

S A

&M

UN

IV C

OL

G S

TA

TIO

N o

n A

pril

27, 2

020

at 0

1:46

:37

(UT

C).

See

http

s://p

ubs.

acs.

org/

shar

ingg

uide

lines

for

opt

ions

on

how

to le

gitim

atel

y sh

are

publ

ishe

d ar

ticle

s.

Page 2: Improving the Oxidative Stability of Shape Memory

material chemistry as opposed to altering the composition byincorporating ethers (which may oxidize in thermosets),carbonates (which may hydrolyze), siloxanes (which mayexhibit significantly altered mechanical behavior and phaseseparation), or olefins (which may phase separate or beotherwise immiscible for processing).4,20,21

This paper presents the results from of utilizing theisocyanaurate functionality in aliphatic polyurethane SMPswith the intent of improving biostability and examining theroles of both diisocyanate and amino alcohol with regards tothermomechanical and biostability behavior. SMP composi-tions were varied to increase the material life span to predictedvalues of nearly 20 years compared to the 18 month life span ofthe controls.

■ METHODS AND MATERIALSMaterials. Hexamethylene diisocyanate (98%, HDI, Sigma), 2,4,4-

trimethylene hexamethylene diisocyanate (98%, TMHDI, Sigma),triethanolamine (99%, TEA, Sigma), N,N,N′,N′-tetrakis(hydroxy-propyl)ethylenediamine (98%, HPED, Sigma), and tris(2-hydrox-yethyl) isocyanurate (Sigma) were used without modification (Figure1). Deuterated dimethyl sulfoxide (d-DMSO, 99.99%, Sigma-Aldrich),ethanol (97%, Sigma-Aldrich), dimethylformamide (DMF, 99.5%,Sigma), acetone (99.5%, VWR), and isopropyl alcohol (99.5%, Sigma-Aldrich) were used as solvents. Ethyl isocyanate (98%, Sigma-Aldrich)was used without purification. Cobalt chloride (CoCl2) and hydrogenperoxide (H2O2, 50%, Sigma-Aldrich) were used for degradationsolutions.General Characterization. Matrix-assisted laser desorption/

ionization (MALDI)−time-of-flight mass spectrometry (MS) wasused to characterization the prepolymer network. Fourier transforminfrared spectroscopy (FTIR)−attenuated total reflectance (ATR)was performed using a Bruker ALPHA infrared spectrometer (Bruker,Billerica, MA); 48 scans per spectra for both background and sampleswere used. Spectra data were collected in absorption mode with aresolution of 4 cm−1. OPUS software was used to examine spectra,identify peaks, and perform baseline and atmospheric corrections.Nuclear magnetic resonance (NMR) (13C 125 MHz and 1H 300MHz) was performed on a Mercury 300 MHz spectrometer operatingin the Fourier transform mode with CDCl3 as the standard. Liquidchromatography−mass spectrometry (LCMS) was performed using asingle quad OrbiTrap (ThermoFisher) with Exactive software and aC18 normal phase silica column. The capillary and heater temperatureswere set for 50°C to prevent thermal-induced fragmentation.Synthesis of Isocyanurate Prepolymer. ISO (15.78 g, 0.060

mol) was added at room temperature to a flask containing DMF (6mL) and acetone (15 mL). HDI (4.2 g, 0.025 mol) was then addeddropwise over the course of 10 min while stirring. The flask was thenheated with stirring to 180 °C over the course of 1 h and held foranother hour before being allowed to cool to room temperature. Theviscous polymer product (clear) was then characterized and used in

subsequent syntheses. Characterization: 1H NMR (d-DMSO): 4.38(2H, d, NCH2CH2OCO), 4.05 (2H, d, NCH2CH2OCO), 4.04 (2H,d HOCH2CH2N), 3.51 (2H, d, HOCH2CH2N), 3.21 (2H, d,OCONHCH2CH2CH2), 1.66 (2H, d, OCONHCH2CH2CH2) 1.21(2H, d, OCONHCH2CH2CH2).

13C NMR (d-DMSO): 157.18,150.13, 61.38, 58.75, 46.47, 42.98, 37.31, 36.40, 26.27. Mn = 2490 g/mol, PDI (2.05).

Synthesis of Shape Memory Polymer. Films were produced byreacting stoichiometric amounts of alcohols and isocyanates (105:100NCO:OH) in solvent. Isocyanurate prepolymer (7.649 g, 3.07 mmol,30% alcohols) and TEA (7.129 g, 47.85 mmol, 70% alcohols) weredissolved in DMF, followed by the addition of HDI (17.22 g, 102.44mmol, 100% isocyanates). The solution was mixed and cast into films.This protocol was repeated for films containing TMHDI and HPED.Nomenclature and ratios of composition are presented in Table 1.

SMP porous scaffolds were synthesized using a traditional two-stepprocess for porous scaffolds.8 An isocyanate premix consisting of thefull measure of isocyanates and 40% of the hydroxyl groups (ratio of105:35 NCO:OH), with the isocyanurate premix contributing 30% ofthe total alcohol concentration, was prepared 48 h in advance.Isocyanurate prepolymer (4.208 g, 1.69 mmol) was dissolved in 6 mLof DMF. TEA (3.922 g, 26.32 mmol) was added and mixed for 5 min.HDI (27.07 g, 161.04 mmol) was added over 2 min and mixed for 5min. The premix was heated to 50 °C and held isothermally undervacuum for 48 h. For the control samples, HPED was substituted forthe isocyanurate.

An alcohol premix was then synthesized as the second step.Isocyanurate prepolymer (2.186 g, 0.88 mmol, 30% Iso), TEA (2.038g, 13.68 mmol), water (1.081, 59.99 mmol), surfactant DC 5943 (2.8

Figure 1. Starting materials and nomenclature for synthesis of prepolymer and SMP.

Table 1. Nomenclature and Compositions of ExaminedSpecies

speciesdiisocyante

(NCO species)TEA

(OH %)HPED(OH %)

iso(OH%) morphology

HH-30 HDI 70 30 0 porousfoam

HTISO-30p

HDI 70 0 30 porousfoam

HTISO-20p

HDI 80 0 20 porousfoam

HTISO-10p

HDI 90 0 10 porousfoam

HT-100s HDI 50 50 0 solid filmHTISO-50s

HDI 50 0 50 solid film

HHISO-50s

HDI 0 50 50 solid film

TTISO-50s

TMHDI 50 0 50 solid film

THISO-50s

TMHDI 0 50 50 solid film

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9079

Page 3: Improving the Oxidative Stability of Shape Memory

g), tin catalyst (0.3.g), and amine catalyst (0.75 g) were mixed for 2min and added to the isocyanate premix with a low boiling pointblowing agent. The mixture was immediately placed in a 90 °C ovenand held for 20 min, with the primary foam blowing occurring withinthe initial 5 min. This protocol was followed for varied ratios of TEAand isocyanurate prepolymer (30% Iso for 30 Iso, 20 Iso, 10 Iso,control compositions were investigated). The SMPs were then cutand washed. The cleaning process consisted of alternating washes inRO H2O and EtOH, at 50°C with sonication for 30 min, two cycles ofeach. The SMPs were dried for 3 days at 50°C under vacuum (30in.Hg) and stored in a sealed box with desiccant. HDI TEA 30 IsoFTIR: 3321, 2926, 2853, 1688, 1638, 1538, 1457, 1362, 1292, 1182,1135, 1035 cm−1. For the control samples, HPED was substituted forthe isocyanurate.Physical Properties. The gel fraction was determined for selected

compositions (n = 3). Samples (∼0.10 mg) were immersed in 30 mLof ethanol and heated to 50 °C for a 5 days. The samples were thenremoved, blotted dry, and placed in a vacuum oven at 50 °C (30in.Hg) for 2 days to remove any solvent before being weighed.Density was determined from porous SMP cubes (1 cm3, n = 3).Porous SMP samples were removed from vacuum at roomtemperature and stored in individual open containers at ambientconditions overnight. The mass of the sample was measuredperiodically.Thermal Analysis. Dynamic mechanical analysis (DMA) of

cylindrical samples (6 mm diameter, 5 mm length) was used toconduct thermomechanical analysis using a Q800 TA DMA (TAInstruments, New Castle, DE). Dry temperature sweep samples wereequilibrated at 0 °C for 15 min and then ramped to 200 °C (films), 90°C (dry foams), and 70 °C (wet foams) at a rate of 2 °C/min. Thestorage modulus (E′) and the loss modulus (E″) were used todetermine the peak tan δ(E′/E″), with the maximum value recordedas the Tg. Differential scanning calorimetry (DSC) was also used tomeasure both wet and dry Tg using a Q-200 DSC (TA Instruments,Inc., New Castle, DE). Samples of ∼5.0 ± 1.0 mg were sealed inTZero aluminum pans and placed in the heating cell. The test profilewas as follows: equilibration at −40 °C, heating to 180 °C at 10 °C/min, cooling 10 °C/min to −40 °C and holding for 5 min, and a finalheating to 180 °C at 10 °C/min. The half-height transition of the final

heating cycle was the reported Tg. Wet samples were weighed andsealed in the same manner and were then heated from −40 to 120 °C.

Shape Recovery. Shape recovery of cylindrical foam samples (6mm diameter, 10 mm length, six samples per series) crimped over awire was examined at 37 °C in RO water to determine the volumerecovery behavior (strain recovery). Samples were crimped to aminimal diameter (∼1.0 mm using a SC150-42 stent crimper(Machine Solutions, Flagstaff, AZ). To crimp, samples wereequilibrated at 100 °C for 10 min and then radially compressedand cooled to room temperature. Samples relaxed for 12 h in a sealedbox with desiccant and were tested over the course of 20 min. ImageJ(NIH, Bethesda, MD) was used for analysis of the change indiameters over time.

Microscopy Analysis. Foam cell structure was determined bycutting axial and transverse samples that were examined usingscanning electron microscopy (SEM). Samples were mounted onto astage and sputter coated with gold using a Cressington sputter coater(Ted Pella, Inc., Redding, CA) for 60 s at 20 mA. Samples were thenexamined using a Joel NeoScope JCM-5000 SEM (Nikon InstrumentsInc., Melville, NY) at 11× magnification and 15 kV under highvacuum.

Tensile Testing. Mechanical testing of SMP dog bones wasperformed using tensile testing. An Instron 5965 electromechanicalscrew driven test frame, equipped with a 500 N load cell, was used totest the dog bones at 5 mm/min at room temperature. Both porousand nonporous samples were cut in Type IV D dog bones for testing.

Cytocompatibility. Green fluorescent protein (GFP)-expressingfibroblasts (NIH3T3/GFP, Cell Biolabs Inc., USA) were expandedand then seeded onto control (unmodified) and degraded films at adensity of 20000 cells/film in normal growth media, which containedhigh glucose Dulbecco’s Modified Eagle Medium (DMEM, GEBiosciences, USA) with 10% FBS (Atlanta Biologicals, USA) and 1%penicillin/streptomycin (100 U/μg/mL, Invitrogen, USA). Afterinitial cell seeding, samples were transferred to new tissue culturewells to ensure measured cell growth was confined to the polymericbiomaterial as opposed to the tissue culture polystyrene.

To investigate cell proliferation, an Alamar Blue assay (ThermoFisher, USA) was utilized per the manufacturer’s instructions tocalculate percent reduction of reszurin (blue) to resorufin (pink).Proliferation within foam samples was normalized to accessible

Figure 2. 1H NMR with 13C spectra inset of isocyanurate prepolymer, with peaks noted for an idealized structure.

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9080

Page 4: Improving the Oxidative Stability of Shape Memory

surface area via considerations of foam volume and percent porosity.At the respective time points (day 0, day 1, day 3, and day 7), cellmorphology was evaluated using fluorescence microscopy with anexcitation of 488 nm (Nikon Ti-U Inverted Microscope, USA); forday 0 time point, cells were allowed to adhere for 3 h prior to imaging.Representative images were selected over the course of imaging threereplicates for each film or foam.Degradation Testing. For analysis of degradation rates, cleaned

samples were completely immersed in respective solutions of 3%H2O2 (real time oxidation) and 20% H2O2 with 0.1 M CoCl2(accelerated oxidation), both stored at 37 °C, based upon previouslypublished procedures.8

Model Compounds. Model compounds were dissolved in 50%H2O2 and placed in a 37 °C oven for 5 weeks, with sampleswithdrawn for analysis on a weekly basis; peroxide was refreshed atthe same time. Hydrolytic samples were added to 4 M NaOH andincubated for the same time. Samples were diluted 100-fold in a 50%methanol (49.95% water, 0.05% formic acid) solution and injectedinto the LCMS column. The injection volume was 10 μL, and theflow gradient was varied from 0% acetonitrile to 95% acetonitrile overa 3 min gradient. Exactive software (ThermoFischer) was used foranalysis of the chromatography scan and the peaks.

■ RESULTS AND DISCUSSIONSynthesis of SMPs. Qualitatively, an increase in solution

viscosity was noted along with the formation of theisocyanurate prepolymer, which was confirmed using MALDImass spec.22,23 1H NMR (Figure 2) displays a peakcorresponding to the HDI backbone urethane at ca. 3.30ppm, indicating that ca. half of the alcohol groups areconsumed during prepolymer synthesis, which corresponds tothe initial stoichiometry. 13C NMR displayed the initialcarbonyl from the isocyanurate at 150.12 ppm and a newpeak at 157.80 ppm corresponding to the urethane carbonyl.The prepolymer was then successfully incorporated into the

SMP foams as determined by the spectroscopic examination ofthe materials (Figure S1) which displayed an increase in thetransmittance at 1456 cm−1, indicative of the quasi-aromaticstructure, and a decrease at 1050 cm−1 corresponding to adecrease in tertiary amine concentration.8 The characteristiccarbonyl peak associated with the urethane linkage occurs at1689 cm−1 with the urea shoulder at ca. 1645 cm−1, as foundwith previous studies.8,24,25

The morphology of the pore cells was found to vary withoutregard to composition (Figure 3) and would be basedprimarily on the amount of residual solvent as well as watercontent and premix viscosity.26,27 For further examinations,pore diameters were limited to materials possessing an averagediameter of 1.1 ± 0.4 mm. As expected from previous workwith porous SMPs, these cross-linked polymers may be

produced with ultralow density and a high gel fraction (allformulations possess gel fractions >99% and densities <0.032g/cm3).24,25

Thermomechanical Characterization. With sufficientpresence of the isocyanurate, the Tg is also high enough to beclinically relevant when dry (greater than room temperature)but will easily plasticize and undergo shape recovery in water(Table 2) and is more hydrophilic relative to the control

materials (Supporting Information Figure S1). While the Tg ofthe HH-100s was found to be 1.3 °C by DMA (Table 3, peak

tan δ value), the addition of the isocyanurate increased the Tgby nearly 80 °C, with HTISO-30s displaying a transition at81.4 °C. While the addition of HPED versus TEA increasedthe thermomechanical properties in the TMHDI series, thiswas not the case for the HDI-based materials. The interplay ofthe hydrogen-bonding and cross-link density with the moreflexible HDI chain, relative to the methylated TMHDIbackbone, may contribute to this behavior, with the increasedpolarity of the isocyanurate offsetting the hydrophobicity of theHPED despite the greater cross-link density. The increase inthermomechanical properties when using TMHDI comparedto HDI is an expected behavior.8

The shoulder that is displayed by the tan δ curves is mostlikely attributable to the urethane linkages due to its proximityto the control peak value (Figure 4). With the increasing

Figure 3. SEM images of 10% (left), 20% (middle), and 30% (right) isocyanurate in SMP foam, with remaining alcohols from TEA and adiisocyanate of HDI.

Table 2. Physical and Thermal Properties of Porous Iso-SMPs from HDI and TEA (n = 3)

compositiongel fraction

(%)density(g/cm3)

wet DSC(°C)

dry DSC(°C)

HH-30p 99.8 0.022 −5.0 55.3HTISO-10p 99.6 0.032 −1.2 34.6HTISO-20p 99.7 0.025 −2.1 37.2HTISO-30p 99.5 0.032 −1.1 57.6

Table 3. Tan δ Peak and Maximum Complex ModulusValue of Nonporous SMPs for Dry Samples Examined byDMA (n = 3)

species tan δ (°C) E* (MPa) (Tg − 20 °C)

HT-100s 1.3 9.2HTISO-30s 81.4 1103.1HHISO-30s 86.8 1036.2TTISO-30s 104.5 2075.6THISO-30s 125.2 2786.1

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9081

Page 5: Improving the Oxidative Stability of Shape Memory

rigidity of the polymer with the use of HPED and TMHDI, thetan δ narrows and begins to separate from the shoulder asdisplayed in the TTISO-30p curve. All ISO-containing samplesalso had complex moduli at least 2 orders of magnitude higherthan the control at 20 °C below Tg, and the inflection points ofthe respective curves mostly correspond to the increased Tg

values.Tensile testing (Figure 5) at ambient conditions further

confirmed the mechanical behavior from DMA testing. Films

displayed distinctly increased mechanical properties with theincorporation of the isocyanurate compared with the controlgroup consisting of only HDI and TEA monomers. Theisocyanurate greatly improved the elastic modulus and ultimatestrength, but at the cost of the strain to failure. Additionally, asthe polymer gets more rigid, the toughness overall decreases,with the exception of the HTISO-30s composition.28 This alsoindicates that further increasing the isocyanurate concentrationin the SMP may reduce mechanical properties even more.

Figure 4. Thermomechanical properties of Iso films, displaying tan δ vs temperature (left) and E′ (storage moduli) vs temperature (right).

Figure 5. Stress−strain behavior of nonporous SMPs, comparing the role of diisocyanate, amino alcohol monomers, and isocyanurate triol onmechanical properties under ambient conditions (A). (B) Porous HTISO-30p compared to HH-30p in both ambient and hydrated (37 °C in PBS)conditions (n = 7, representative curves displayed).

Table 4. Mechanical Properties of Porous and Nonporous SMPs Examined at Ambient and in Vitro Conditions (37 °C in PBS)(n = 7)

species testing condition elastic modulus (MPa) strain to failure (%) ultimate strength (MPa) toughness (J/m3)

HT-100s 22 °C, ambient 9 ± 1 65.7 ± 25 4.5 ± 1 792.18HTISO-30s 22 °C, ambient 502.5 ± 23. 22.7 ± 1 45.7 ± 1 815.35HHISO-30s 22 °C, ambient 671.3 ± 2 20.7 ± 4 46.0 ± 4 724.04TTISO-30s 22 °C, ambient 826.2 ± 154 16.4 ± 5 47.0 ± 1 191.42THISO-30s 22 °C, ambient 889.9 ± 92 7.5 ± 2 48.4 ± 7 178.33HH-30p 22 °C, ambient 0.1 ± 0 130.0 ± 13 0.1 ± 0 32.50HTISO-30p 22 °C, ambient 1.1 ± 0.5 39.4 ± 11 0.2 ± 0 41.21HH-30p 37 °C, submerged in PBS 0.15 ± 0. 64.7 ± 12 0.1 ± 0 5.00HTISO-30p 37 °C, submerged in PBS 0.1 ± 0. 57.6 ± 19 0.1 ± 0 5.16

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9082

Page 6: Improving the Oxidative Stability of Shape Memory

HISO-30p samples were examined using tensile testingcompared with the control SMPs (HH-30p) in both ambientand in vitro conditions (submerged samples at 37 °C). Whilethere were significant mechanical differences when the sampleswere dry, upon exposure to moisture, both series of SMPspossessed similar elastic moduli (0.05 MPa), strain to failure(∼60%), ultimate strength (0.02 MPa), and toughness (ca. 5J/m3) (Table 4). Interestingly, toughness decreased withplasticization of the polymer bonds when immersed. Furtherstudies are needed to understand why this may occur and howit could impact translation of the materials toward a clinicalsetting.Shape Memory Characterization. Despite the changes in

dry mechanical properties, the Iso-SMPs possessed full strainrecovery and strain fixity (Figure 6). While the rate of strain

recovery is statistically slower compared to the controlformulation over the initial 3 min, by 4 min and onwardthere is no difference in behavior. It is possible that theinterplay between the increased rigidity of the polymerbackbone containing isocyanurate is balanced by the increasedhydrophilicity. For comparison with previous studies,morphology was not factored into the recovery kinetics, but

further work is needed to understand the effects of macro-scopic factors such as membrane intactness and diffusivity andpore morphology.8,24,25,29

The presence of the isocyanurate also causes distinct uptakein water, determined through a change both in mass and inmechanical properties using samples that were allowed toequilibrate with the room’s conditions over the course of 24 h.The SMPs displayed increased masses by ∼3% over the courseof hours, and mechanical properties qualitatively changedwithin 2 h (less than a full standard deviation tested over thecourse of this experiment) (Figure 7). Strain to failureincreased from ca. 20% to nearly 100% over the course of2.5 h, while elastic moduli decreased from ca. 0.1 MPa tonearly 0.01 MPa in the same time. This corroborates the rapid1% increase in mass displayed during atmospheric wateruptake over this time period and may be sufficient moisture toresult in SMP plasticization similar to what was found duringthe submersion testing of the control SMPs (HH-30p).

Degradation and Cytocompatibility. Degradation test-ing using gravimetric analysis indicates that while the controlSMPs will undergo rapid oxidative mass loss, depending on theporosity of the SMP, within days or weeks, the addition of theisocyanurate, oxidative stability will greatly increase in bothfilms and porous SMPs (Figure 8). At day 110, films possessed∼70% of the original mass (HTISO-30s) or up to 90%(THISO-30s). For the porous SMPs, while the control sampleshad undergone total mass loss by 15 days, the HTISO-20p andHTISO-30p retained >90% mass loss at this point. Only afterthis period did any significant mass loss begin to occur,indicating that the initiation period hypothesized to corre-spond to the consumption of tertiary amines and theirsubsequent conversation to N-oxides is still occurring, albeitat a slower rate.8−10

Model compound studies were used to propose themechanisms of degradation and to predict whether any toxicbyproducts would be produced. Hydrolytic degradation studiesusing model compounds indicated that the most likely route topolymer fragmentation is due to the urethane (carbamate)hydrolysis rather than a ring-opening of the isocyanurate(Figure S2).30 The hydrolysis of isocyanurates has beenreported for disubstituted structures, resulting in the formationof linear biuret structures, such as in barbituric acids.31 In thesestructures, ring-opening yields malonuric acid, which willfurther fragment to carbon dioxide, acetyl urea, acids, and

Figure 6. Shape recovery behavior of porous HTISOp series SMPs (n= 3).

Figure 7. (A) Gravimetric analysis examining water uptake of HTISOp series and (B) plots of elastic modulus and strain to failure of HTISO-30p.Samples were allowed to equilibrate with ambient conditions over the course of 25 h (gravimetric) and 2.5 h (tensile) (n = 3 at each time point).

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9083

Page 7: Improving the Oxidative Stability of Shape Memory

diamides.31,32 While some ring-opening product was obtained,which may occur due to resonance of the ring, the primaryroute of degradation seems to be cleavage of the urethane toproduce a primary amine and an alcohol (Figure S3).30,31

Oxidatively, the SMPs are distinctly stable, with thedisplayed degradation occurring at the remaining tertiaryamine cross-linking sites. The isocyanurate model compoundmay undergo fragmentation to produce disubstituted iso-cyanurates as a result of oxidative degradation, as found withthermal stability studies of similar compounds, but the relativerate of decomposition makes it unlikely that substantialamounts of cyanic acid would form (Figure S4).33,34 Whilethis step is shown, only the initial fragmentation was detectedover the 5-week experiment and seems to agree with detectedcompounds for isocyanurate derivatives.35

The cytocompatibility results confirm that the TEAproduces a more compatible surface compared with theHPED and that the presence of the isocyanurate may improveinitial cell adhesion (Figure 9). This improvement in celladhesion may be due to increased hydrophobicity (polargroups of the isocyanurate) relative to the amino alcohols. Thepresence of the isocyanurate does not seem to alter thebiocompatibility of polyurethanes. We hypothesize that long-term cellular response would be improved, as oxidation of theISO SMPs would produce fewer N-oxides compared to thecontrol formulation; the N-oxides may act in a similar manner

to surfactants and make the surface more difficult for cellularadhesion.8

The model compound study shows degradation occurringover the course of weeks, in conditions that are notphysiologically representative on even the accelerated scale,as demonstrated by the bulk degradation studies. Previouswork with the degradation of polyurethane foams in non-physiological conditions has been the source of regulatorydiscussion, substantial research, and the recalling of manyimplanted devices. In this case, the model studies are notmeant to demonstrate the rates of formation of toxic products,but rather to demonstrate the possible products and predictthe most likely pathway of degradation. These pathways arethen used to develop a degradation model for toxicity risks. Forthese materials, the pathway to degradation is primarilyoxidative; the primary hydrolytic degradation pathway isthrough carbamate hydrolysis, yielding an alcohol, amine,and carbon dioxide. Previous studies of aliphatic carbamateshave indicated that there is hydrolytic stability in 0.1 M NaOHout to at least 250 days (1% mass loss of the bulk material inthis time); the presented structures were obtained in 4 MNaOH, with carbamate hydrolysis products being the mostabundant.30,31 The ring-opening appears to be less preferentialthan the carbamate hydrolysis, indicating that incorporation ofthe isocyanurate has not sacrificed the hydrolytic stability.

Figure 8. Oxidative degradation of nonporous SMP containing isocyanurate (A) and of porous SMPs containing HDI, TEA, and isocyanurate (B).Samples were degraded in 20% H2O2 catalyzed by 0.1 M CoCl2 at 37 °C. Predicted real time degradation rates for HTISO-30p and HTISO-20pare displayed relative to HH-30p for porcine aneurysm conditions (C) (n = 7).

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9084

Page 8: Improving the Oxidative Stability of Shape Memory

In the control materials, total mass loss in acceleratedoxidative solution occurs by ∼40 days. In the THISO-30sfilms, ∼95% of the original mass is retained at this point. At110 days in accelerated oxidative solution, THISO-30s has lostonly 5.8% of the original mass. For HTISO-30s, 82.1% of theoriginal nonporous films’ mass is retained at 110 days. HH-30pare completely degraded within ∼25 days; HTISO-30p retains97.7% of the original mass at the same time point. At 50 days,HTISO-30p retains ∼75% of its original mass, while HTISO-20p possesses nearly 50%; tuning the rate of oxidative massloss can be achieved through incorporation of varyingconcentrations of the ISO prepolymer to achieve higheroxidative stability. The longer degradation times ensure lesstoxic risk associated with the released degradation productsdue to lower accumulation.The gravimetric change that is displayed may be attributable

to the oxidation of the TEA, and the increased oxidativestability is due to the isocyanurate. The expected degradationproducts are therefore those previously discussed in the studyof the control SMP degradation, including primary amines,aldehydes, and carboxylic acids as a result of fragmentation ofthe tertiary amine to a secondary amine and the correspondingaldehyde, followed by further fragmentation to a primaryamine. The oxidation of the isocyanurate structure is unlikelyand will mostly likely undergo excretion before substantialdegradation, as discussed below. This ring structure mayundergo hydrolysis in the presence of hydroxyl ions due to thecharge distribution between the surrounding carbonyls and the

amine (equilibrium results in the partially charged species andwater), resulting in a 1,6 ring-opening to form 1,5 disubstitutedbiurets.31

Both degradation schemes indicate that the isocyanurate isnot preferentially going to fragment in the in vivo environment,meaning that clearance of the degradation products will likelyinvolve the cyclic structure. In studies of similar molecules, inthis case small molecule isocyanurates with substituted endgroups, clearance occurred via the renal system without changeto the structure for the concentrations that will be producedfrom oxidative degradation of the SMPs.36,37 The cyanuratehad clearance rates of 2 mL/(min kg),with a half-life of 4 h inpigs, without metabolizing the compounds, indicating rapidclearance for similar products as well.37

Based upon the degradation profiles of the control SMPs(HH-30p), it appears that ca. 100% mass loss will occur within1 year; HTISO-20p will have lost ca. 35% mass at the sametime and 30 ISO will have lost ca. 10%. For decreasing thecytotoxicity risk associated with the degradation, the HTISO-20p rate of exposure is 8 times less (0.64 mg total of SMPmaterial/day) and HTISO-30p is 26 times less (0.19 mg/day)than that of the control (5.3 mg/day). Model compoundstudies of biurets and ureas found them to be cytotoxic tohuman breast cancer cells at concentrations of ca. 0.05 mM(14.2 mg/L), which is more than an order of magnitude higherthan the total production rate of degradation products, andstudies on bulk SMPs have confirmed biocompatibility.37

Production of isocyanurate degradation products, assuminguniform production of products, is reduced by another order ofmagnitude based upon composition of the SMPs. While theseSMPs are not as biostable as nonporous polyurethanes used inalternative applications, the use of the isocyanurate doesprovide a means to improve biostability for these porousmaterials, in addition to having a lower theoretical toxicity.

■ CONCLUSIONSWhile the oxidative degradation of porous SMPs is rapid inoxidative environments, reduction in the mass loss may beachieved by the incorporation of isocyanurate triols into theSMP. With some SMP systems, this will increase mechanicaltoughness and other properties without sacrificing shapememory and biocompatibility, both of which are necessaryfor translating a material system out of the lab and into theclinic. These new SMPs possess reduced rates of oxidativedegradation, and it is predicted that they will possess highhydrolytic stability as well due to the main path of hydrolyticdegradation being hydrolysis of the urethane bonds. Overall,this seems like a valid method for pursuing more biostableSMPs for vascular devices.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.macro-mol.8b01925.

Additional material analysis and analysis of isocyanuratefragmentation (PDF)

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected].

Figure 9. Fluorescent 3T3 fibroblasts seeded on films at 7 daysincubation under light microscopy (A) and corresponding metabolicactivity as determined through Alamar Blue reduction (B).

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9085

Page 9: Improving the Oxidative Stability of Shape Memory

ORCIDA. C. Weems: 0000-0002-4683-3820A. K. Gaharwar: 0000-0002-0284-0201NotesThe authors declare the following competing financialinterest(s): Andrew C. Weems holds stock in Shape MemoryMedical, Inc. (SMM), which holds a license for the shapememory polymers developed in Maitland's laboratory. DuncanMaitland owns stock and is on the board of directors at SMM.

■ ACKNOWLEDGMENTSThe authors thank Alexandra Danielle Easley for foamdiscussions, William Wustenburg for toxicological and medicaldevice risk discussions, and Lawrence Dangott for massspectrometry assistance. ThermoFisher generously providedthe OrbiTrap instrument and time on the experiment. Weacknowledge the NASA Harriett G. Jenkins Fellowship(NNX15AU29H, A.C.W.).

■ REFERENCES(1) Dempsey, D. K.; Carranza, C.; Chawla, C. P.; Gray, P.; Eoh, J.H.; Cereceres, S.; Cosgriff-Hernandez, E. M. Comparative Analysis ofIn Vitro Oxidative Degradation of Poly(Carbonate Urethanes) forBiostability Screening. J. Biomed. Mater. Res., Part A 2014, 102, 3649−3665.(2) Padsalgikar, A.; Cosgriff-Hernandez, E.; Gallagher, G.; Touchet,T.; Iacob, C.; Mellin, L.; Norlin-Weissenrieder, A.; Runt, J.Limitations of Predicting In Vivo Biostability of MultiphasePolyurethane Elastomers using Temperature-Accelerated DegradationTesting. J. Biomed. Mater. Res., Part B 2015, 103 (1), 159−168.(3) Smith, R. S.; Zhang, Z.; Bouchard, M.; Li, J.; Lapp, H. S.;Brotske, G. R.; Lucchino, D. L.; Weaver, D.; Roth, L. A.; Coury, A.;Biggerstaff, L.; Sukavaneshvar, S.; Langer, R.; Loose, C. VascularCatheters with a Nonleaching Polysulfobetaine Surface ModificationReduce Thrombus Formation and Microbial Attachment. Sci. Transl.Med. 2012, 4 (153), 153ra132.(4) Chaffin, K. A.; Chen, X.; McNamara, L.; Bates, F. S.; Hillmyer,M. A. Polyether Urethane Hydrolytic Stability after Exposure toDeoxygenated Water. Macromolecules 2014, 47 (15), 5220−5226.(5) Chaffin, K. A.; Buckalew, A. J.; Schley, J. L.; Chen, X.; Jolly, M.;Alkatout, J. A.; Miller, J. P.; Untereker, D. J.; Hillmyer, M. A.; Bates, F.S. Influence of Water on the Structure and Properties of PDMS-Containing Multiblock Polyurethanes. Macromolecules 2012, 45 (22),9110−9120.(6) Nichols, C. I.; Vose, J. G.; Mittal, S. Incidence and costs relatedto lead damage occurring within the first year after a cardiacimplantable electronic device replacement procedure. J. Am. HeartAssoc. 2016, 5, e002813.(7) Martin, J. R.; Gupta, M. K.; Page, J. M.; Yu, F.; Davidson, J. R.;Guelcher, S. A.; Duvall, C. L. A Porous Tissue Engineering ScaffoldSelectively Degraded by Cell-Generated Reactive Oxygen Species.Biomaterials 2014, 35 (12), 3766−3776.(8) Weems, A. C.; Wacker, K. T.; Carrow, J. K.; Boyle, A. J.;Maitland, D. J. Shape Memory Polyurethanes with Oxidation-InducedDegradation: In Vivo and In Vitro Correlations for EndovascularMaterial Applications. Acta Biomater. 2017, 59, 33−44.(9) Colladon, M.; Scarso, A.; Strukul, G. Mild Catalytic Oxidation ofSecondary and Tertiary Amines to Nitrones and N-Oxides with H2O2Mediated by Pt (II) Catalysts. Green Chem. 2008, 10, 793.(10) Bergstad, K.; Backvall, J. E. Mild and Efficient Flavin-CatalyzedH2O2 Oxidation of Tertiary Amines to Amine N-Oxides. J. Org. Chem.1998, 63, 6650−6655.(11) Englert, C.; Hartlieb, M.; Bellstedt, P.; Kempe, K.; Yang, C.;Chu, S. K.; Ke, X.; Garcia, J. M.; Ono, R. J.; Fevre, M.; Wojtecki, R. J.;Schubert, U. S.; Yang, Y. Y.; Hedrick, J. L. Enhancing theBiocompatibility and Biodegradability of Linear Poly(Ethylene

Imine) through Controlled Oxidation. Macromolecules 2015, 48(20), 7420−7427.(12) Driest, P. J.; Lenzi, V.; Marques, L. S. A.; Ramos, M. M. D.;Dijkstra, D. J.; Richter, F. U.; Stamatialis, D.; Grijpma, D. W. AliphaticIsocyanurates and Polyisocyanurates Networks. Polym. Adv. Technol.2017, 28, 1299−1304.(13) Zhang, L.; Rowan, S. J. Effect of Sterics and Degree ofCrosslinking on the Mechanical Properties of Dynamic Poly-(Alkylurea-Urethane) Networks. Macromolecules 2017, 50 (13),5051−5060.(14) Preis, E.; Schindler, N.; Adrian, S.; Scherf, U. MicroporousPolymer Networks made by Cyclotrimerization of Commercial,Aromatic Diisocyanates. ACS Macro Lett. 2015, 4, 1268−1272.(15) Qiu, Y.; Qian, L.; Xi, W. Flame-Retardant Effect of a NovelPhosphaphenanthrene/Triazine-Trione Bi-Group Compound on anEpoxy Thermoset and its Pyrolysis Behavior. RSC Adv. 2016, 6,56018−56027.(16) Nguyen, L. T.; Truong, T. T.; Nguyen, H. T.; Le, L.; Nguyen,V. Q.; Van Le, T.; Luu, A. T. Healable Shape Memory (Thio)-Urethane Thermosets. Polym. Chem. 2015, 6, 3143−3154.(17) Ghosh, S. K.; Bureekaew, S.; Kitagawa, S. A Dynamic,Isocyanurate-Functionalized Porous Coordination Polymer. Angew.Chem. 2008, 120 (18), 3451−5454.(18) Peng, C.; Chang, K.; Weng, C.; Lai, M.; Hsu, C.; Hsu, S.; Li, S.;Wei, Y.; Yeh, J. UV-Curable Nanocasting Technique to Prepare Bio-Mimetic Super-Hydrophobic Non-Fluorinated Polymeric Surfaces forAdvanced Anticorrosive Coatings. Polym. Chem. 2013, 4, 926−932.(19) Tian, K.; Wu, Z.; Xie, F.; Hu, W.; Li, L. Nitrogen-DopedPorous Carbons Derived from Triarylisocyanurate-Cored Polymerswith High CO2 Adsorption Properties. Energy Fuels 2017, 31, 12477−12486.(20) Li, Z.; Yang, J.; Ye, H.; Ding, M.; Luo, F.; Li, J.; Li, J.; Tan, H.;Fu, Q. Simultaneous Improvement of Oxidative and HydrolyticResistance of Polycarbonate Urethanes based on Polydimethylsilox-ane/Poly(Hexamethylene Carbonate) mixed Macrodiols. Biomacro-molecules 2018, 19, 2137−2145.(21) Wang, Y.; Hillmyer, M. A. Oxidatively Stable PolyolefinThermoplastic and Elastomers for Biomedical Applications. ACSMacro Lett. 2017, 6, 613−618.(22) Pavlicevic, J. Preparation and Thermal Stability of Elastomersbased on Irregular Poly(Urethane-Isocyanurate) Networks. Mater.Manufact. Process. 2009, 24, 1217−1223.(23) Jiang, Z.; Liu, G. Microencapsulation of AmmoniumPolyphosphate with Melamine-Formaldehyde-Tris(2-Hydroxythyl)Isocyanurate Resin and its Flame Retardancy in Polypropylene. RSCAdv. 2015, 5, 88445−88455.(24) Weems, A. C.; Szafron, J. M.; Easley, A. E.; Herting, S.; Smolen,J.; Maitland, D. J. Shape Memory Polymers with Enhanced Visibilityfor Magnetic Resonance and X-Ray Imaging Modalities. ActaBiomater. 2017, 54, 45−57.(25) Weems, A. C.; Raymond, J. E.; Easley, A. D.; Wierzbicki, M. A.;Gustafson, T.; Monroe, M. B. B.; Maitland, D. J. Shape MemoryPolymers with Visible and Near-Infrared Imaging Modalities:Synthesis, Characterization and In Vitro Analysis. RSC Adv. 2017,7, 19742−19753.(26) Quell, A.; de Bergolis, B.; Drenckhan, W.; Stubenrauch, C.How the Locus of Initiation Influences the Morphology and the PoreConnectivity of a Monodisperse Polymer Foam.Macromolecules 2016,49, 5059−5067.(27) Viswanathan, P.; Johnson, D. W.; Hurley, C.; Cameron, N. R.;Battaglia, G. 3D Surface Functionalization of Emulsion-TemplatedPolymeric Foams. Macromolecules 2014, 47, 7091−7098.(28) Han, C.; Ran, X.; Zhang, K.; Zhuang, Y.; Dong, L. Thermal andMechanical Properties of Poly(ε-caprolactone) Crosslinked with γRadiation in the Presence of Triallyl Isocyanurate. J. Appl. Polym. Sci.2007, 103, 2676−2681.(29) Boyle, A. J.; Weems, A. C.; Nash, L. D.; Maitland, D. J. Solventstimulated actuation of polyurethane-based shape memory polymer

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9086

Page 10: Improving the Oxidative Stability of Shape Memory

foams using dimethyl sulfoxide and ethanol. Smart Mater. Struct.2016, 25 (7), 075014.(30) Argabright, P. A.; Phillips, B. L. On the Basic Hydrolysis ofDisubstituted Isocyanurates. J. Heterocycl. Chem. 1970, 7, 999−1000.(31) Garrett, E. R.; Bojarski, J. T.; Yakatan, G. J. Kinetics ofHydrolysis of Barbituric Acid Derivatives. J. Pharm. Sci. 1971, 60 (8),1145−1154.(32) Hammond, B. G.; Barbee, S. J.; Inoue, T.; Ishida, N.; Levinskas,G. J.; Stevens, M. W.; Wheeler, A. G.; Cascieri, T. A Review of theToxicology Studies on Cynaurate and its Chlorinated Derivative.Environ. Health Perspect. 1986, 69, 287−292.(33) Duff, D. W.; Maciel, G. E. Monitoring the ThermalDegradation of an Isocyanurate-Rich MDI-based Resin by 15N and13C CP/MAS NMR. Macromolecules 1991, 24, 651−658.(34) Yamane, S.; Shinzawa, H.; Ata, S.; Suzuki, Y.; Nishizawa, A.;Mizukado, J. A Thermal Oxidative Degradation Study of TriallylIsocyanurate Crosslinking Moietry in Fluorinated Rubber by Two-Dimensional Infrared Correlation Spectroscopy. Vib. Spectrosc. 2018,98, 30−34.(35) Robbins, Z.; Bodnar, W.; Zhang, Z.; Gold, A.; Nylander-French, L. A. Trisaminohexyl Isocyanurate, a Urinary Biomarker ofHDI Isocyanurate Exposure. J. Chromatogr. B: Anal. Technol. Biomed.Life Sci. 2018, 1076, 117−129.(36) Dorne, J. L.; Doerge, D. R.; Vandenbroeck, M.; Fink-Gremmels, J.; Mennes, W.; Knutsen, H. K.; Vernazza, F.; Castle, L.;Edler, L.; Benford, D. Recent Advances in the Risk Assessment ofMelamine and Cyanuric Acid in Animal Feed. Toxicol. Appl.Pharmacol. 2013, 270 (3), 218−229.(37) Johanson, G.; Kristjansson, V.; Savolainen, K.; Skaug, V. TheNordic Expert Group for Criteria Documentation of Health Risks fromChemicals. 128: Triglycidyl Isocyanaurate; National Institute forWorking Life: 2001; p 18.

Macromolecules Article

DOI: 10.1021/acs.macromol.8b01925Macromolecules 2018, 51, 9078−9087

9087