Accepted Manuscript Graphene oxide reduces the hydrolytic degradation in polyamide-11 Samuel J.A. Hocker, Natalie Hudson-Smith, Patrick Smith, Chris Komatsu, Laura R. Dickinson, Hannes C. Schniepp, David E. Kranbuehl PII:
S0032-3861(17)30809-1
DOI:
10.1016/j.polymer.2017.08.034
Reference:
JPOL 19936
To appear in:
Polymer
Received Date: 15 June 2017 Revised Date:
11 August 2017
Accepted Date: 14 August 2017
Please cite this article as: Hocker SJA, Hudson-Smith N, Smith P, Komatsu C, Dickinson LR, Schniepp HC, Kranbuehl DE, Graphene oxide reduces the hydrolytic degradation in polyamide-11, Polymer (2017), doi: 10.1016/j.polymer.2017.08.034. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
3
Samuel JA Hocker,a,c Natalie Hudson-Smith,b Patrick Smith,b Chris Komatsu,b Laura R. Dickinson,a
4
Hannes C. Schniepp,a and David E. Kranbuehlb*
5
a
Department of Applied Science, The College of William & Mary, Williamsburg VA
6
b
Department of Chemistry, The College of William & Mary, Williamsburg VA
7
c
8
E-mail:
[email protected]
9
Phone: 1-757-221-2542
11
SC
M AN U
Advanced Materials and Processing Branch, NASA Langley Research Center, Hampton, VA
Abstract Graphene
oxide
TE D
10
RI PT
2
Graphene Oxide Reduces the Hydrolytic Degradation in Polyamide-11
1
(GO)
was
incorporated
into
polyamide-11 (PA11)
via
in-situ
polymerization. The GO-PA11 nano-composite had elevated resistance to hydrolytic
13
degradation. At a loading of 1 mg/g, GO to PA11, the accelerated aging equilibrium molecular
14
weight of GO-PA11 was higher (33 and 34 kg/mol at 100 and 120 ˚C, respectively) compared to
15
neat PA11 (23 and 24 kg/mol at 100 and 120 ˚C, respectively). Neat PA11 had hydrolysis rate
16
constants (kH) of 2.8 and 12 (×10-2 day-1) when aged at 100 and 120 ˚C, respectively, and re-
17
polymerization rate constants (kP) of 5.0 and 23 (×10-5 day-1), respectively. The higher
18
equilibrium molecular weight for GO-PA11 loaded at 1 mg/g was the result of a decreased kH,
19
1.8 and 4.5 (×10-2 day-1), and an increased kP, 10 and 17 (×10-5 day-1) compared with neat PA11
20
at 100 and 120 ˚C, respectively. The decreased rate of degradation and resulting 40 % increased
21
equilibrium molecular weight of GO-PA11 was attributed to the highly asymmetric planar GO
AC C
EP
12
1
ACCEPTED MANUSCRIPT
nano-sheets that inhibited the molecular mobility of water and the polymer chain. The
23
crystallinity of the polymer matrix was similarly affected by a reduction in chain mobility during
24
annealing due to the GO nanoparticles’ chemistry and highly asymmetric nano-planar sheet
25
structure.
26
Introduction
RI PT
22
Polyamide-11 (PA11) is a widely-used engineering polymer. PA11 comprises the pressure
28
sheath liner in contact with the production fluid at elevated temperature and pressure in flexible
29
hoses for underwater transport of gas, oil, and crude to offshore platforms. As such, PA11 fulfills
30
a vital role in the global economy. Degradation and failure of these hoses have far-reaching
31
financial and environmental implications. Currently, the molecular weight of PA11 is used to
32
predict the operational lifetimes of the PA11 liner 1,2 where hydrolysis of PA11 is the mechanism
33
for degradation in anaerobic environments such as the condition of underwater crude oil risers.
34
Improving the properties of PA11 as a pressure sheath motivates research on the hydrolytic
35
degradation of graphene oxide (GO) loaded PA11: GO-PA11 polymer nano-composite.
TE D
M AN U
SC
27
Polymer nano-composites have become high potential alternatives to traditionally filled
37
polymers or polymer blends 3. Since Usuki, et al. 4, polymer nano-composites have shown
38
valuable performance property enhancements such as stiffness and thermal management at much
39
lower percent mass loadings than often used glass and carbon fiber micron scale fillers. This
40
lower percent mass loading can lead to cheaper processing and novel material solutions to
41
engineering problems.
AC C
EP
36
42
Graphene is an atomic single layer of carbon, a 2-D carbon-honeycomb crystal structure
43
with remarkable properties. It would be advantageous if graphene’s properties; an extremely
44
high modulus of 1 terapascal (TPa), an ultimate strength of 130 gigapascal (GPa), a gas 2
ACCEPTED MANUSCRIPT
45
impermeable honeycomb network, and superior electrical and thermal conductivity, 5,6 -could be
46
lent to polymers by means of creating graphene-polymer nano-composites 3,7-10. The challenge is
47
to disperse graphene as single nano-sheets into polar solvents, then into the polymer matrix. A wide interest in graphene’s exceptional strength involves the analogous compound GO.
49
GO has nearly the same strength and barrier properties as pristine graphene, given the very high
50
aspect ratio planar honeycomb structure. GO has the same honeycomb structure as graphene with
51
oxygen-containing functional groups on the surface of the planar carbon. These oxygen
52
containing functional groups are ketones, 6-membered lactol rings, alcohols, epoxides, and
53
hydroxyl groups 11. As a result, GO is a polar hydrophilic material compared to graphene, which
54
is nonpolar and hydrophobic. GO, unlike graphene, can be dispersed into water and solvents
55
commonly used in polymer precursor resins or mixtures to create GO-polymer nano-composites.
56
Also, the oxide groups on the GO surface can be reacted with functional groups for improved
57
dispersion and incorporation into a polymer matrix.
58
TE D
M AN U
SC
RI PT
48
Several methods of making graphitic oxide have been developed, starting with Brodie 13,14
, Hofmann
16
, Hummers
17
1859, modified Brodie
60
and most recently Tour
61
chemistry: in depth studies and reviews have been done illustrating such differences
62
Importantly, synthesis of graphitic oxide is suitable for industrial scale GO production 21,22.
, modified Hummers
in 18
,
in 2010. Each synthesis route results in a different graphitic oxide 11,19,20
.
AC C
EP
19
, Staudenmaier
15
59
12
8
63
GO-polymer nano-composites are easily made by exfoliating graphitic oxide into GO
64
nanosheets using a compatible solvent in-situ with the polymer. Many GO-polymer nano-
65
composites have been made using polyurethanes, polyimides, polyamides, and other polymers
66
3,7-10
67
properties of the polymer, to result in significantly improved performance properties 23-44.
. Well dispersed GO nanoparticles in a polymer can change the chemical and morphological
3
ACCEPTED MANUSCRIPT
68
In previous research on GO-PA11, Jin, et al.
39
extruded thermally reduced GO into
commercial PA11 at loadings ranging from 1 to 30 mg/g reduced GO-PA11. They showed that
70
water vapor and oxygen permeation resistance of PA11 films were increased by 49 % at a
71
1 mg/g loading and reported on the tensile properties of functionalized graphene loaded PA11.
72
Yuan, et al. 45 prepared 1.25 to 5 mg/g GO-PA11 by in-situ melt polycondensation and discussed
73
improved toughness and correlated changes in crystal structures that were formed when GO was
74
loaded into PA11.
SC
75
RI PT
69
In this work, the hydrolytic degradation of GO-PA11 and the effect GO had on the PA11 matrix crystallization was explored.
The large immobile nano thin GO sheets and the
77
intermolecular interaction between the GO’s surface C=O groups with the polyamide’s N-H
78
groups significantly reduced molecular mobility in the GO-polymer and resulted in a reduction
79
in the rate of crystallization and most importantly the rate of degradation by hydrolysis.
80
Methods and Measurements
TE D
M AN U
76
GO was incorporated into PA11 via in-situ melt condensation polymerization (GO-PA11).
82
Accelerated aging of PA11 and GO-PA11 was performed by immersing 0.3 mm thick × 1 cm2
83
samples in 15 mL of deionized water at elevated temperatures of 100 and 120 ˚C. The extent of
84
hydrolytic degradation was measured by monitoring the change in mass averaged molecular
85
weight (Mm) of PA11 over time. For PA11, a hydrolytic aging equilibrium molecular
86
weight (Mme) was established after extended aging times in anaerobic conditions.
AC C
87
EP
81
GO was synthesized using Hummers method
17
for a C/O ratior of 2.3, measured by
88
Galbraith laboratories via elemental analysis.
89
ultrasonication (Cole-Parmer EW-04711-40 homogenizer 750-Watt) for 30 minutes. The
90
resulting solution of GO in water was filtered using G4 glass fiber filter paper (Fisher Scientific) 4
The GO was dispersed in water by
ACCEPTED MANUSCRIPT
to remove unreacted graphitic particles. The GO particle water dispersion was measured by mass
92
to have a concentration of 0.3 mg/g. A Laurell: WS-650Sz-6NPP-Lite spin processor was used to
93
apply a single layer of GO particles to V5 cleaved sheet of mica (Ted Pella). A NT-MDT
94
NTEGRA atomic force microscope was used to characterize the dimensions of the GO sheets.
95
The sheet sizes in the collected filtered GO/water dispersion were observed to be 0.3-1 µm
96
laterally and 1 nm in height, Figure 1.
97 98
Figure 1: AFM height image (from GO-water dispersion) distributed on mica.
TE D
M AN U
SC
RI PT
91
Dry 11-aminoundecanoic acid powder, Sigma Aldrich, was mixed with the filtered
100
GO/water dispersion at a concentration of 1 and 5 mg of GO per gram of powder. Water was
101
added to the mixture for a total volume of 15 mL. An ultrasonication tip was submerged in the
102
mixture and sonication at 35 % of maximum power proceeded for 30 minutes with continuous
103
stirring at 300 rpm and a maximum temperature of 30 ˚C.
AC C
EP
99
104
The homogeneous golden-brown mixture of GO-monomer settled with a clear water
105
supernatant. The mixture was poured into a Teflon® lined vessel and placed at 100 ˚C in an oven
106
for 60 minutes. The dried powder was uniformly brown. The GO-monomer powder was placed
107
under argon at 235 ˚C for 4 hours of melt condensation static polymerization, and thereby GO5
ACCEPTED MANUSCRIPT
108
PA11 was made. PA11 fabrication followed the same method. For convention, a concentration of
109
1 mg/g GO-PA11 was termed 1-GO-PA11 and 5 mg/g GO-PA11 was 5-GO-PA11. Previous work had found that GO will thermally reduce carbon oxygen ratio in-situ in a
111
polymer matrix 46. GO was heated under nitrogen gas to 250 ˚C and found to have a C/O ratio of
112
5.3 via elemental analysis. GO was assumed to have increased in hydrophobicity by reducing
113
within PA11 in-situ during polymerization.
RI PT
110
Aging samples were prepared by slowly hot pressing the PA11 or GO-PA11 between two
115
Teflon® surfaces at pressures up to 16 MPa at 250 ˚C under argon. The system was allowed to
116
cool under flowing argon to room temperature. The approximate dimension of each film was 60
117
cm2 with a 300 µm thickness. Coupons of 1 cm2 by 300 µm thick were cut from the 60 cm2 by
118
300 µm thick PA11 or GO-PA11 films. For each material tested, ten coupons were immersed in
119
deionized water over a period of 3 - 4 months at elevated temperature to accelerate aging. Two
120
independent sets were studied at the temperatures of 100 and 120 ˚C. High-pressure rated #40
121
glass tubes with Teflon® plugs (Ace Glass) were used as containment vessels. Before sealing the
122
pressure tubes, dissolved oxygen was removed by sparging the water with argon in-situ until the
123
measured oxygen concentration dropped below 50 ppm (Oakton Waterproof DO 450 Optical
124
Dissolved Oxygen Portable Meter). Samples were removed periodically from each pressure
125
tube, and then each pressure tube was degassed and resealed for continued aging of the
126
remaining samples.
AC C
EP
TE D
M AN U
SC
114
127
Mass average molecular weight (Mm) was measured by in-line Shodex size exclusion
128
chromatography columns (HFIP-LG, HFIP-805, and HFIP-803) paired with a Wyatt miniDAWN
129
light scattering detector and Wyatt Optilab 803 dynamic refractive index detector. The
130
miniDAWN had a laser wavelength of λ = 690 nm and three detection angles θ of 45, 90, and
6
ACCEPTED MANUSCRIPT
131
135˚. The solvent used to dissolve the PA11 samples was 1,1,1,3,3,3-hexafluroisopropan-2-ol
132
(HFIP) doped with 0.05 M potassium tri-fluroacetate (KTFA) salt to remove nylon
133
agglomerations
134
helium sparge under atmospheric pressure. The polymer solution concentration was 2 mg/mL
135
(PA11/KTFA-HFIP). A 100 µL aliquot of solution was injected for measurement. A
136
0.235 mL/g determined the concentration of the PA11 at each fraction per the refractive index
137
measurement. The
138
concentrations. The measured Mm values had a 6% error margin, determined by the standard
139
deviation divided by the mean for six consecutive Mm measurements on a single solution of
140
PA11.
dn
/dc of
/dc was experimentally measured for PA11/KTFA-HFIP using the solution
SC
dn
RI PT
. The KTFA doped HFIP was degassed in the pump reservoir via constant
M AN U
47
The GO-PA11/KTFA-HFIP solutions were darkened by the presence of GO. All visible GO
142
in each solution was removed by a 0.45 µm porosity polytetrafluoroethylene filter. Additionally,
143
the 0.2 µm stainless steel frit column guard on the SEC-MALLS was unaffected by injections of
144
clear solutions that were passed through the 0.45 µm filter. The unaffected 0.2 µm frit indicated
145
that the polymer solutions being measured on the SEC-MALLS were free of GO contamination.
TE D
141
Differential scanning calorimetry (DSC) was used to measure the heat of fusion and re-
147
crystallization in PA11. The DSC (TA Instruments Q20) was calibrated using indium. A ramp
148
rate of 3 ˚C/min was used to heat, cool, and re-heat the samples under nitrogen from 40 to
149
220 ˚C. The integration limits were 140 to 200 ˚C for the heating ramps and 130 to 190 ˚C for
150
the cooling ramp. The root of the peak onset maximum slope with the peak baseline set to zero
151
was the fusion or crystallization onset temperature.
AC C
EP
146
152
Optical microscopy images were taken with an IX71 Olympus Inverted microscope and a
153
LUCPLFLN, 40x (0.6 NA) objective. Transmission bright field and cross polarized light images
7
ACCEPTED MANUSCRIPT
154
were taken. For PA11 and 1-GO-PA11, optical glass microscope slides were prepared by melt
155
pressing the PA11 or GO-PA11 for 30 seconds at 250 ˚C under argon, then ambient cooling to
156
room temperature under argon.
157
Microtome to obtain a sample less than 150 nm, for optical transmission.
RI PT
The 5-GO-PA11 was sliced using a Leica Ultracut UCT
Attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR) was used
159
to probe the chemical functional groups and intermolecular interactions. Spectra were taken
160
using an IR Tracer-100 Shimadzu FTIR with a MIRacle 10 Single Reflectance ATR accessory at
161
a resolution of 2 cm-1 from 600 to 4000 cm-1 and 32 scans were averaged. The spectra were
162
adjusted to a common baseline and normalized to the 2851 cm-1 peak. To analyze the carbonyl
163
oxygen peak, the work of Skrovanek, et al.
164
peak was de-convoluted using MagicPlot software, and the relative concentration of free and
165
hydrogen bonded amide oxygen groups were measured at 1684, 1654, and 1635 cm-1. The
166
heights of the de-convoluted peaks were used to compare the relative concentration of hydrogen
167
bonded carbonyl oxygen groups. The half-height peak width of the 3304 cm-1 N-H absorption
168
was also used to probe the hydrogen bonding behavior of the amide hydrogen.
M AN U
SC
158
was followed. The 1635 cm-1 carbonyl oxygen
TE D
48
The equilibrium molecular weight (Mme) hydrolysis rate constant (kH), and solid state
170
polymerization (kP) rate constant were determined during accelerated molecular weight
171
degradation by fitting the literature mathematical model of amide hydrolysis to the measured Mm
172
per aging time
173
hydrolysis kinetics described by Equation 2 and the associated first order ordinary differential
174
Equation 7. The values kH and kP were fit to the changing values of molecular weight over time
175
in days and reflected the effect of the aqueous aging environment on the hydrolysis process.
AC C
EP
169
49-51
. The previously published model was derived from the condensation-
8
ACCEPTED MANUSCRIPT
In the model, Equation 10, was derived from the hydrolysis kinetics of Equation 1 - 3, and
177
the associated first order ordinary differential solution, Equation 9. The values kH and kP were fit
178
to the changing values of molecular weight over time via the relationship described in Equation
179
10 where at was the average number of amide bonds at time t. Subscript “0” referred to the
180
starting value, “e” referred to the equilibrium value, and “t” referred to the value associated at
181
time t.
RI PT
176
Equation 1
M AN U
183
SC
182
184 185 186
Equation 2
Equation 3
190 191 Equation 4
193 194
Equation 5
195 196 197 198 199 200
AC C
192
EP
189
TE D
187 188
Equation 6
Equation 7
201 202 9
ACCEPTED MANUSCRIPT
203
Equation 8
204 205 Equation 9
RI PT
206 207 208
Equation 10
SC
209
M AN U
210 211
The hydrolysis degradation model in terms of the mass average number of amide bonds at
212
time t (at) was defined by Equation 10. The derivation of Equation 10 started with the steady
213
state approximation at equilibrium, Equation 3
214
assumed to be constant and thereby Equation 6 was used to form the first order ordinary
215
differential Equation 7. At the steady state, equilibrium, dx/dt was defined to be zero and thereby
216
rearranging Equation 7 in terms of of kP resulted in Equation 8. Equation 9 was the solution of
217
the differential equation that resulted from substituting Equation 8 into Equation 7. Equation 10
218
was Equation 9 rearranged in terms of at. Equation 5 established that at was proportional to the
219
measured Mm by the molecular weight of the monomer. Using Equation 5, the changes in the
220
Mm data over time were fit to Equation 10 to determine kH using a non-linear least squares fit.
221
The model fit determined both the Mme and kH. Then, the best fit of Mme and kH were used in
222
Equation 8 to calculate the re-polymerization rate constant kP.
.
Then the concentration of water was
AC C
EP
TE D
1,50
223
10
ACCEPTED MANUSCRIPT
224
Results and Discussion The Mme of PA11 is very important to the performance properties of PA11. If the Mme is
226
above the ductile-brittle transition Mm then PA11 can retain its ductility 1. The Mme of 1-GO-
227
PA11 was 10 kg/mol larger than that obtained for PA11 at both 100 and 120 ˚C, respectively.
228
This was an impressive increase of 40 % at both temperatures.
RI PT
225
Using Equation 10 and a non-linear least squares fit, the kH and solid state kP rate constants
230
were determined and listed in Table 1. Figure 2 and Table 1 showed that the kH of 1-GO-PA11
231
was reduced by a factor of 2 and the Mme of 1-GO-PA11 was increased by 40 % at both
232
temperatures. At 120 ˚C, both kH and kP were reduced for 1-GO-PA11 compared with PA11.
233
The higher loading, 5-GO-PA11 showed no change in the rate of hydrolytic degradation
234
compared with PA11 at both temperatures. The absence of this effect at the higher concentration,
235
5-GO-PA11, was attributed to a poor nano-particle dispersion as shown in Figure 3. The optical
236
images of 5-GO-PA11 showed that GO was highly agglomerated, the GO sheets were in contact
237
with each other, Figure 3. The 1-GO-PA11 lacks sharp features in the optical micrograph,
238
compared with the defined sharp features sized from 3 µm to 15 µm in 5-GO-PA11, GO
239
agglomerates. This agglomeration resulted in no effect on the Mme compared with the finely
240
dispersed GO in the 1-GO-PA11 nano-composite.
AC C
EP
TE D
M AN U
SC
229
11
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 2: Dependence of Mm on time when aged in water at 100 and 120 ˚C for PA11, 1-GO-PA11, and
243
5-GO-PA11. Mme of 1-GO-PA11 was approximately 10 kg/mol higher than PA11 and 5-GO-PA11 when
244
aging in water at 100 ˚C and 120 ˚C, respectively.
M AN U
241 242
245
Table 1: Rate constants and Mme for GO-PA11 aged in water; determined using Equation 8 and a non-
247
linear least squares fit.
248
AC C
120
PA11 1-GO-PA11 5-GO-PA11 PA11 1-GO-PA11 5-GO-PA11
EP
Temperature, °C 100
TE D
246
kH,×10-2 day-1 2.8 ±0.5 1.8 ±0.5 2.9 ±0.6 12 ±2 4.5 ±1.1 13 ±2
12
kP, ×10-5 day-1 5.0 ±1.5 10.0 ±4.0 6.8 ±2.4 23 ±7 17 ±5 19 ±5
Mme, kg/mol 24 ±3 34 ±4 24 ±3 24 ±3 34 ±3 23 ±3
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 3: Transmission optical microscopy of PA11 bright field (a), cross-polarized (b), 1-GO-PA11
251
bright field 6c), and 5-GO-PA11 bright field (d). Each image is sized at 60 x 60 µm.
AC C
252
EP
249 250
253
PA11 is a semicrystalline polymer. Amide hydrolysis in PA11 is understood to occur in the
254
amorphous regions rather than the impermeable crystalline domains. A key component of the
255
crystalline formation is hydrogen bonding. Hydrogen bonding in the PA11 matrix is an
256
intermolecular interaction between the amide hydrogen and the carbonyl oxygen of a
257
neighboring amide bond. Hydrogen bonding also occurs between nearby amide bonds in the
13
ACCEPTED MANUSCRIPT
258
amorphous regions. The hydrogen bonding between adjacent amide groups is responsible for the
259
higher melting temperatures of polyamide crystalline structures versus a non-polar
260
semicrystalline polymer such as polyethylene. Commercially prepared PA11 is plasticized with N-n-butylbenzenesulfonamide (BBSA) and 49,52
RI PT
261
shows a similar increase in the Mme over neat PA11
. The BBSA increases the amorphous
263
content in PA11 by disrupting hydrogen bonding. BBSA inserts itself to replace the amide/amide
264
interaction with a sulfonamide/amide interaction. The C-N-H – O=C hydrogen bond is replaced
265
by a stronger S-N-H – O=C hydrogen bond, and a weaker C-N-H – O=S hydrogen bond.
266
Effectively, BBSA blocks intermolecular hydrogen bonding between neighboring amide bonds
267
for lower crystallinity and increased amorphous content 53.
M AN U
SC
262
268
In the 1-GO-PA11, there were three species involved in the amide hydrolysis reaction:
269
polyamide chains, GO, and water. The elementary chemical mechanisms for chain scission were
270
identified: base catalyzed
271
energies had been determined to be 21, 31, and 99 kJ/mol for the base catalyzed, acid catalyzed,
272
and water assisted mechanisms of amide hydrolysis, respectively
273
found activation energies of 66, 78, and 147 kJ/mol. Thus, given equivalent concentrations of
274
OH- and H+ the base catalyzed pathway was favored; and the water assisted hydrolysis of an
275
amide bond was the least favored pathway.
277
59-62
, and water assisted
65
63,64
. The activation
. Computationally, Zahn
EP
TE D
, acid catalyzed
AC C
276
54-58
Equation 11: Elementary chemical mechanism of resonance stabilization of the amide bond.
278
14
63
ACCEPTED MANUSCRIPT
Equation 12: Base catalyzed elementary chemical mechanism of amide hydrolysis.
SC
RI PT
279
285
286 287 288
TE D EP
283 284
Equation 13: Acid Catalyzed elementary chemical mechanism of amide hydrolysis.
Equation 14: Elementary chemical mechanism for water assisted Hydrolysis of the amide bond.
AC C
282
M AN U
280 281
Equation 11 shows the initial resonance stabilization of the amide bond. stabilization chemically shortens and strengthens the amide bond.
15
Resonance
ACCEPTED MANUSCRIPT
54
289
The base catalyzed hydrolysis mechanisms were drawn in Equation 12. Zahn
290
computationally studied the mechanism of base catalyzed hydrolysis of the amide bond, and
291
found that the base catalyzed Pathway II was favored. Zahn
292
reaction led to the simultaneous protonation of the amide nitrogen and deprotonation of the
293
hydroxyl group. Since the amide anion was a poorer leaving group than an alkoxide ion, the base
294
catalyzed hydrolysis the proton transfer to the amide anion was considered the rate limiting step
295
66,67
also found that a proton transfer
RI PT
54
SC
.
The four-step acid catalyzed hydrolysis mechanism was drawn in Equation 13. The first step
297
was the protonation of the carbonyl oxygen of the amide bond. Zahn 68 found in a computational
298
study that the second step, nucleophilic attack of a water molecule to the carbon atom of the
299
amide group, was the rate determining step. The third step was the intermediate de-protonation
300
of the carbonyl oxygen and protonation of the amide nitrogen, quickly followed by the
301
dissociation of the protonated amine and the carboxylic acid.
TE D
M AN U
296
The water assisted hydrolysis mechanism was drawn in Equation 14. The water assisted
303
mechanism involved two water molecules that acted in a concerted process and utilized a
304
Grotthuss mechanism of proton hopping through hydrogen bonding. The first step and rate
305
determining step was the nucleophilic attack of the amide carbon and simultaneous protonation
306
of the amide nitrogen 63.
AC C
EP
302
307
The crystalline regions in PA11 were impermeable to water molecules: therefore, hydrolysis
308
took place predominately in the molecularly mobile amorphous regions. Unaged PA11 and GO-
309
PA11 had the same crystalline content and therefore no change in hydrolysis was expected.
310
However, the well dispersed 1-GO-PA11 showed unexpected decreases in the kH at both
311
temperatures and a 40 % higher Mme than PA11, Figure 2.
16
ACCEPTED MANUSCRIPT
From the amide bond hydrolysis elementary mechanisms for base hydrolysis Equation 12,
313
acid hydrolysis Equation 13, and water assisted hydrolysis Equation 14, it was clear that both the
314
availability of water molecules and hydrogen bonding were required for amide hydrolysis. Thus,
315
the BBSA’s disruption of inter-chain hydrogen bonding also inhibits hydrolysis. Changes in the
316
hydrogen bonding behavior was another possible explanation for GO inhibiting the hydrolysis
317
process in 1-GO-PA11, in addition to its large immobile nano-sheet structure that inhibited
318
polymer chain and water mobility. Skrovanek, et al.
48
SC
319
RI PT
312
first studied the use of FTIR to detect hydrogen bonding interactions 48
found de-convoluting the 1635 cm-1 peak of the
within a PA11 polymer. Skrovanek, et al.
321
carbonyl oxygen revealed three peaks that could be characterized at 1635, 1654, and 1684 cm-1.
322
The 1635 cm-1 sub-peak was a lower energy carbonyl vibration correlated to aligned hydrogen
323
bonding. The 1654 cm-1 peak was un-aligned amorphous hydrogen bound carbonyl oxygens. The
324
1684 cm-1 peak was correlated with carbonyl oxygen groups that were free of hydrogen bonding.
325
The amide bond 1635 cm-1 carbonyl oxygen peak in PA11, 1-GO-PA11 and 5-GO-PA11 was
326
deconvoluted based on the results of Skrovanek, et al.
327
carbonyl oxygen peaks for PA11 and GO-PA11 from deconvolution of the 1635 cm-1 amide
328
C=O. Further details were provided in the supplementary information.
TE D
EP
AC C
329
M AN U
320
17
48
. No difference was observed at the
330 331
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 4: ATR-FTIR spectra of PA11 and 1-GO-PA11, and aged PA11 and aged 1-GO-PA11.
Infrared spectroscopy was used to probe the PA11 inter-molecular interactions, and
333
hydrogen bonding behavior at the amide bonds. Figure 4 shows the FTIR spectra for PA11 and
334
1-GO-PA11, pristine and aged. The characteristic peaks were: 1158 cm-1, an interaction of the N-
335
H stretch and O=C-N deformation; 1539 cm-1, amide II C-N stretch; 1635 cm-1, amide I C=O
336
stretch; 2851 cm-1, symmetric C-H aliphatic vibration; 2919 cm-1, asymmetric C-H aliphatic
337
vibration; and 3304 cm-1, amide III N-H stretch.
EP
AC C
338
TE D
332
18
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
339 340
Figure 5: The peak widths centered at 3304 cm-1 for PA11, 1-GO-PA11, and 5-GO-PA11; and grouped
341
by unaged and aged at 120 ˚C for 90 and 120 days, respectively. The 3304 cm-1 peak widths are
342
associated with the distribution of hydrogen bonding orientations of the amide bond hydrogen.
Figure 5 compared the half height peak widths of the 3304 cm-1 amide (N-H) stretching peak
344
for the PA11 and 1-GO-PA11, pristine and aged. The peak widths were associated with the
345
distribution of hydrogen bonding orientations of the amide bond hydrogen. With aging, PA11
346
showed a decline in the distribution of hydrogen bonding orientations; from 54 to 34 ± 2 cm-1.
347
The 1-GO-PA11 showed less of a decline in the 3304 cm-1 peak widths; from 54 to 39 ± 2 cm-1.
348
The broader distribution of hydrogen bonding orientations in the GO-PA11 had also been
349
correlated previously to the lower overall crystallinity content 48.
AC C
EP
TE D
343
350
In aged 1-GO-PA11 the N-H stretch had a broader distribution than aged PA11, Figure 5.
351
Further 1-GO-PA11 had an increased intensity at the 1158 cm-1 peak, an interaction of the N-H
352
stretch and O=C-N deformation. These results showed that the GO with C=O groups present on
353
the surface
69
were interacting with the N-H amide hydrogen of polyamide. This interaction
19
ACCEPTED MANUSCRIPT
354
reduced the molecular mobility of the PA11 chain and thereby reduced the rate of hydrolysis and
355
rate of the crystallization. A transmission optical microscope was used in cross-polarization mode and the
357
birefringence of the crystalline regions within PA11 and 1-GO-PA11 was observed. In
358
transmission mode, the light passed through the sample. With cross polarizers, the birefringent
359
and therefore visible crystalline regions of PA11 and 1-GO-PA11 were imaged. Figure 6 showed
360
that well dispersed GO within the polymer matrix altered the crystalline morphology of PA11.
361
The 1-GO-PA11 showed less birefringence than PA11. The aged PA11 showed the most
362
birefringence, and banded spherulites, broader bands of birefringence, for both the 100 and 120
363
˚C with widths of 0.9 ± 0.3 and 1.0 ± 0.2 µm, respectively. The banded spherulites were
364
associated with bundling of laminar stacks and higher crystallinity
365
laminar stacks in aged PA11 were absent in the aged 1-GO-PA11 optical images. The bundles of
366
laminar stacks in the 100 and 120 ˚C aged 1-GO-PA11 appeared much thinner with widths of 0.7
367
± 0.2 and 0.7 ± 0.1 µm, respectively. Zhang, et al.
368
montmorillonite-PA11 nano-composites. These smaller crystalline domains agreed with the
369
lower crystallinity DSC results for 1-GO-PA11 in Table 2.
AC C
EP
TE D
M AN U
SC
RI PT
356
20
72
70,71
. The larger bundled
reported a similar crystalline behavior in
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 6: Cross polarized transmission optical microscopy images of 100 and 120 ˚C 80-90 days aged
372
PA11 and 1-GO-PA11. Each image is sized at 60 x 60 µm.
EP
370 371
When fully dissociated and dispersed, the GO nanoparticles acted as large immobile 2-D
374
nano-thin barriers within the PA11 matrix. Consequently, they were predicted to have inhibited
375
molecular mobility within the polymer. As such, the results suggested that the GO nano-sheets
376
reduced molecular mobility for both PA11 chains and water for less accessible hydrolysis sites.
377
GO’s effect on the rate of crystallization was characterized in PA11 and GO-PA11 - unaged and
378
aged using DSC, which provided evidence for GO’s role in reducing chain mobility.
AC C
373
21
ACCEPTED MANUSCRIPT
45,70,73-81
Nanoparticles had been known to affect the crystalline morphology of PA11
. Figure 7
380
showed the DSC curves of the un-aged and 80-90 days aged samples in three columns from left
381
to right to indicate the heat (fusion), cool (crystallization), and heat (fusion) curves. The peak
382
integration values of specific heat of fusion (∆Hfus) and crystallization (∆Hc) were listed in Table
383
2.
RI PT
379
In Table 2, a lower total ∆Hfus indicated a decrease in the crystallinity of the GO-PA11 after
385
heat, cool, and re-heat ramps. Peak and onset temperatures were respectively tabulated in Table 3
386
and Table 4 for the fusion and crystallization of PA11 and GO-PA11. The effect of the previous
387
thermal history on the semi-crystalline behavior of PA11 was revealed in the first heat ramp. The
388
first heat ramp both revealed the previous thermal history on the semi-crystalline behavior of
389
PA11 and melted the PA11 to erase that previous thermal history. The 3 ˚C/min controlled
390
cooling ramp measured the affect that GO has upon the crystallization behavior of PA11, both
391
nucleation and growth. The second heating ramp compared how the total crystallization growth
392
was affected by the GO loading.
TE D
M AN U
SC
384
During aging at 100 and 120 ˚C the PA11 matrix was annealed resulting in an increase in
394
crystallinity. Quantified by the peak total area, the first heating ramp revealed the extent of
395
annealing for the PA11 and GO-PA11 samples via their ∆Hfus. Before aging, PA11 had 25%
396
crystallinity, using 189 J/g as 100% crystalline PA11
397
crystallinity increased to 34 and 48% at 100 and 120 ˚C; a relative increase of 36 and 92%,
398
respectively. The 1-GO-PA11 started with a crystallinity of 25% and had lower relative increases
399
of 12 and 76% at the same annealing conditions.
AC C
EP
393
22
71
. Upon aging and annealing PA11 the
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 7: DSC scans for un-aged, 100 ˚C, and 120 ˚C for 80-90 days aged PA11, 1-GO-PA11, and 5-
402
GO-PA11. The 3 columns from left to right were 1st heat ramp from 40 ˚C to 220 ˚C, followed by a cool
403
ramp to 40 ˚C, and a 2nd heat ramp to 220 ˚C. The rows top to bottom were un-aged, 100 ˚C, and 120 ˚C
404
aged samples.
AC C
405
EP
400 401
23
ACCEPTED MANUSCRIPT
406 407
Table 2: Integration values of DSC heat flow peaks from 140 to 200˚C for GO-PA11 and PA11, aged
408
(80-90 days) and un-aged samples. The error margin was 2.5 J/g per measurement and peak integration.
120 ˚C
409
Cooling Ramp ∆Hc, J/g 50.0 42.5 44.0 60.0 50.0 53.5 65.5 53.5 49.5
2nd Heat Ramp ∆Hfus, J/g 48.0 49.0 49.0 60.0 55.0 56.0 61.0 59.0 51.0
RI PT
100 ˚C
1st Heat Ramp ∆Hfus, J/g 47.5 46.5 46.0 63.0 53.0 57.5 91.0 82.0 83.0
SC
Unaged
GO Loading, mg/g 0 1 5 0 1 5 0 1 5
M AN U
Sample
410
Table 3: DSC peak temperatures for GO-PA11 and PA11, aged (80-90 days) and un-aged samples. The
411
peak temperatures for melting(Tm,fus) and re-crystallization (Tm,c) were accurate within 0.5 ˚C.
100 ˚C
412
AC C
120 ˚C
1st Heat Ramp Tm,fus, ˚C 187.5 190.0 187.5 189.5 191.0 189.5 184.5, 190.5 182.0, 187.0, 191.0 183.0, 188.5
TE D
Unaged
GO Loading, mg/g 0 1 5 0 1 5 0 1 5
EP
Sample
Cooling Ramp Tm,c, ˚C 166.5 169.5 169.0 168.5 173.0 174.0 170.0 173.5 174.5
2nd Heat Ramp Tm, fus, ˚C 187.5 188.5 187.0 189.5 190.0 187.5 190.0 191.0 187.5
413
These results showed that the GO nano-sheets hindered annealing in PA11 during aging at
414
these temperatures. The 1-GO-PA11 samples had the lowest ∆Hfus in both 100 and 120 ˚C
415
conditions and indicated a large immobilizing effect on the polymer chains. The decreased effect
416
of the 5-GO-PA11 was attributed to agglomeration and poor dispersion of 2-D nano-particles as
417
previously described.
24
ACCEPTED MANUSCRIPT
418
Table 4: DSC onset melting and formation temperatures for GO-PA11 and PA11, aged (80-90 days) and
419
un-aged samples. The onset melting(To,fus) and re-crystallization (To,c) temperatures were accurate within
420
0.5 ˚C.
120 ˚C
421
2nd Heat Ramp To,fus, ˚C 177.5, 178.5 181.0 178.0 178.0, 183.0 181.5 177.5 180.0, 184.5 182.5 171.0
RI PT
100 ˚C
Cooling Ramp To,c, ˚C 167.0 175.0 176.0 170.5 179.0 180.0 172.0 179.5 181.0
SC
Unaged
1st Heat Ramp To,fus, ˚C 175.5, 180.5 178.5, 180.0 175.0, 176.5 182.5, 183.5 179.5, 183.5 177.5, 179.5 181.0, 186.5 174.0, 177.5, 181.5 171.0
GO Loading, mg/g 0 1 5 0 1 5 0 1 5
M AN U
Sample
422
The melting temperatures (Tm,fus) in Table 3 for PA11 and GO-PA11 were indicative of the
423
thermal stability of the polymer crystalline regions. A lower pre-peak Tm,fus, 184.5 ˚C, became
424
increasingly prominent for the PA11 when aged at 120 ˚C. The 184.5 ˚C pre-peak was attributed
425
to poorly formed crystalline structures
426
pre-peak at 184.5 ˚C in the first heating ramp after aging at 120 ˚C. Conversely, the poorly
427
formed crystals were not favored in the 1-GO-PA11 and there was an additional intermediate
428
melting peak at Tm,fus 187.0 ˚C. The highest Tm,fus of 191.5 ˚C indicated that the annealed GO-
429
PA11 had the most stable crystalline structure. The increased annealed crystalline melt
430
temperature after aging suggested intermolecular attractive interactions between the GO sheets
431
and the polyamide matrix as was seen in the ATR-FTIR spectra.
. As seen in Figure 7, the PA11 had the largest
AC C
EP
TE D
71,72,77,82,83
432
After the first heating ramp on the DSC, the polymer sample melted and its thermal history
433
erased. With controlled cooling the heat of re-crystallization (∆Hc) can be characterized. The
434
decrease in the ∆Hc for GO-PA11 compared to PA11 indicated that the chain mobility was
435
inhibited by the inclusion of GO nanoparticles. 25
ACCEPTED MANUSCRIPT
The GO-PA11 had a higher onset temperature of re-crystallization during the cooling ramp
437
(To,c) and lower ∆Hc value than PA11, Table 4. The pristine PA11 had a To,c of 167 ˚C that
438
increased to 170 (+2%) and 172 (+3%) ˚C after aging at 100 and 120 ˚C. The increase in To,c
439
after aging correlated to lower molecular weights. For a polyamide, crystalline regions form
440
more quickly at lower molecular weights due to increased mobility and ease of alignment of the
441
shorter polymer chains 84. Both GO-PA11 showed a 9 ˚C higher To,c with a broader reaction peak
442
than unloaded PA11. The 1-GO-PA11 samples initially had a To,c of 175 ˚C. After aging at 100
443
˚C, the To,c increased to 179 (+2%) ˚C; and aging at 120 ˚C increased the To,c to 180 (+3%) ˚C.
444
The increased To,c of the GO-PA11 samples were evidenced that the GO nano-particles were
445
nucleating sites resulting in higher crystallization onset temperatures.
M AN U
SC
RI PT
436
In summary the large immobile GO, a 2-dimensional sheet in the amorphous phase, in
447
addition to inhibiting molecular mobility, competed for hydrogen bonding with the polyamide’s
448
amide hydrogens. Both effects inhibited hydrolysis of the amide bond. It is the decrease in chain
449
and water molecular mobility combined with hydrogen bonding interactions between GO and the
450
N-H bonds of the polymer chains that led to a lower hydrolysis rate for the well dispersed 1-GO-
451
PA11, Table 1. The effects of GO’s large sheet shape and its surface functional groups
452
interacting with the polymer chain, were similarly shown to have reduced polymer chain
453
mobility in polyimide 85.
454
Summary & Conclusion
AC C
EP
TE D
446
455
A concentration of 1 mg/g GO in PA11 produced a 40% increase over PA11’s equilibrium
456
molecular weight when aged in water at 100 and 120 ˚C, respectively. No change was observed
457
in the 5 mg/g GO-PA11 due to GO agglomeration. Fitting the previously reported kinetic model
458
to aging results showed an increase in equilibrium molecular weight, when the GO particles were 26
ACCEPTED MANUSCRIPT
459
well dispersed. This was due to a reduction in the model’s hydrolytic rate constant relative that
460
for re-polymerization. Measured by the heat of fusion, GO-PA11’s decreased crystallization indicated that the GO
462
decreased molecular mobility within the nanocomposite, specifically the polymer chain
463
molecular mobility during the annealing process.
464
nanocomposites showed a slight decrease in the amount of PA11 crystallinity prior to aging.
465
Upon aging at both 100 and 120 ˚C, the GO-PA11 had significantly less crystallinity and
466
demonstrated a reduction in the rate of annealing. After removing the thermal history by heating
467
above the melting temperature and cooling, the materials were nearly comparable and suggested
468
no effect on the crystallinity. The decrease in the annealing rate of crystallization in the aged
469
samples was attributed to GO’s large immobile nano thickness sheet structure, which inhibited
470
chain molecular mobility within the polymer nano-composite.
472
M AN U
SC
Both 1-GO-PA11 and 5-GO-PA11
ATR-FTIR spectra showed that the graphene oxide’s surface C=O groups hydrogen bond
TE D
471
RI PT
461
with the polyamide’s N-H groups. This interaction further inhibited polymer mobility. Two factors significantly reduced molecular mobility in the GO-polymer and resulted in a
474
reduction in the rate of crystallization and most importantly the rate of degradation by
475
hydrolysis: the large immobile nano thin GO sheets and the intermolecular interaction between
476
the GO’s surface C=O groups with the polyamide’s N-H groups.
478 479
AC C
477
EP
473
27
ACCEPTED MANUSCRIPT
References
480
5 6 7
8 9
10
11 12 13
14
15 16
SC
4
M AN U
3
TE D
2
Meyer, A., Jones, N., Lin, Y. & Kranbuehl, D. Characterizing and Modeling the Hydrolysis of Polyamide-11 in a pH 7 Water Environment. Macromolecules 35, 2784-2798, doi:10.1021/ma010541o (2002). API. Technical Report 17TR2, The Ageing of PA-11 in Flexible Pipes. (American Petroleum Institute, Washington, DC, 2003). Potts, J. R., Dreyer, D. R., Bielawski, C. W. & Ruoff, R. S. Graphene-based polymer nanocomposites. Polymer 52, 5-25, doi:10.1016/j.polymer.2010.11.042 (2011). Usuki, A. et al. Synthesis of nylon 6-clay hybrid. Journal of Materials Research 8, 11791184, doi:10.1557/jmr.1993.1179 (1993). Bunch, J. S. et al. Impermeable Atomic Membranes from Graphene Sheets. Nano Letters 8, 2458-2462, doi:10.1021/nl801457b (2008). Leenaerts, O., Partoens, B. & Peeters, F. M. Graphene: A perfect nanoballoon. Applied Physics Letters 93, 193107-193104, doi:10.1063/1.3021413 (2008). Dhand, V., Rhee, K. Y., Ju Kim, H. & Ho Jung, D. A Comprehensive Review of Graphene Nanocomposites: Research Status and Trends. Journal of Nanomaterials 2013, 1-14, doi:10.1155/2013/763953 (2013). Kim, H., Abdala, A. A. & Macosko, C. W. Graphene/Polymer Nanocomposites. Macromolecules 43, 6515-6530, doi:10.1021/ma100572e (2010). Mittal, V. Functional Polymer Nanocomposites with Graphene: A Review. Macromolecular Materials and Engineering 299, 906-931, doi:10.1002/mame.201300394 (2014). Yoo, B. M., Shin, H. J., Yoon, H. W. & Park, H. B. Graphene and Graphene Oxide and Their Uses in Barrier Polymers. Journal of Applied Polymer Science 131, 39628-39651, doi:10.1002/app.39628 (2014). Dreyer, D. R., Park, S., Bielawski, C. W. & Ruoff, R. S. The chemistry of graphene oxide. Chem. Soc. Rev. 39, 228-240, doi:10.1039/B917103G (2010). Brodie, B. C. On the Atomic Weight of Graphite. Phil. Trans. R. Soc. Lond. 149, 249-259, doi:10.1098/rstl.1859.0013 (1859). Kuila, T., Khanra, P., Mishra, A. K., Kim, N. H. & Lee, J. H. Functionalizedgraphene/ethylene vinyl acetate co-polymer composites for improved mechanical and thermal properties. Polymer Testing 31, 282-289 (2012). Terrones, M. et al. Interphases in Graphene Polymer-based Nanocomposites: Achievements and Challenges. Advanced Materials 23, 5302-5310, doi:10.1002/adma.201102036 (2011). Staudenmaier, L. Verfahren zur Darstellung der Graphitsäure. Berichte der deutschen chemischen Gesellschaft 31, 1481-1487, doi:10.1002/cber.18980310237 (1898). Hofmann, U. & König, E. Untersuchungen über Graphitoxyd. Zeitschrift für anorganische und allgemeine Chemie 234, 311-336, doi:10.1002/zaac.19372340405 (1937).
EP
1
AC C
482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519
RI PT
481
28
ACCEPTED MANUSCRIPT
22 23
24
25 26
27
28 29
30
31
32
RI PT
21
SC
20
M AN U
19
TE D
18
Hummers Jr, W. S. & Offeman, R. E. Preparation of Graphitic Oxide. Journal of the American Chemical Society 80, 1339, doi:10.1021/ja01539a017 (1958). Kovtyukhova, N. I. et al. Layer-by-Layer Assembly of Ultrathin Composite Films from Micron-Sized Graphite Oxide Sheets and Polycations. Chemistry of Materials 11, 771778, doi:10.1021/cm981085u (1999). Marcano, D. C. et al. Improved Synthesis of Graphene Oxide. ACS Nano 4, 4806-4814, doi:10.1021/nn1006368 (2010). Szabó, T., Talyzin, A. V., You, S. & Luzan, S. M. Effect of synthesis method on solvation and exfoliation of graphite oxide. Carbon 52, 171-180, doi:10.1016/j.carbon.2012.09.018 (2013). Zhao, J., Pei, S., Ren, W., Gao, L. & Cheng, H.-M. Efficient Preparation of Large-Area Graphene Oxide Sheets for Transparent Conductive Films. ACS Nano 4, 5245-5252, doi:10.1021/nn1015506 (2010). Robinson, J. T. et al. Wafer-scale Reduced Graphene Oxide Films for Nanomechanical Devices. Nano Letters 8, 3441-3445, doi:10.1021/nl8023092 (2008). Adelnia, H., Gudarzi, M. M. & Gavgani, J. N. Intumescent flame retardant polyurethane/reduced graphene oxide composites with improved mechanical, thermal, and barrier properties. Journal of Materials Science 49, 243-254, doi:10.1007/s10853013-7698-6 (2013). Kim, H., Miura, Y. & Macosko, C. W. Graphene/Polyurethane Nanocomposites for Improved Gas Barrier and Electrical Conductivity. Chemistry of Materials 22, 3441-3450, doi:10.1021/cm100477v (2010). Kim, H. & Macosko, C. W. Processing-property relationships of polycarbonate/graphene composites. Polymer 50, 3797-3809, doi:10.1016/j.polymer.2009.05.038 (2009). Liu, H., Kuila, T., Kim, N. H., Ku, B.-C. & Lee, J. H. In situ synthesis of the reduced graphene oxide–polyethyleneimine composite and its gas barrier properties. J. Mater. Chem. A 1, 3739-3746, doi:10.1039/c3ta01228j (2013). Park, O.-K. et al. Synthesis and properties of iodo functionalized graphene oxide/polyimide nanocomposites. Composites Part B 56, 365-371, doi:10.1016/j.compositesb.2013.08.065 (2014). Stankovich, S. et al. Graphene-based composite materials. Nature 442, 282-286, doi:10.1038/nature04969 (2006). Zheng, D. et al. In situ thermal reduction of graphene oxide for high electrical conductivity and low percolation threshold in polyamide 6 nanocomposites. Composites Science and Technology 72, 284-289, doi:10.1016/j.compscitech.2011.11.014 (2012). Zhu, J. et al. Multifunctional polyimide/graphene oxide composites via in situ polymerization. Journal of Applied Polymer Science 131, 40177-40184, doi:10.1002/app.40177 (2014). Ding, P. et al. Highly thermal conductive composites with polyamide-6 covalentlygrafted graphene by an in situ polymerization and thermal reduction process. Carbon 66, 576-584, doi:10.1016/j.carbon.2013.09.041 (2014). Ji, W.-F. et al. Preparation and comparison of the physical properties of PMMA/thermally reduced graphene oxides composites with different carboxylic group
EP
17
AC C
520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562
29
ACCEPTED MANUSCRIPT
37
38
39
40 41 42
43
44
45
46
RI PT
SC
36
M AN U
35
TE D
34
EP
33
content of thermally reduced graphene oxides. Composites Part A 65, 108-114, doi:10.1016/j.compositesa.2014.05.017 (2014). Chen, W.-Q. et al. In situ random co-polycondensation for preparation of reduced graphene oxide/polyimide nanocomposites with amino-modified and chemically reduced graphene oxide. Journal of Materials Science 50, 3860-3874, doi:10.1007/s10853-015-8890-7 (2015). Huang, H.-D. et al. Improved barrier properties of poly(lactic acid) with randomly dispersed graphene oxide nanosheets. Journal of Membrane Science 464, 110-118, doi:10.1016/j.memsci.2014.04.009 (2014). Mittal, V., Chaudhry, A. U. & Luckachan, G. E. Biopolymer - Thermally reduced graphene nanocomposites: Structural characterization and properties. Materials Chemistry and Physics 147, 319-332, doi:10.1016/j.matchemphys.2014.05.007 (2014). Ramanathan, T. et al. Functionalized graphene sheets for polymer nanocomposites. Nature Nanotech 3, 327-331, doi:10.1038/nnano.2008.96 (2008). Compton, O. C., Kim, S., Pierre, C., Torkelson, J. M. & Nguyen, S. T. Crumpled Graphene Nanosheets as Highly Effective Barrier Property Enhancers. Advanced Materials 22, 4759-4763, doi:10.1002/adma.201000960 (2010). Etmimi, H. M., Mallon, P. E. & Sanderson, R. D. Polymer/graphite nanocomposites: Effect of reducing the functional groups of graphite oxide on water barrier properties. European Polymer Journal 49, 3460-3470, doi:10.1016/j.eurpolymj.2013.08.004 (2013). Jin, J., Rafiq, R., Gill, Y. Q. & Song, M. Preparation and characterization of high performance of graphene/nylon nanocomposites. European Polymer Journal 49, 26172626, doi:10.1016/j.eurpolymj.2013.06.004 (2013). Kim, H. & Macosko, C. W. Morphology and Properties of Polyester/Exfoliated Graphite Nanocomposites. Macromolecules 41, 3317-3327, doi:10.1021/ma702385h (2008). Chen, J.-t. et al. Enhancing polymer/graphene oxide gas barrier film properties by introducing new crystals. Carbon 75, 443-451, doi:10.1016/j.carbon.2014.04.024 (2014). Yousefi, N. et al. Highly aligned, ultralarge-size reduced graphene oxide/polyurethane nanocomposites: Mechanical properties and moisture permeability. Composites Part A 49, 42-50, doi:10.1016/j.compositesa.2013.02.005 (2013). Li, Y., Umer, R., Samad, Y. A., Zheng, L. & Liao, K. The effect of the ultrasonication pretreatment of graphene oxide (GO) on the mechanical properties of GO/polyvinyl alcohol composites. Carbon 55, 321-327, doi:10.1016/j.carbon.2012.12.071 (2013). Rafiq, R., Cai, D., Jin, J. & Song, M. Increasing the toughness of nylon 12 by the incorporation of functionalized graphene. Carbon 48, 4309-4314, doi:10.1016/j.carbon.2010.07.043 (2010). Yuan, D., Wang, B., Wang, L., Wang, Y. & Zhou, Z. Unusual toughening effect of graphene oxide on the graphene oxide/nylon 11 composites prepared by in situ melt polycondensation. Composites Part B 55, 215-220, doi:10.1016/j.compositesb.2013.05.055 (2013). Glover, A. J., Schniepp, H. C., Kranbuehl, D. E., Cai, M. & Overdeep, K. R. In SituReduction of Graphene Oxide in Polymers. Macromolecules 44, 9821-9829, doi:10.1021/ma2008783 (2011).
AC C
563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605
30
ACCEPTED MANUSCRIPT
51 52 53
54
55
56
57
58 59
60
61
RI PT
SC
50
M AN U
49
TE D
48
Chen, J., Radke, W. & Pasch, H. Analysis of Polyamides by Size Exclusion Chromatography and Laser Light Scattering. Macromolecular Symposia 193, 107-118 (2003). Skrovanek, D. J., Painter, P. C. & Coleman, M. M. Hydrogen-Bonding in Polymers. 2. Infrared Temperature Studies of Nylon-11. Macromolecules 19, 699-705, doi:10.1021/ma00157a037 (1986). Hocker, S., Rhudy, A. K., Ginsburg, G. & Kranbuehl, D. E. Polyamide hydrolysis accelerated by small weak organic acids. Polymer 55, 5057-5064, doi:10.1016/j.polymer.2014.08.010 (2014). Jacques, B., Werth, M., Merdas, I., Thominette, F. & Verdu, J. Hydrolytic ageing of polyamide 11. 1. Hydrolysis kinetics in water. Polymer 43, 6439-6447, doi:10.1016/S0032-3861(02)00583-9 (2002). Laidler, K. J. Chemical Kinetics. 3 edn, (Harper & Row, 1987). Jarrin, J., Driancourt, A., Brunet, R. & Pierre, B. in MERL Oilfield engineering with polymers (London, UK, 1998). De Groote, P. H., Rouxhet, P. G., Devaux, J. & Godard, P. Infrared study of the hydrogen bonding association in polyamides plasticized by benzenesulfonamides. Part I: Selfassociation in amide and sulfonamide systems; Part II: Amide-sulfonamide interaction. Applied Spectroscopy 55, 877-887, doi:10.1366/0003702011952677 (2001). Zahn, D. Car-Parrinello molecular dynamics simulation of base-catalyzed amide hydrolysis in aqueous solution. Chemical Physics Letters 383, 134-137, doi:10.1016/j.cplett.2003.11.015 (2004). Xiong, Y. & Zhan, C. G. Theoretical studies of the transition-state structures and free energy barriers for base-catalyzed hydrolysis of amides. Journal of Physical Chemistry A 110, 12644-12652, doi:10.1021/jp063140p (2006). Bender, M. L. & Thomas, R. J. Concurrent Alkaline Hydrolysis and Isotopic Oxygen Exchange of a Series of p-Substituted Acetanilides. Journal of the American Chemical Society 83, 4183-4189, doi:10.1021/ja01481a021 (1961). Brown, R. S., Bennet, A. J. & Slebockatilk, H. Recent Perspectives Concerning the Mechanism of H3O+ Promoted and OH- Promoted Amide Hydrolysis. Acc. Chem. Res. 25, 481-488, doi:10.1021/ar00023a001 (1992). Pollack, R. M. & Bender, M. L. Alkaline Hydrolysis of Para Nitroacetanilide and Para Fromylacetanilide. Journal of the American Chemical Society 92, 7190-7194 (1970). Bagno, A., Lovato, G. & Scorrano, G. Thermodynamics of Protonation and Hydration of Aliphatic Amides. J. Chem. Soc.-Perkin Trans. 2, 1091-1098, doi:10.1039/p29930001091 (1993). Fersht, A. R. Acyl-Transfer Reactions of Amides and Esters with Alcohols and Thiols Reference System for Serine and Cysteine Proteinases - Concerning N Protonation of Amides and Amide-Imidate Equilibria. Journal of the American Chemical Society 93, 3504-3515, doi:10.1021/ja00743a035 (1971). Martin, R. B. O-Protonation of Amides in Dilute Acids. J. Chem. Soc.-Chem. Commun., 793-794, doi:10.1039/c39720000793 (1972).
EP
47
AC C
606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647
31
ACCEPTED MANUSCRIPT
67 68
69
70
71 72
73
74
75
76
RI PT
66
SC
65
M AN U
64
TE D
63
Kresge, A. J., Fitzgera.Ph & Chiang, Y. Position of Protonation and Mechanism of Hydrolysis of Simple Amides. Journal of the American Chemical Society 96, 4698-4699, doi:10.1021/ja00821a069 (1974). Zahn, D. On the Role of Water in Amide Hydrolysis. Eur. J. Org. Chem. 2004, 4020-4023, doi:10.1002/ejoc.200400316 (2004). Pan, B., Ricci, M. S. & Troutt, B. L. A Molecular Mechanism of Hydrolysis of Peptide Bonds at Neutral pH Using a Model Compound. Journal of Physical Chemistry B 115, 5958-5970, doi:10.1021/jp1076802 (2011). Duan, P. G., Dai, L. Y. & Savage, P. E. Kinetics and mechanism of N-substituted amide hydrolysis in high-temperature water. J. Supercrit. Fluids 51, 362-368, doi:10.1016/j.supflu.2009.09.012 (2010). Schowen, R. L., Jayarama.H, Kershner, L. & Zuorick, G. W. Solvent Isotope Effects in Amide Hydrolysis. Journal of the American Chemical Society 88, 4008-4012, doi:10.1021/ja00969a020 (1966). Schowen, R. L., Jayaraman, H. & Kershner, L. Catalytic efficiencies in amide hydrolysis. The two-step mechanism. 88, 3373-3375 (1966). Zahn, D. Theoretical study of the mechanisms of acid-catalyzed amide hydrolysis in aqueous solution. Journal of Physical Chemistry B 107, 12303-12306, doi:10.1021/jp034175h (2003). Oh, Y. J. et al. Oxygen functional groups and electrochemical capacitive behavior of incompletely reduced graphene oxides as a thin-film electrode of supercapacitor. Electrochimica Acta 116, 118-128, doi:10.1016/j.electacta.2013.11.040 (2014). Panaitescu, D. M., Frone, A. N. & Nicolae, C. Micro- and nano-mechanical characterization of polyamide 11 and its composites containing cellulose nanofibers. European Polymer Journal 49, 3857-3866, doi:10.1016/j.eurpolymj.2013.09.031 (2013). Zhang, Q. Z., Mo, Z. S., Liu, S. Y. & Zhang, H. F. Influence of annealing on structure of Nylon 11. Macromolecules 33, 5999-6005, doi:10.1021/ma000298d (2000). Zhang, Q., Yu, M. & Fu, Q. Crystal morphology and crystallization kinetics of polyamide11/clay nanocomposites. Polymer International 53, 1941-1949, doi:10.1002/pi.1581 (2004). Yang, Z., Huang, S. & Liu, T. X. Crystallization Behavior of Polyamide 11/Multiwalled Carbon Nanotube Composites. Journal of Applied Polymer Science 122, 551-560, doi:10.1002/app.34118 (2011). Kolesov, I. et al. Crystallization of a polyamide 11/organo-modified montmorillonite nanocomposite at rapid cooling. Colloid Polym Sci 291, 2541-2549, doi:10.1007/s00396013-2977-y (2013). Capsal, J.-F., Pousserot, C., Dantras, E., Dandurand, J. & Lacabanne, C. Dynamic mechanical behaviour of polyamide 11/Barium titanate ferroelectric composites. Polymer 51, 5207-5211, doi:10.1016/j.polymer.2010.09.011 (2010). Lacrampe, M.-F., Prashantha, K. & Krawczak, P. Highly dispersed polyamide11/halloysite nanocomposites: Thermal, rheological, optical, dielectric, and mechanical properties. Journal of Applied Polymer Science 130, 313-321, doi:10.1002/app.39160 (2013).
EP
62
AC C
648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690
32
ACCEPTED MANUSCRIPT
81
82
83
84 85
RI PT
SC
80
M AN U
79
TE D
78
Mago, G., Kalyon, D. M. & Fisher, F. T. Nanocomposites of Polyamide-11 and Carbon Nanostructures: Development of Microstructure and Ultimate Properties Following Solution Processing. Journal of Polymer Science Part B-Polymer Physics 49, 1311-1321, doi:10.1002/polb.22311 (2011). Ma, Y. L., Hu, G. S., Ren, X. L. & Wang, B. B. Non-isothermal crystallization kinetics and melting behaviors of nylon 11/tetrapod-shaped ZnO whisker (T-ZnOw) composites. Materials Science and Engineering a-Structural Materials Properties Microstructure and Processing 460, 611-618, doi:10.1016/j.msea.2007.01.133 (2007). Wu, M. et al. Nonisothermal crystallization kinetics of ZnO nanorod filled polyamide 11 composites. Materials Chemistry and Physics 109, 547-555, doi:10.1016/j.matchemphys.2008.01.002 (2008). Zhang, G. S., Li, Y. J. & Yan, D. Y. Polymorphism in nylon-11/montmorillonite nanocomposite. Journal of Polymer Science Part B-Polymer Physics 42, 253-259, doi:10.1002/polb.10680 (2004). Liu, T.-x., Chen, D., Phang, I. Y. & Wei, C. Studies on crystal transition of polyamide 11 nanocomposites by variable-temperature X-ray diffraction. Chin J Polym Sci 32, 115-122, doi:10.1007/s10118-014-1383-3 (2013). Latko, P., Kolbuk, D., Kozera, R. & Boczkowska, A. Microstructural Characterization and Mechanical Properties of PA11 Nanocomposite Fibers. Journal of Materials Engineering and Performance 25, 68-75, doi:10.1007/s11665-015-1817-2 (2016). Kawaguchi, A., Ikawa, T., Fujiwara, Y., Tabuchi, M. & Monobe, K. Polymorphism in Lamellar Single-Crystals of Nylon-11. J. Macromol. Sci.-Phys. B20, 1-20, doi:10.1080/00222348108219425 (1981). Fornes, T. D. & Paul, D. R. Crystallization behavior of nylon 6 nanocomposites. Polymer 44, 3945-3961, doi:10.1016/s0032-3861(03)00344-6 (2003). Hocker, S., Hudson-Smith, N., Schniepp, H. C. & Kranbuehl, D. E. Enhancing polyimide's water barrier properties through addition of functionalized graphene oxide. Polymer 93, 23-29, doi:http://dx.doi.org/10.1016/j.polymer.2016.04.008 (2016).
EP
77
AC C
691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719
33
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
1. Graphene oxide reduces extent of hydrolytic degradation in polyamide-11. 2. Graphene oxide reduces molecular mobility in polyamide-11. 3. Polyamide's amide hydrogen bonds with the graphene oxide surface