Sie sind auf Seite 1von 9

Polymer 54 (2013) 1141e1149

Contents lists available at SciVerse ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Polyamides based on the renewable monomer, 1,13-tridecane diamine


II: Synthesis and characterization of nylon 13,6
Satyabrata Samanta a, Jie He a, Sermadurai Selvakumar b, Jessica Lattimer c, Chad Ulven c, Mukund Sibi b,
James Bahr a, Bret J. Chisholm a, d, *
a

Center for Nanoscale Science and Engineering, North Dakota State University, Fargo, ND 58102, USA
Department of Chemistry and Biochemistry, North Dakota State University, Fargo, ND 58102, USA
Department of Mechanical Engineering, North Dakota State University, Fargo, ND 58102, USA
d
Department of Coatings and Polymeric Materials, North Dakota State University, Fargo, ND 58102, USA
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 30 October 2012
Received in revised form
28 November 2012
Accepted 6 December 2012
Available online 19 December 2012

Nylon 13,6 was successfully synthesized and its chemical composition, thermal properties, crystal structure, and moisture absorption characterized. Melting temperature and glass transition temperature were
determined to be 206  C and 60  C, respectively, while the equilibrium melting temperature was determined to be 248  C. Characterization of the crystallization kinetics showed that nylon 13,6 exhibits very fast
crystallization compared to the industrially important nylons, nylon 6 and nylon 6,6. In addition, the
moisture absorption of nylon 13,6 was dramatically lower than nylon 6 and nylon 6,6 which is consistent
with the much lower amide content of nylon 13,6. The crystal structure was determined to be different from
the expected g-form, which possess a pseudohexagonal unit cell. Instead, the x-ray scattering pattern was
more similar to the a-form. A similar result has been obtained for other oddeeven nylons including nylon
5,6, nylon 5,10, nylon 9,2, nylon 11,10, and nylon 11,12.
2012 Elsevier Ltd. All rights reserved.

Keywords:
Nylon
Polyamide
Semicrystalline

1. Introduction
Since the pioneering work by Carothers [1,2], polyamides,
commonly referred to as nylons, have become a mainstay in the
plastic, lm, and ber industries [3]. While many different nylon
homopolymers and copolymers have been reported and commercialized, nylon 6,6 and nylon 6 have been by far the most commercially successful. Compared to other semicrystalline, aliphatic
polymers such as polyethylene, polypropylene, and polycaprolactone, aliphatic nylons exhibit relatively high melting and
glass transition temperatures. The higher melting and glass transition temperatures can be attributed to hydrogen bonding between
amide groups that effectively restrict molecular motions in the
solid-state. The extensive hydrogen bonding associated with polyamides also provides strength, toughness, ductility, wear and
abrasion resistance, and resistance to oils, greases, and other
hydrocarbons. The primary drawback associated with nylon 6 and
nylon 6,6 is extensive moisture absorption. At 23  C and 100 percent

* Corresponding author. Department of Coatings and Polymeric Materials, North


Dakota State University, Fargo, ND 58102, USA.
E-mail address: Bret.Chisholm@ndsu.edu (B.J. Chisholm).
0032-3861/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.polymer.2012.12.034

relative humidity, the equilibrium moisture absorption of nylon 6,6


and nylon 6 has been determined to be 8.5 and 9.5 weight percent,
respectively [4]. The high level of moisture absorption of nylon 6 and
nylon 6,6 can result in poor dimensional stability of molded objects
and signicant changes in mechanical properties resulting from
plasticization by absorbed water. Due to their lower amide content,
aliphatic nylon polymers derived from longer chain diacids and
diamines exhibit much lower water absorption than nylon 6,6 and
nylon 6. For example, the equilibrium moisture absorption of nylon
6,12, nylon 11, and nylon 12 at 23  C and 100 percent relative
humidity has been reported to be 3.0, 1.9, and 1.6 weight percent,
respectively [4].
Due to concerns over the long-term availability of chemicals
derived from fossil resources, a renewed interested in the development of materials from renewable resources has occurred [5e9].
With regard to polyamides, plant oils represent an excellent source
for obtaining aliphatic diacids and diamines. As illustrated in Fig. 1a,
unsaturated fatty acids obtained from the hydrolysis of plant oil
triglycerides can be readily utilized to produce long-chain aliphatic
dicarboxylic acids through an oxidation process. Oxidation of the
most common unsaturated fatty acid, i.e. oleic acid, produces the
nine carbon dicarboxylic acid, azelaic acid. The ten carbon dicarboxylic acid, sebacic acid, can be obtained by alkali pyrolysis of

1142

S. Samanta et al. / Polymer 54 (2013) 1141e1149

Fig. 1. Illustration of the synthesis of dicarboxylic acids and diamines from a fatty acid (oleic acid): (a) dicarboxylic acid synthesis; (b) diamine synthesis.

ricinoleic acid, which is the major fatty acid component of castor


oil [10,11].
Dicarboxylic acids can be readily used as a substrate to produce
diamines. Fig. 1b shows the synthetic scheme commonly used to
produce diamines from dicarboxylic acids on an industrial scale. The
nine carbon diamine, 1,9-nonane diamine, has recently been used
for the production of a semi-aromatic nylon commercialized by
Kuraray under the tradename, Genestar.
The authors have been interested in nylons derived, at least in
part, from erucic acid, which is a major component of crambe and
rapeseed oil [12]. Crambe oil and industrial rapeseed oil contain 55e
60 weight percent and 40e55 weight percent erucic acid, respectively [13]. Oxidation of erucic acid yields the thirteen carbon
dicarboxylic acid, brassylic acid, which can be subsequently used to
produce the thirteen carbon diamine, 1,13-tridecane diamine (TDA).
Most of the published research involving euric acid as a renewable
source for the production of nylons appears to have been done
within laboratories at the United States Department of Agriculture
(USDA) and Southern Research Institute (SRI) [12e18]. These
researchers focused their efforts on the study of nylon 13 [17], nylon
13,13 [16,18], and nylon 6,13 [14]. The physical properties of these
three nylons were fully evaluated and compared to nylon 11 and
nylon 6,10. The most signicant difference between the three nylons
derived from brassylic acid and nylon 11 or nylon 6,10 was the lower
water absorption for the brassylic acid-based nylons. Of these three
brassylic acid-based nylons, nylon 13,13 appears to have been
investigated to the greatest extent as suggested by scale-up efforts
directed toward this polymer [16]. In addition to scale-up, a detailed
cost analyses for the production of nylon 13,13 was published [15].
The physical properties of the nylon 13,13 produced at the multipound scale were thoroughly evaluated and compared to two
commercially available long-chain nylons, nylon 11 and nylon 12. It
was found that the physical and mechanical properties of nylon
13,13 were similar to nylon 11 and nylon 12. The primary differences
were lower melting point, slightly lower density, and lower moisture absorption for nylon 13,13.
In addition to researchers at the USDA and SRI laboratories,
a few other groups have investigated nylon 13,13 or other nylons
based brassylic acid. Wang and coworkers [19] characterized the

thermal properties, crystal structure, and mechanical properties of


nylon 13,13 samples that were uniaxially drawn by solid-state
extrusion. The glass transition temperature (Tg) and melting
temperature (Tm) were determined to be 56  C and 174  C,
respectively. The crystal structure for nylons have been classied,
according to Kinoshita [20], into two basic forms, namely, the
a and g forms. It was determined that the crystal structure for
nylon 13,13 was of the g form with a monoclinic unit cell. The
crystal structure and Tm observed by Wang et al. [19] was also
observed by Prieto and coworkers [21]. Gidanian and Howard
synthesized the brassylic acid-based polyamides, nylon 6,13 and
nylon 10,13, and compared the physical properties to the
commercial polyamide, nylon 6,12. The Tm and Tg of nylon 6,13
was determined to be 205  C and 47  C, respectively, while the Tm
and Tg of nylon 10,13 was determined to be 187  C and 43  C,
respectively. Kim and Yu [22] prepared a series of homopolyamides
and copolyamides based on brassylic acid and characterized
various properties. The hompolyamides prepared were nylon
12,13, nylon 10,13, nylon 8,13, and nylon 6,13. The Tms of these
polymers ranged from 175  C to 208  C with Tm increasing with
a decrease in the number of carbon atoms of the diamine
component, while the Tgs increased from 42  C to 51  C with
decreasing diamine chain length. All of these long-chain, brassylic
acid-based polyamides exhibited moisture absorption in the range
of 1.0e1.6 weight percent. Copolyamides were produced from
hexamethylene diamine and mixtures of brassylic acid with
adipic, terephthalic, or isophthalic acid. All copolymer systems
exhibited eutectic behavior with respect to the inuence of
copolymer composition on Tm. Cui and coworkers [23] synthesized seven different polyamides based on brassylic acid with the
length of the diamine component ranging from 3 carbons to 11
carbons. For this series of brassylic acid-based nylons, Tg ranged
between 29  C and 36  C while Tm ranged between 176  C and
191  C. Analysis of the crystal structure of these polymers revealed
that all of the nylons except nylon 6,13 exhibited the g form with
a monoclinic unit cell. The wide-angle X-ray diffraction pattern for
nylon 6,13 exhibited two distinct diffractions at 0.44 nm and
0.38 nm. This diffraction pattern is similar to that expected for the
a-structure; however, as pointed out by Puiggali and coworkers

S. Samanta et al. / Polymer 54 (2013) 1141e1149

[24], polyamides based on odd diamines and/or odd dicarboxylic


acids cannot form the fully extended planar zigzag conformation of
molecular chains that produce the parallel planar sheets of
hydrogen bonded molecules that dene the a-form. Based on their
model, this a-like structure can result from a molecular conrmation in which consecutive amide groups (those separated by the
odd methylene segment) rotate slightly to produce structures that
possess two hydrogen bond directions. Besides nylon 6,13,
several other polyamides based on an odd diamine or odd dicarboxylic have been shown to exhibit this a-like diffraction pattern
[25e29].
In spite of the tremendous importance of polyamides in society
and the enormous efforts made regarding the research and development of nylons, the authors were unable to nd any journal
reports describing the synthesis and characterization of nylon 13,T
or nylon 13,6. In addition, none of the previous reports based on
polyamides produced from either brassylic acid or TDA provided
information about crystallization.
2. Experimental
2.1. Materials
Adipic acid (AA) was purchase from Fluka (99.5% purity), while
isopropanol (IPA) (ACS grade) and hexane (ACS grade) were
purchase from BDH. Deuterated triuoroacetic acid was obtained
from SigmaeAldrich (99.5% purity).
2.2. Synthesis of the monomer salt of AA and TDA
TDA was synthesized using the procedure described by He
et al. [37]. The monomer salt of AA and TDA was produced as
follows: A TDA solution was prepared by dissolving 5.35 g of TDA
in 50 mL of hot IPA. An AA solution was prepared by dissolving
3.65 g of AA in 75 mL of hot IPA. The two solutions were combined
in a round-bottom ask and the mixture reuxed for 2 h. The
solution was then allowed to cool to room temperature before the
volume was reduced to 50 mL using a rotary evaporator. To this
concentrated solution, 50 mL of hexane was added and then the
solution stored in a refrigerator for overnight. The precipitate
produced was isolated and washed with IPA (20 mL) using
vacuum ltration. The salt was then dried under vacuum and
characterized using 1H NMR. Yield: 7.8 g (87%). 1H NMR (400 MHz,
CD3OD) d 1.30e1.38 (m, 18 H); 1.59e1.64 (m, 8H); 2.17 (t, 6.5 Hz,
4H); 2.87 (t, 7.4 Hz, 4H).
2.3. Synthesis of nylon 13,6
Nylon 13,6 was synthesized using a two-step process. First, an
oligomer/prepolymer was produced by dissolving 5e10 g of the
AA/TDA salt in a minimum amount of hot, nitrogen-purged
water (3e4 weight excess); transferring the solution to
a stainless steel pressure vessel equipped with a magnetic stir
bar; degassing the solution ve times; heating the vessel to
220  C; holding the temperature at 220  C for 2 h; releasing the
pressure and allowing the water to removed; and further
heating the material at 220  C under high vacuum for 1.5 h.
After allowing the vessel to cool to room temperature under
vacuum, the brittle oligomer/prepolymer was removed as small
chips. Next, 3e10 g of prepolymer was added to a glass tube
equipped with an overhead stirrer, vacuum line, nitrogen inlet
and outlet, and an oil bath. High vacuum was applied to the
tube and before heating with the oil bath preheated to 250  C.
Once the prepolymer was molten, the overhead stirring was
started and the material held under vacuum at 250  C for 1.5 h.

1143

After this period, the heat to the oil bath was turned off and the
material allowed to cool overnight while maintaining the
vacuum. The polymer was recovered the next day by sacricing
the glass tube.

2.4. Characterization
A JEOL-ECA 400 (400 MHz) NMR spectrometer was used to
generate 1H NMR spectra. Data acquisition was completed using 16
scans, and the polymer sample was dissolved in deuterated triuoroacetic acid. Infrared spectra were obtained on at, bulk
specimens with a Bruker Vertex 70 spectrophotometer equipped
with a Pike Technologies MIRacle ATR attachment that utilized
a Ge crystal. Before characterization using thermal techniques, all
polymer samples were dried in a vacuum oven at 100  C overnight.
Differential scanning calorimetry (DSC) experiments were performed using a TA Q2000 instrument. The experiments were conducted under a N2 atmosphere, and the instrument was calibrated
with an indium standard. Sample sizes ranged from 4.5 to 7.5 mg.
For characterization of the melting transition, specimens were
cooled from 260  C to 0  C at a cooling rate of 20  C/min and then
reheated to 260  C at different heating rates (i.e. 5, 20, and 40  C/
min.). Isothermal crystallization kinetics were determined by
cooling specimens from 260  C to the predetermined crystallization
temperature at a rate of 100  C/minute and observing the heat ow
as a function of time.
The thermal stability of nylon 13,6 was characterized using
a Q500 Thermogravimetric Analyzer from TA Instruments (New
Castle, DE). Samples (8.5e10 mg) were heated from 25  C to
1000  C at a heating rate of 10  C/min under a nitrogen or air
atmosphere.
X-Ray Diffraction (XRD) data was collected in BraggBrentano
geometry using a Rigaku Ultima IV instrument (Cu Ka radiation,
voltage 40 kV, current 44 mA). The samples were scanned from 2 to
60 2-Theta with a step size of 0.025 and a scan time of 10 s/step.
The samples were rotated at 25 rpm. Slits used were a 2/3 mm
divergence slit, 10 mm height limiting slit, 5 soler slits, 2/3 mm
scattering slit, and 0.3 mm receiving slit.
The solution viscosity of the nylon 13,6 sample produced for the
study was determined using an Ubbelohde viscometer (model 1B)
from Technical Glass Products Inc. (Dover, NJ). A 0.10 wt.% polymer
solution in dichloracetic acid was prepared and passed through
a 0.2 um Millipore lter. The intrinsic viscosity [h] of the polymer
was measured at 25  0.1  C using the single-point viscosity
equation described by Soloman and Cuita [30]:

i.
h 
c
h Sqrt 2 hsp  ln hrel

(1)

where hsp and hrel are the specic viscosity and relative viscosity,
respectively, and c is the concentration of the polymer solution in
units of g/ml.
The moisture absorption of the nylon 13,6 sample was characterized using the method previously described by Kugel et al. [31].
This method involved melt extrusion and compression of 3 g
samples of nylon 13,6 into small plaques using a Tinius Olsen
MP600 Extrusion Plastometer. The analysis of moisture absorption
was done using an Arizona Instruments Computrac 4000XL Moisture Analyzer. Sample plaques were soaked in distilled water for
one week and adsorbed moisture was removed by towel drying the
specimens prior to analysis. Samples were heated to 210  C, while
maintaining this temperature, mass loss was recorded. Once the
mass loss slowed to 0.015% moisture/minute the analysis was
complete and the total mass loss measured was recorded as the
total moisture absorbed.

1144

S. Samanta et al. / Polymer 54 (2013) 1141e1149

3. Results and discussion

1
0.95

Nylon 13,6 was successfully produced by rst preparing the 1/1


mole/mole salt of AA and TDA, producing oligomeric species from
AA and TDA by heating an aqueous solution of the AA/TDA salt in
a closed vessel, and, nally, producing high molecule weight
polymer from the oligomer via melt condensation polymerization.
Production of the oligomeric species was needed to inhibit monomer sublimation during the melt condensation polymerization. The
intrinsic viscosity of the polymer produced for the investigation
was 0.65 dl/g.
1
H NMR and FT-IR were used to conrm the successful synthesis
of nylon 13,6. Fig. 2 displays the 1H NMR of the sample along with
the peak assignments [32]. The peaks at approximately 2.6 ppm are
consistent with the methylene protons next to the nitrogen atom of
amine end-groups. Fig. 3 displays the FT-IR spectrum which
exhibits the characteristic bands associated with an aliphatic
polyamide. Bands with peak maximums at 3300 cm1 (NeH
stretching), 1634 cm1 (C]O stretching), and 1537 cm1 (CeN
stretching and COeNeH bending) all correspond to motions associated with the amide group, while bands at 2918 cm1 (CH2
asymmetric stretching) and 2848 cm1 (CH2 symmetric stretching)
correspond to motions associated with methylene groups.
3.2. Thermal properties
Fig. 4 displays a DSC thermogram produced by heating an as
polymerized sample from 25  C to 260  C at a heating rate of 40  C/
minute. According to the thermogram, nylon 13,6 is a semicrystalline polymer with a Tg of approximately 60  C and Tm of
approximately 206  C. In addition to a Tg and melting endotherm,
a sharp crystallization exotherm was observed with a peak
temperature of 191  C. Since this exotherm occurs at a temperature
higher than the onset of the melting endotherm (approximately
150  C), it was attributed to a crystallite reorganization process. To
conrm this conclusion, DSC thermograms were obtained by
cooling specimens from the melt at 20  C/minute and then

O
C

Transmission

3.1. Chemical characterization

0.9

3300

0.85

2848

0.8

2918

0.75

1537

0.7
1634

0.65
0.6
500

1000

1500

2000

2500

3000
-1

3500

4000

Wavenumber (cm )

Fig. 3. FT-IR spectrum of nylon 13,6.

reheating at different heating rates. As shown in Fig. 5, the


magnitude of the crystallization exotherm decreased with
decreasing heating rate, which is the expected outcome if the
crystallization exotherm is the result of crystallite reorganization
process. By heating at a slower rate, more time is available for
polymer chains present in the lowest melting, least thermally
stable crystallites to reorganize into higher melting, more thermally stable crystallites. Thus, the heat of crystallization is less
apparent in the thermogram. The observation of a meltingrecrystallization process has been observed for several other
polyamides including nylon 12,12 [33], nylon 12,14 [34], nylon 6,6
[35], and nylon 12,10 [36]. The authors have conducted the identical DSC experiment on nylon 13,T [37]. A comparison of the
thermograms obtained for nylon 13,T to those obtained for nylon
13,6 showed that the melting-recrystallization phenomenon was
much more prevalent with nylon 13,6. This result can be attributed
to the higher thermal stability of crystallites based on terephthalamide repeat units as opposed to adipamide repeat units. The
stronger intermolecular interactions associated with aromatic rings
reduces that ability of the polymer chains to reorganize into thicker,
more stable crystallites.

O
a

CH2

CH2

CH2

CH2

N
H

CH2

CH2

CH2

CH2

CH2

Tm = 206 C

N
H

Heat Flow
endo

Tg = 60 C
c

10

PPM
1

Fig. 2. H NMR spectrum of nylon 13,6.

50

100

150

200

250

Temperature ( C)
Fig. 4. DSC thermogram of nylon 13,6 generated by heating at 40  C/min.

S. Samanta et al. / Polymer 54 (2013) 1141e1149

1145

Table 1
Temperatures corresponding to 5 and 50 weight loss for nylon 6,6, nylon 13,T, and
nylon 13,6 in a nitrogen atmosphere.

5 C/min
o

20 C/min

Polymer

Temp. at 5%
wt. loss ( C)

Temp. at 50%
wt. loss ( C)

Nylon 6,6
Nylon 13,T
Nylon 13,6

411
396
359

471
467
443

endo

Heat Flow

40 C/min

3.3. Crystallization kinetics

100

120

140

160

180

200

220

Temperature ( C)
Fig. 5. DSC thermograms of nylon 13,6 produced using different heating rates.

The thermal stability of nylon 13,6 was characterized using TGA


and both air and nitrogen as an atmosphere. In general, the thermal
stability appears to be high enough to enable melt processing of the
polymer without major issues with thermal degradation. Since the
Tm of the polymer is 206  C, it should be easily processed at
temperatures between 230 and 250  C. According to the TGA
thermograms shown in Fig. 6, the weight loss in both air and
nitrogen are negligible over this temperature range. The onset of
thermal degradation for nylon 13,6 occurs above 300  C with 50
percent weight loss occurring at approximately 437  C. For
comparison purposes, the authors previous thermal stability data
obtained for nylon 6,6 [31] and nylon 13,T [37] can be considered.
Table 1 lists the temperatures corresponding to 5 and 50 percent
weight loss for nylon 6,6, nylon 13,T, and nylon 13,6. The higher
thermal stability obtained for nylon 13,T as compared to nylon 13,6
can be attributed to the higher thermal stability of the aromatic
terephthalamide units as compared to the aliphatic adipamide
units.

Xt 1  expKt n

100

Nitrogen

60

Air
40

20

Tm T m1  1=g Tc=g

0
0

100

200

(2)

where X(t) is the relative extent of crystallization at time t, K is the


crystallization rate parameter, and n is the Avrami exponent
[39,40]. The Avrami exponent provides information about the
crystal growth geometry and the nucleation mechanism.
Prior to conducting the isothermal crystallization experiments,
it was necessary to determine the equilibrium melting temperature
(T m) for nylon 13,6. The T m for a polymer, which is dened as the
melting temperature of innitely thick crystals, can be obtained by
measuring the Tm of a polymer as function of isothermal crystallization temperature in accordance with the HoffmaneWeeks [41]
equation:

80

Weight %

The crystallization kinetics for nylon 13,6 was characterized


using both isothermal and nonisothermal methods. A convenient
method for comparing overall crystallization rates of polymers with
different melting temperatures involves simple nonisothermal DSC
experiments in which Tm is compared to the maximum temperature of the crystallization exotherm (Tc) obtained upon controlled
cooling from the melt. For nylon 13,6, the difference between the
Tm and Tc (i.e. DT) obtained at a cooling rate of 20  C was 29  C. This
value is signicantly lower than the DT values obtained by the
authors for nylon 6,6 (DT 33) and nylon 6 (DT 46) [37] indicating faster overall crystallization for nylon 13,6, but higher than
that obtained for the semi-aromatic analog, nylon 13,T (DT 22)
[37].
Khanna [38] proposed a parameter referred to as the crystallization rate coefcient (CRC) to directly compare the crystallization
rates of different polymers. The CRC is determined using DSC and
measuring Tc at three different cooling rates (i.e. 1.25, 10 and 40  C/
min). CRC is dened as the slope of the plot of cooling rate as
a function of Tc. The basic premise behind the CRC is that the Tc
measured for relatively fast crystallizing polymers will be less
dependent on cooling rate than relatively slow crystallizing polymers. Thus, faster crystallizing polymers possess higher values for
CRC. The authors determined the CRC for nylon 6 [31], nylon 6,6
[31], nylon 13,T [37], and nylon 13,6 to be 54/h, 82/h, 164/h, and
156/h, respectively. The trend in CRC value with nylon composition
resulted in the same ranking with respect to overall crystallization
rate as that obtained by comparing values of DT.
As a means to obtain more detailed information of the crystallization kinetics as well as information about the crystallization
mechanism, isothermal experiments were conducted. By conducting isothermal experiments, the Avrami parameters, K and n,
can be obtained by plotting the data in accord with the Avrami
equation:

300

400

500

600

(3)

700

Temperature ( C)
Fig. 6. TGA thermograms produced using both an air and nitrogen environment.

where Tm is the experimental melting temperature, Tc is the


isothermal crystallization temperature, and g is a factor that
depends on the nal lamellar thickness of crystallites. Using the

S. Samanta et al. / Polymer 54 (2013) 1141e1149

equation, T m can be determined by extrapolating the plot of Tm as


a function Tc to the Tm Tc line. As shown in Fig. 7, T m was
determined to be 248  C. Based on this result, samples for
measuring isothermal crystallization kinetics were heated above
T m to 260  C to ensure complete melting of even the thickest
crystallites.
For the isothermal crystallization experiments, crystallization
isotherms were readily obtained at crystallization temperatures
between 186  C and 190  C (Fig. 8). With the isothermal data, the
crystallization half-time (t1/2), which is the time required to obtain
50% of the total extent of crystallization, was obtained. Fig. 9
displays t1/2 as a function of supercooling (i.e. DT T m  crystallization temperature). As shown in Fig. 9, t1/2 decreased with
increasing DT, which is consistent with a nucleation controlled
crystallization process.
Using data from the crystallization isotherms corresponding to
relatively low extents of crystallization, the Avrami crystallization
rate parameter, K, and the Avrami exponent, n, were obtained by
plotting ln[ln(1  X(t))] as a function of ln[time (min.)], as shown
in Fig. 10. At each crystallization temperature, the plots yielded
straight lines in which n was obtained from the slope and K obtained from the intercept. Table 2 lists the Avrami parameters obtained along with t1/2 data. The Avrami exponent increased from
2.41 to 3.01 as the crystallization temperature increased from the
lowest temperature (i.e. 186  C) to the highest crystallization
temperature (i.e. 190  C). According to theory, the Avrami exponent
provides information about the nucleation mechanism and the
geometry of crystal growth. The Avrami exponents obtained for
nylon 13,6 suggests spherulitic crystal growth with a change in the
nucleation mechanism from homogenous to heterogeneous with
increasing temperature [42]. Since the Avrami exponent depends
both on crystal growth geometry and the nucleation mechanism
and is based on some assumptions, such as the absence of
secondary crystallization and changes in crystal perfection, other
experimental techniques, such as optical microscopy, are needed to
fully understand the mechanism of crystallization. Nonetheless, it
is interesting to note that the Avrami exponents obtained for nylon
13,6 were signicantly different from those obtained for a similar
polyamide, nylon 13,T [37]. The Avrami exponents obtained by the
authors for nylon 13,T were in the range of 1.5e2.0, which is
signicantly lower than those for nylon 13,6. These lower Avrami
exponents obtained for nylon 13,T indicate one-dimensional
crystal growth. Thus, replacing the more rigid terephthalamide

120

100

X(t) (%)

1146

186 C

60

187 C
o

188 C
40

189 C
o

190 C
20

0
0

Time (min.)
Fig. 8. Crystallization isotherms obtained for nylon 13,6.

units with the more exible adipamide units enabled crystal


growth to occur in three dimensions as opposed to one dimension.
Using the Arrhenius relationship:

K 1=n ko exp  DE=RTc

(4)

where ko is the temperature-independent pre-exponential factor, R


is the gas constant, K is the Avrami kinetic rate parameter, n is the
Avrami exponent, and Tc is the crystallization temperature, the
crystallization activation energy (DE) was determined. From the
slope of the plot shown in Fig. 11, DE was determined to be 855.9 kJ/
mol. This value is considerably higher than that obtained for nylon
13,T [37], which is consistent with the faster rate of crystallization
for nylon 13,T.

120

100

260

80

60

1/2

240

(sec.)

248.1 C

Melting Temperature ( C)

80

220

40

200
20

180

180

200

220

240
o

260

Crystallization Temperature ( C)
Fig. 7. Plot of Tm as a function of isothermal crystallization temperature extrapolated
to the Tm crystallization temperature line. This plot enables the determination of the
equilibrium melting temperature of nylon 13,6.

0
57

58

59

60

61

62

63

( C)
Fig. 9. Plot of crystallization half-time (t1/2) as a function of supercooling (i.e.
DT T m  crystallization temperature) for nylon 13,6.

S. Samanta et al. / Polymer 54 (2013) 1141e1149


0.4

1147

-1
o

186 C

0.2

187 C

-1.5

188 C
o

189 C
o

-2

190 C
-0.2

(lnK)/n

ln[-ln(1-X(t))]

-0.4

-2.5

-0.6

-3
-0.8

-3.5
-1

-1.2
-2

-1.6

-1.2

-0.8

-0.4

0.4

0.8

-4
2.155

1.2

2.16

ln[time (min.)]

2.165

2.17

2.175

2.18

-1

1000/T (K )
c

Fig. 10. Plots based on the Avrami equation that enables n to be obtained from the
slope of the lines and K to be obtained from the intercept.

Fig. 11. Plot based on the Arrhenius relationship that allows the crystallization activation energy to be determined from the slope of the line.

3.4. Crystal structure

Table 2
Kinetic parameters for nylon 13,6 obtained from isothermal crystallization
experiments.
Crystallization temp.
( C)

Supercooling
( C)

t1/2
(s)

K
(s1)

186
187
188
189
190

62
61
60
59
58

25.2
27.6
37.5
66.6
115.8

2.91
3.15
8.65
4.40
4.26

n






104
104
105
106
107

2.41
2.32
2.48
2.85
3.01

investigators have observed crystal structures for oddeeven polyamides that deviate from the anticipated g-form. Two strong
diffraction signals close to the characteristic spacings of the a-form
have been observed for nylon 5,6 [24,25], nylon 5,10 [26], nylon 9,2
[27], nylon 11,10 [29], and nylon 11,12 [29]. As discussed by MoralesGmez et al. [25], the a-like diffraction pattern from these odde
even nylons is the result of a molecular conformation that
possesses close to an all trans conguration with the exception that
there is a slight deviation toward 150 (or 150 ) for the two
torsional angles vicinal to the odd diamide unit. This deviation from
the all trans conformation enables hydrogen-bonding interactions
to be maximized. Thus, the WAXD data obtained for nylon 13,6
appears to be consistent with the results obtained for other odde
even polyamides that do not exhibit the pattern consistent with
the g-form. Therefore, based on model developed by Puiggali et al.
[24] for the a-like structure observed for nylon 5,6, the crystal
structure for nylon 13,6 most likely involves two hydrogen-bonding

1400
1200
1000

Intensity

WAXD was utilized to characterize the crystal structure of nylon


13,6. In general, polyamides exhibit one of two main types of crystal
structures, which are referred to as the a and g structures [43]. The
a structure, which is characterized by a fully extended planar zigzag
conformation of chain segments, is typically observed with evene
even nylons [44]. The crystal symmetry for eveneeven nylons
with this structure is triclinic with one repeat unit per unit cell,
while the crystal symmetry for even nylons, such as nylon 6, is
monoclinic with four repeat units per unit cell. Diffraction patterns
characteristic of the a structure exhibit intense reections with dspacings of 4.4 and 3.7 
A, which correspond to spacings between
crystallographic planes comprised of non-hydrogen bonded and
hydrogen bonded segments, respectively. In contrast, the g structure, which is common for even nylons with more than seven
carbons atoms as well as eveneodd, oddeeven, and oddeodd
nylons, exhibits an intense reection with a d-spacing of 4.15 
A.
The g structure exhibits close to hexagonal packing of polymer
chains.
Fig. 12 displays the one-dimensional WAXD prole obtained for
nylon 13,6. As shown in Fig. 12, the WAXD data shows two strong
reections at 2q values of 20.3 and 22.4 that correspond to dspacings of 4.3 and 4.0 
A, respectively. This result is somewhat
surprising considering the fact that oddeeven nylons typically
exhibit the g-form structure depicted by a single reection with
a d-spacing of 4.15 
A. As discussed by Kinoshita, the g-form maximizes hydrogen bonding for oddeeven polyamides [43]. Other

800
600
400
200
0
0

10

20

30

40

50

2 theta
Fig. 12. The WAXD pattern obtained for nylon 13,6. The sample was obtained from the
synthesis process without any post-polymerization thermal treatment.

1148

S. Samanta et al. / Polymer 54 (2013) 1141e1149

Moisture Uptake After One Week of Immersion (wt. %)

nylon 6

nylon 6,6

nylon 13,6

nylon 13,T

Fig. 13. Moisture absorption after immersion in water for one week at ambient
conditions for nylon 13,6, nylon 6 [31], nylon 6,6 [31], and nylon 13,T [37].

directions and rotation of consecutive amide groups in different


senses from the plane dened by the methylene carbons.
3.5. Moisture absorption
It is well known that the properties of polyamides are highly
dependent on the hydrogen-bonding that occurs between repeat
units. Hydrogen-bonding contributes to many of the desirable
properties of nylons such as high thermal properties, toughness,
and abrasion resistance. Unfortunately, that amide groups that are
source of the hydrogen bonds also lead to signicant moisture
absorption. Absorbed water molecules effectively displace
hydrogen bonds between amide groups and, as a result, can result
in major changes in mechanical properties, thermal properties, and
dimensional stability. The amount of moisture absorption is largely
dependent on the concentration of amide groups in the polymer
backbone. Thus, in general, the higher the amide content, the
higher the equilibrium moisture content at a given relative
humidity. For example, at 23  C and 100 percent relative humidity,
the equilibrium moisture content for nylon 4,6, nylon 6,6, nylon
6,10, and nylon 6,12 was reported to be 15, 8.5, 1.4, and 1.2 weight
percent, respectively [4].
Based on the relatively low amide content of nylon 13,6, it was
expected that this polymer would absorb a relatively low amount of
moisture. To quantify the moisture uptake of nylon 13,6, a technique was utilized that had been previously used to quantify the
moisture uptake for nylon 6 [31], nylon 6,6 [31], and nylon 13,T [37].
As shown in Fig. 13, the moisture absorption after immersion in
water for one week at ambient conditions for nylon 13,6 was much
lower than that for either nylon 6 or nylon 6,6, but not as low as for
nylon 13,T. This results indicates that properties of nylon 13,6
should be much less sensitive to moisture than observed for the
commodity polyamides, nylon 6 and nylon 6,6.

process involving oligomerization using a pressure vessel followed by high temperature melt condensation polymerization.
Characterization of the thermal properties showed that nylon 13,6
possesses a Tm of approximately 206  C and a Tg of approximately
60  C. In addition, the thermal stability was determined to be
adequate enough to enable melt processing without signicant
thermal degradation. Due to the much lower amide content of
nylon 13,6, water absorption was found to be dramatically lower
than the industrially important nylons, nylon 6 and nylon 6,6. This
result indicates that objects produced from nylon 13,6 should have
better dimensional stability in humid environments than objects
derived from nylon 6 or nylon 6,6. Further, the crystallization rate of
nylon 13,6 was shown to be signicantly faster than either nylon 6
or nylon 6,6, suggesting that the polymer may be useful for applications involving processing by injection molding.
Although most oddeeven nylons exhibit a single intense crystallographic reection with a d-spacing of 4.15 
A characteristic of
the g crystal structure, several oddeeven nylons, such as nylon 5,6,
nylon 5,10, nylon 9,2, nylon 11,10, and nylon 11,12, exhibit two
strong diffraction signals with d-spacings similar to those characteristic of the a crystal structure. Since polyamides based on odd
diamines and/or odd dicarboxylic acids cannot form the fully
extended planar zigzag conformation of molecular chains that
produce the parallel planar sheets of hydrogen bonded molecules
that dene the a-form, the crystal structure of these oddeeven
nylons do not t either the a or g-forms. It has been concluded
that these oddeeven nylons that exhibit an a-like diffraction
pattern possess a molecular conrmation in which consecutive
amide groups are rotated slightly to produce structures that
possess two hydrogen bond directions. Nylon 13,6 was found to
exhibit two d-spacings of 4.3 
A and 4.0 
A, which is consistent with
the a-like diffraction pattern analogous to nylon 5,6, nylon 5,10,
nylon 9,2, nylon 11,10, and nylon 11,12, indicating that the polymer
chains within the crystals are rotated slightly to maximize
hydrogen-bonding interactions.
Disclaimer
This report was prepared as an account of work sponsored by an
agency of the United States Government. Neither the United States
Government nor any agency thereof, nor any of their employees,
makes any warranty, express or implied, or assumes any legal
liability or responsibility for the accuracy, completeness, or
usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately
owned rights. Reference herein to any specic commercial product,
process, or service by trade name, trademark, manufacturer, or
otherwise does not necessarily constitute or imply its endorsement,
recommendation, or favoring by the United States Government or
any agency thereof. The views and opinions of authors expressed
herein do not necessarily state or reect those of the United States
Government or any agency thereof.
Acknowledgment
The authors acknowledge the Department of Energy for nancial support through grant DE-FG36-08GO88160.
References

4. Conclusion
To the best of the authors knowledge, this document is the rst
description of the synthesis and characterization of nylon 13,6.
Successful polymerization was accomplished using a two-step

[1]
[2]
[3]
[4]

Carothers WH. J Am Chem Soc 1929;51:2548e59.


Carothers WH, Garvin GA. J Am Chem Soc 1929;51:2560e70.
Kohan MI. Nylon plastics handbook. New York, NY: Carl Hanser Verlag; 1995.
Williams JCL. In: Kohan MI, editor. Nylon plastics handbook. Cincinnati, OH:
Hanser/Gardner Publications; 1995. p. 324.
[5] Datta NC. Chem Ind Dig 2011;24:75e84.

S. Samanta et al. / Polymer 54 (2013) 1141e1149


[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]

Tschan MJ-L, Brule E, Haquette P, Thomas CM. Polym Chem 2012;3:836e51.


Xia Y, Larock RC. Gree Chem 2010;12:1893e909.
Derksen JT, Cuperus FP, Kolster P. Prog Org Coat 1996;27:45e53.
Montero de Espinosa L, Meier MR. Eur Polym J 2011;47:837e52.
Vasishtha AK, Trivedi RK, Das G. J Am Oil Chem Soc 1990;67:333e7.
Azcan N, Demirel E. Ind Eng Chem Res 2008;47:1774e8.
Carlson KD, Sohns VE, Perkins RB, Huffman EL. Ind Eng Chem Prod Res Dev
1977;16:95e101.
Princen LH, Rothfus JA. J Am Oil Chem Soc 1984;61:281e9.
Perkins RB, Roden JJ, Tanquary AC, Wolff IA. Mod Plast 1969;46:136e42.
Sohns VE. J Am Oil Chem Soc 1971;48:362Ae4A.
Nieschlag HJ, Rothfus JA, Sohns VE, Beltron Perkins R. Ind Eng Chem Prod Res
Dev 1977;16:101e7.
Greene JL, Burks RE, Wolff IA. Ind Eng Chem Prod Res Dev 1969;8:171e6.
Greene JL, Huffman EL, Burks RE, Sheeman WC. J Polym Sci Part A 1967;5:
391e4.
Wang Y, Liu M, Wang Z, Li X, Zhao Q, Fu P-F. J Appl Polym Sci 2007;104:
1415e22.
Kinoshita Y. Makromol Chem 1959;33:21e31.
Prieto A, Iribarren I, Munoz-Guerra S. J Mater Sci 1993;28:4059e62.
Kim K-S, Yu AJ. J Appl Polym Sci 1979;23:439e44.
Cui X, Yan D, Xiao C. e-Polymers 2004. no. 068.
Puiggali J, Franco L, Aleman C, Subirana JA. Macromolecules 1998;31:
8540e8.
Morales-Gamez L, Soto D, Franco L, Puiggali J. Polymer 2010;51:5788e98.

1149

[26] Villasenor P, Franco L, Subirana JA, Puiggali J. J Polym Sci Part B: Polym Phys
1999;37:2383e95.
[27] Franco L, Subirana JA, Puiggali J. Macromolecules 1998;31:3912e24.
[28] Franco L, Cooper SJ, Atkins EDT, Hill MJ, Jones NA. J Polym Sci Part B: Polym
Phys Ed 1998;36:1153e65.
[29] Cui X, Yan D. Eur Polym J 2005;41:863e70.
[30] Solomon OF, Ciuta IZ. J Appl Polym Sci 1962;6:683e6.
[31] Kugel A, He J, Samanta S, Bahr J, Lattimer J, Fuqua MA, et al. Polym-Plast
Technol 2012;51:1266e74.
[32] Holmes BS, Moniz WB, Ferguson RC. Macromolecules 1982;15:129e32.
[33] Song JB, Chen QY, Ren MQ, Sun XH, Zhang HL, Zhang HF, et al. Chin J Polym Sci
2006;24:187e93.
[34] Li Y, Zhang G, Yan DJ. Appl Polym Sci 2003;88:1581e9.
[35] Magill JH, Girolamo M, Keller A. Polymer 1981;22:43e55.
[36] Franco L, Puiggali J. J Polym Sci Part B: Polym Phys 1995;33:2065e73.
[37] He J, Samanta S, Sermadurai S, Lattimer L, Ulven C, Sibi M, et al. Green Materials,
in press.
[38] Khanna YP. Polym Eng Sci 1990;30:1615e9.
[39] Avrami M. J Chem Phys 1939;7:1103e13.
[40] Avrami M. J Chem Phys 1940;8:212e25.
[41] Hoffman JD, Weeks JJ. Res Natl Bur Stand US 1962;66A:13.
[42] Hiemenz PC. Polymer chemistry: the basic concepts. New York, NY: Marcel
Dekker; 1984. pp. 227.
[43] Kinoshita Y. Makromol Chem 1959;33:1e20.
[44] Bunn CW, Garner EV. Proc R Soc London 1947;189:39e68.

Das könnte Ihnen auch gefallen