Você está na página 1de 25

Accepted Manuscript

A method to reveal bulk and surface crystallinity of Polypropylene by FTIR


spectroscopy - Suitable for fibers and nonwovens

Franz J. Lanyi, Nicolai Wenzke, Joachim Kaschta, Dirk W. Schubert

PII: S0142-9418(18)30459-8
DOI: 10.1016/j.polymertesting.2018.08.018
Reference: POTE 5580

To appear in: Polymer Testing

Received Date: 20 March 2018

Accepted Date: 16 August 2018

Please cite this article as: F.J. Lanyi, N. Wenzke, J. Kaschta, D.W. Schubert, A method to reveal bulk
and surface crystallinity of Polypropylene by FTIR spectroscopy - Suitable for fibers and nonwovens,
Polymer Testing (2018), doi: 10.1016/j.polymertesting.2018.08.018.

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.
ACCEPTED MANUSCRIPT
A method to reveal bulk and surface crystallinity of Polypropylene by FTIR
spectroscopy - suitable for fibers and nonwovens
Franz J. Lanyi1, Nicolai Wenzke1, Joachim Kaschta1 and Dirk W. Schubert1,2
1
Institute of Polymer Materials, Friedrich-Alexander-University Erlangen-Nürnberg,
Martensstraße 7, 91058 Erlangen, Germany
2
Bavarian Polymer Institute (BPI), Dr.-Mack-Straße 77, 90762 Fürth, Germany
e-mail: Franz.Lanyi@fau.de

PT
Corresponding author: Franz J. Lanyi

RI
Keywords: Fourier-transform infrared spectroscopy; X-ray diffraction; isotactic
Polypropylene; fiber spinning; crystallinity; processing

SC
Abstract

The crystallinity of polymers is usually determined via Differential Scanning Calorimetry


(DSC), X-ray diffraction (XRD) or density measurements. Even though DSC is an easy and

U
fast way to determine crystallinity, it suffers of several disadvantages in the case of fibers and
nonwovens. In DSC studies, the sample preparation of fiber material may be difficult, results
AN
are not calibration free and heat is introduced in the measuring process, which might lead to
transitions in instable crystal structures in the polymer. XRD is able to overcome these
problems but is also time consuming and complicated in terms of sample preparation and
M

evaluation of the measurements. Furthermore, it is not possible to determine surface


crystallinity or crystallinity gradients inside the specimen with both methods.
Based on approaches found in literature, a method was developed which is capable of
D

measuring the degree of crystallinity in the bulk and at the surface of Polypropylene via
Fourier transformed infrared spectroscopy. A broad set of calibration samples with a wide
TE

range of degrees of crystallinity but without a crystallinity gradient was produced by cooling
samples with cooling rates from 1 to 3500 K·min-1. The degree of crystallinity was then
determined from the ratio of the maximum peak height at 974 cm-1 and the maximum peak
height at 998 cm-1 multiplied by 61.4 %.
EP

Consecutively, the transferability of the method to the measurement of fibers and nonwovens
is demonstrated and verified via procedures which are already well described in literature.
This method has an excellent reproducibility, spatial resolution and is capable of determining
C

the degree of crystallinity in less than five minutes. Furthermore, this method is easy to
implement and does not require further sample preparation.
AC

1. Introduction

The production of fibers from a polymer melt is an important process in many industrial
sectors where fibers and nonwovens are utilized as hygiene and medical products, as
geotextiles or as filter materials. The properties of such highly oriented products, like
nonwovens, strongly depend on their crystallinity. As fibers are exposed to high cooling rates
and draw ratios during their production, crystallinity normally is not a bulk property but a
characteristic that changes over the cross-section of the sample. Therefore, bulk properties
often are not informative for the end consumer characteristics like a soft feel, scratch
resistance [1], adhesion of chemicals or blood cells [2] or the migration of additives to the
surface [3], respectively.
ACCEPTED MANUSCRIPT
When looking into literature, usually X-ray diffraction (XRD) [4,5] or Differential Scanning
Calorimetry [4,6,7] (DSC) are the most commonly used methods to determine crystallinity of
polymers. Also density measurements [8], Raman spectroscopy [9,10] and for the
determination of orientation and structure development of crystalline regions, dichroism or
birefringence measurements [11,12] are used. In DSC studies, the sample preparation of fiber
material can be difficult as it can be hard to ensure a good heat conduction between the fiber
samples and the DSC pan [13]. In addition, energy in the form of heat is inevitably introduced
into the sample during the measurement. This may result in a change in the polymers
structure, e.g. due to the conversion of metastable crystal phases in the polymer [14,15].

PT
Moreover, the determination of the absolute value of the crystallinity depends strongly on the
heat of fusion, which can vary by several 10% relatively depending on the method used for
the determination. These problems can be solved by using XRD, but the measurements are

RI
time-consuming and the evaluation of the measured data is often complex. Furthermore, the
standard XRD can only be used to a limited extent to determine the surface crystallinity of
polymers.

SC
The use of IR spectroscopy for the determination of the degree of crystallinity of
Polypropylene is not new and the first applications date back to the 1950s [8,16] but until
today it is mostly applicated to bulk materials or films [8,17–22]. Even though some results
have been published, the calibration of the methods is in many cases at least questionable as

U
most reference methods used to determine the calibration are rather inaccurate. Quynn et al.
[8] used an inaccurate density measurement, compared to an XRD measurement, to validate
AN
the calibration line. Huy et al. [17] have used two single-point DSC measurements for the
determination of the calibration line which is questionable as it lacks of a reliable set of
samples and DSC samples might change their crystallinity due to the heat input during the
M

measurement. In the authors opinion, the only reliable calibration curve for the crystallinity
determination of Polypropylene via FTIR was published by Lamberti et al. [20]. In this paper
a broad range of samples with different degrees of crystallinity was measured via wide angle
X-ray scattering for an isotactic Polypropylene with a molecular weight of 481000 g·mol-1
D

and a polydispersity of 6.4. There are comparatively few studies dealing with the
determination of the crystallinity of fibers using FTIR spectroscopy, as most studies use
TE

birefringence [7,11,12,23,24] or (wide-angle) x-ray scattering (photography) [5,7,12,23,25–


28], both as a qualitative measure for structure and orientation development, as well as X-ray
diffraction [7,23] for quantitative determination of crystalline structures. Recently, Kilic et al.
EP

published a study [29] where FTIR-spectroscopy was utilized to determine the surface
crystallinity of melt spun Polypropylene fibers, but only one data point is shown where FTIR
spectroscopy is utilized to determine the crystallinity of fibers. However, the obtained results
are contradictory to established methods used for the determination of crystallinity. Karacan
C

[30] published a study where FTIR spectroscopy is used to determine the crystallinity of melt
AC

spun Polypropylene fibers, too. Instead of calibrating the method with a calibration free
measuring method, a ratio of the integral of the peaks correlating with crystalline regions and
an integral correlating with the amorphous and crystalline peaks was determined. As there is
no publication on which bands exclusively correlate to amorphous and crystalline regions and
some bands like the peak at 973 cm-1 correspond to both amorphous and crystalline regions,
this method is questionable.
This study shows a novel approach to calibrate FTIR measurements via a comprehensive set
of crystallinity gradient free calibration samples. Specimens were produced at different
cooling rates resulting in a wide range of different degrees of crystallinity from 40 % up to 60
%, reflecting the full spectrum between supercooled samples and the maximum experimental
crystallinity of 63 % [31]. XRD measurements were used to calibrate FTIR-spectroscopy to
reveal surface and bulk crystallinity of the Polypropylene samples. The special application to
ACCEPTED MANUSCRIPT
fibers and nonwovens spun from the melt is shown and used to verify the method’s
applicability.

2. Experimental section

Materials

For this study a fiber spin grade, Moplen HP561R (melt flow index of 25 g·(10 min)-1 at 230

PT
°C with a weight of 2.16 kg), kindly provided by Lyondell Basell, was used. This isotactic
Polypropylene has a weight average of 180 kg·mol-1 and a number average of 91.6 kg·mol-1
resulting in a polydispersity of 2.0. Molecular properties were determined via size exclusion
chromatography using a PL-GPC-220 from Agilent coupled with a multi-angle light

RI
scattering device Dawn Eos from Wyatt Technology.

Preparation of calibration samples

SC
In order to get a reliable calibration curve, it was necessary to use samples that cover a wide
range of crystallinities. Those samples shall not have a crystallinity gradient. Therefore, thin

U
Polypropylene films (d < 300 µm) were utilized, which are placed in a DSC pan, heated up to
200 °C and subsequently cooled under controlled conditions to room temperature.
AN
The Polypropylene films were produced in a LaboPress 200 T hot press from Vogt
Maschinenbau GmbH at a temperature of 240 °C to yield films with an area of (8 x 8) cm and
a uniform thickness of approximately 280 µm. The material was heated for 1 minute without
any pressure, for 2 minutes with a pressure of 10 bar and for additional 2 minutes, a pressure
M

of 200 bar was applied. The resulting film was then cooled between two ceramic plates to
yield films with a uniform surface structure. From these films, samples with a diameter of 4
mm are punched out and put into DSC Tzero Aluminum pans from TA instruments.
D

The samples were then heated up to 200 °C, held for 5 minutes at 200 °C and cooled down via
different procedures. Low cooling rates (1 K·min-1 and 20 K·min-1) were achieved via
TE

controlled cooling programs performed with a DSC Q 2000 differential scanning calorimeter
from TA Instruments. High cooling rates were achieved by heating the DSC pans in a hot air
oven to 200 °C and cooling them in a stirred salt – ice water mixture leading to cooling rates
of approximately 3500 K·min-1. Compared to cooling rates in fiber spinning (approximately
EP

up to 20000 K·min-1) this is still rather low. Cooling rates were determined as described by
Böckh and Wetzel [32]. Samples were produced with and without stirring as well as with and
without subsequent tempering for 60 min at elevated temperatures of 50 °C and 80 °C. A
C

similar annealing procedure which led to an increase of approximately 5 – 10 % in


crystallinity can be found in a publication by Minogianni et al. [33]. An overview of all
AC

produced calibration samples with experimental settings can be seen in Table 1.

Table 1: Overview of the manufacturing methods used for sample preparation for the
calibration set.

Sample ID Cooling method Pan type Cooling rate


1 Salt-ice-water + stirring T zero Aluminum ~ 3500 K·min-1
2 Salt-ice-water T zero Aluminum ~ < 3500 K·min-1
3 Salt-ice-water T zero hermetic ~ < 3500 K·min-1
4 Salt-ice-water + 50 °C annealing T zero Aluminum ~ < 3500 K·min-1 + 50 °C tempering
5 Salt-ice-water + 80 °C annealing T zero Aluminum ~ < 3500 K·min-1 + 80 °C tempering
6 DSC (cooling rate: 20 K·min-1) T zero Aluminum 20 K·min-1
ACCEPTED MANUSCRIPT
7 DSC (cooling rate: 1 K·min-1) T zero Aluminum 1 K·min-1

Fiber spinning

All fiber samples were produced using a multifilament pilot spinning line (Figure 1), which is

spin pump allows to set the throughput  to 0.525 cm³·min-1 per hole. The spinneret consists
fed by a twin-screw extruder. The extrusion temperature was set to 250 °C. After extrusion a

PT
of eight dies with a diameter of 300 µm and a length to diameter ratio of 2. After extrusion,
fibers are quenched in a cooling tunnel with a length of 1.9 m at an air temperature of 12 °C
and a linearly increasing cooling profile towards the aspirator at approximately 65 % relative

RI
air humidity. The fibers are aerodynamically drawn with four gun aspirators and laid down
randomly on a screening belt. The take-up velocity can be adjusted via the applied air


pressure and was calculated via the equation of continuity after spinning:



SC

 
(1)
2⋅⋅ 2 

U
The fiber diameter dfiber for this calculation was determined via a VHX-1000 optical
AN
microscope (magnification of 200) from Keyence from using 30 measurements per setting.
The speed of the belt was adjusted to produce nonwoven samples with a base weight of
20 g·m-2. After laydown, fibers are secured within a frame and released from the sieve belt.
The fibers are then non-continuously calendered via a press. The calender press consists of a
M

converted punch-cutter with a structured plate on the upper part and a flat plate on the lower
part. Both plates are individually heatable via heating cartridges. The structured plate and the
flat plate were both heated to 140 °C as a standard setting. For calendering, the calender press
D

is put into a hydraulic press to apply a pressure ppress of 160 N·mm-² which correlates to a line
pressure in an industry-like calender made of steel (with Young’s modulus Esteel = 210000
TE

MPa, radius r = 300 mm and Poisson’s number ν of 0,3) of approximately 190 N·mm-1
determined via an estimation of the Hertzian stress (Equation (2)) [34].

  ⋅  ⋅  ⋅ (1 − !  )
 =
EP

(2)
#
C

The residence time in the calender is approximately 0.65 s.


AC

Figure 1: Sketch of the pilot spinning plant, which was used for the production of fibers and
nonwovens according to the spunbonding principle.
After production, the fibers and nonwovens were either stored at room temperature (23 °C), in
a hot air oven at 50 °C to promote further crystallisation or in a refrigerator at -12 °C to
prevent any crystallization as the storing temperature is below glass transition temperature of
10 °C.

X-ray diffraction (XRD) measurements

All X-ray diffraction measurements were performed on a Philips Theta-Theta Goniometer,


using CuKα radiation, between 5 – 40 ° with a step size of 0.01 °, each step being measured
ACCEPTED MANUSCRIPT
for 10 seconds. Before evaluation, all data was background corrected via the program EVA
from Bruker with a correction parameter of 0.025. In the XRD measurement, the degree of
crystallinity XC is determined by the ratio of the area of the peaks which is proportional to the
crystalline part and that of the halo which is proportional to the amount of the amorphous part

'()
[35] (Equation (3)).
$% =
'() + '+ , -
(3)

The amount of crystalline and amorphous phases of the samples was determined via a fitting

PT
procedure consisting of a linear combination of Gaussian peak functions (Equation (4)) for the
crystalline phases and a Bigaussian peak function (Equation (5)) for the amorphous phase.
The most crystalline sample (cooled with 1 K·min-1) was evaluated first as the amorphous

RI
halo and the crystalline peaks are most pronounced for that sample. The parameters w1, w2, xc
describing the shape of the amorphous halo of the sample cooled with 1 K·min-1 besides the
peak height h were then utilized for the other samples as the width w1 and w2 and the center

SC
position xc of the scattering of the amorphous halo should not change over all measured
samples.

1 23()(456 )7
'⋅0 87
.(2/) =
U 
(4)
9⋅:
4 ln(2)
AN
456 7
M

?.A 
. (2/) = ℎ ⋅ 0 8B CD (2/ < F( ) GH
456 7 (5)
?.A 
. (2/) = ℎ ⋅ 0 87 CD (2/ ≥ F( )
D
TE

The coefficient of determination R2 was between 0,948 and 0,983 for all evaluated samples
EP

resulting in a typical error of about1 % in crystallinity, which is highly comparable to other


publications [36].
Other evaluation methods, e.g. utilizing a simple Gaussian peak function for the amorphous
halo or utilizing a simultaneous fitting approach (as described in a publication by Carrubba et
C

al. [37]) yielded highly comparable fitting results and calibration curves.
AC

Fourier transformed infrared spectroscopy (FTIR) measurements

All FTIR measurements were carried out on a Nicolet 6700 FTIR spectrometer at 23 °C. Each
measurement consisted of 128 scans at a resolution of 0.48 cm-1. Due to the circumstance that
all peaks which were needed to determine the crystallinity are located in the region between
900 and 1050 cm-1, spectra are recorded in a wave number range from 1400 – 700 cm-1 to
save measuring time and make the method more productive.
For attenuated total reflection (ATR) measurements, a diamond and a germanium single
bounce crystal were used, both with an angle of reflection θ of 42°. For a refractive index of
nDia = 2.4 and nPP = 1.49 a penetration depth of 2.55 µm results at a wave number of 980 cm-1.
The corresponding penetration depth for germanium (nGe = 4.0) is 0.4 µm. In this study, the
IR-radiated volume of both crystals is considered as a surface near region.
ACCEPTED MANUSCRIPT
101 ⋅ JKHMN
 (J)KLMN =
H+ 
2H() ⋅  ⋅ Psin (/) − 
(6)
H()
All measurements were baseline and ATR corrected in order to correct the different
penetration depths according to equation (6) by the program OMNIC 9 from Thermo Fisher.
To perform transmission measurements with single fibers, 5 – 10 mg of the fibers were
pressed in a Graseby Specac manual hydraulic press under vacuum with a mass of 7 tons for 2
minutes at room temperature to yield thin chips with a typical thickness of approximately 10

PT
µm. For transmission measurements of the produced nonwovens no further sample
preparation step was needed. ATR measurements were carried out with the fibers pressed
against the crystal without further sample preparation.

RI
According to Lambert-Beer’s Law the absorbency E at a wavelength λ is proportional to the
penetration depth dp or the IR-radiated volume, the volumetric concentration c of the

SC
substance to be detected and the coefficient of absorption ε:
#(J) = S ⋅ T ⋅  
(7)

U
For the crystallinity Xc a similar equation applies:
AN
#(J) = S ⋅ $( ⋅  (8)
M

If the penetration depth is constant throughout the spectrum, the penetration depth can be
eliminated by building a peak ratio from a peak, which is only dependent on the penetration
D

#J()  SV ⋅ $( ⋅ 
depth:

=
TE

#J  U,  S ⋅ 1 ⋅ 
(9)
 
EP

Rearranging the equation leads to:

#J()  S #J() 
$( = ⋅ =W⋅
#J  U,  SV #J  U, 
(10)
C
AC

The correction factor K has to be determined via a calibration procedure.


A comparable technique for determining the degree of crystallinity via FTIR spectroscopy has
already been published by Burfield et al. [38], Huy et al. [17] or Kilic et al. [29]. For the
quantitation the ratio of the peak height of the peak at 998 cm-1 h998 and the peak height of the
peak at 974 cm-1 h974 is determined. The peak at 998 cm-1 corresponds to the crystalline phase
and the height of this peak grows linearly with the degree of crystallinity [16,22,38]. As the α
and β crystal phase of the Polypropylene have similar spectra [17], the peak height h998 can
therefore be seen as a measure for the overall crystallinity of the sample and cannot be split in
an α and β-phase part. The peak at 974 cm-1 correlates linearly with the penetration depth of
the IR-radiated volume and is independent of the degree of crystallinity [8,19,39]. Therefore,
the ratio of the maximum peak heights of the peak at 998 cm-1 h998 and 974 cm-1 h974 is
ACCEPTED MANUSCRIPT
considered as a linearly proportional measure of the degree of crystallinity XC throughout the

ℎYYZ
following steps (Equation (11)).
$( ~
ℎY[1
(11)

The peak height of both peaks is determined by drawing a linear baseline through the
measured points at 1025 cm-1 and 950 cm-1 and measuring the vertical drop from the highest
measured signal value in the range of 960 cm-1 and 985 cm-1 (for the band at 974 cm-1) and in
the range of 985 cm-1 and 1005 cm-1 (for the band at 998 cm-1). Figure 2 shows a section of

PT
the FTIR spectrum of a rapidly cooled sample, which should have a low crystallinity, and a
spectrum of a sample, which was slowly cooled and therefore should have a high crystallinity.

RI
Figure 2: Section of the IR spectrum for a slowly and a rapidly cooled sample.

SC
3. Method calibration

Since the XRD measurement takes approximately 10 hours to complete, it had to be verified
that during the measurement no changes in the degree of crystallinity, due to relaxation or

U
recrystallization phenomena of metastable crystal phases, occur. Figure 3 illustrates the time
dependent post-crystallisation of one sample from the calibration set.
AN
Figure 3: Time dependent post-crystallization of a rapidly cooled sample. The peak ratio
M

correlates linearly with the degree of crystallinity. Each sample was measured three times
with a diamond crystal.
Accordingly, all samples from the calibration set were measured 2 and 3 days after production
D

via FTIR with the germanium and the diamond crystal to yield information from two different
penetration depths and to secure that there was no change in the degree of crystallinity. Figure
TE

4 shows FTIR results for all calibration samples. It can be seen that all samples besides
number 3 do not have a different degree of crystallinity in a penetration depth of 0.7 µm and
2.55 µm and are expected to be crystallinity gradient free for further measurements. The
EP

gradient in sample 3 can be explained by the use of the hermetic pan which allows the PP film
to directly connect to the bottom side of the pan but not to the upper side which is resulting in
unequal cooling conditions and therefore in a crystallinity gradient.
C

Figure 4: Peak ratio determined via FTIR spectroscopy for all the samples shown in table 1.
AC

After 2 days, XRD measurements were performed to reveal the crystallinity of the samples
from the calibration set without a need to further validate the determination of the degree of
crystallinity as it is already calibration free. When plotting the peak ratio, which was
determined via FTIR measurements, versus the crystallinity, which was determined via XRD
measurements, a calibration line can be determined via a linear fit through the origin (Figure
5). The slope of this line can now be utilized to determine the degree of crystallinity of

ℎYYZ
samples made from this polymer:
$% = (61.4 ± 1.1) % ⋅
ℎY[1
(12)

Huy et al. have found 56 % to be the factor for another Polypropylene. However, they
determined the crystallinity XC through a calibration via DSC measurements [17]. In this
ACCEPTED MANUSCRIPT
study, no heat of fusion for Polypropylene is specified. Too high values of the heat of fusion
may result in an underestimation of the calibration factor K.

Figure 5: Calibration curve to yield the correction factor K via a linear fit of the crystallinity
determined by XRD plotted versus the peak ratio determined by FTIR for the corresponding
calibration samples.

4. Method validation

PT
To verify this method, fibers and nonwovens spun from the melt were produced and stored
under various conditions and phenomena, which are well known in literature. The new
method was applied to determine the ongoing changes.

RI
Figure 6: Influence of storage time and temperature on the post crystallization of the surface
of melt spun Polypropylene fibers.

SC
From Figure 6 it can be seen that post-crystallization in the regions close to the fibers surface
takes place when storing them above their glass temperature. The results show the degree of
crystallinity reaches a saturation level after 2-3 days. This phenomenon has already been

U
discussed in literature [18]. Fichera and Zannetti [6] have shown this phenomenon via DSC
studies of isotactic Polypropylene which was quenched from the melt and subsequently
AN
annealed at elevated temperatures. The maximum plateau value of the degree of crystallinity
is dependent on the annealing temperature. The higher the annealing temperature, the higher
the final crystallinity will be and the faster it is reached. When storing the fibers below the
M

glass temperature no molecular movement of the chains should take place which results in an
inhibited post-crystallization or mesophase transition as it is evident from the constant degree
of crystallinity of the fibers stored at -12 °C.
D

These results show two major advantages of the FTIR method over DSC and XRD for
measuring the crystallinity, especially for rapidly quenched samples like fibers, nonwovens or
TE

films. Due to the introduced heating during DSC measurements, post-crystallization or


mesophase transitions may occur which will lead to falsification of the measurement.
Furthermore, it is non-destructive, that means one and the same sample can be measured
again and again which makes the results much more comparable as a straight baseline can be
EP

drawn via a reproducible reference sample. As the fiber diameter of 15 µm is not changing
over the storage time but a change in the crystallinity can be detected, it can be concluded that
the method is able to reveal the crystallinity independently from the fiber diameter.
C

Figure 7: Time-resolved measurement of post-crystallization of melt spun Polypropylene


AC

fibers directly after fiber laydown.


Figure 7 shows the time-resolved measurement of the post-crystallization of fibers directly
measured after laydown. It can be seen that during the first hour the relative degree of
crystallinity increases by 5 – 10%. These measurements would not be possible with DSC as it
is destructive and both XRD and DSC measurements are not capable of determining the
degree of crystallinity in less than five minutes. An improvement of the measurement
parameters, e.g. decreasing the number of reproduction measurements, reducing the scanned
wave number region or reducing the resolution of the measurement might result in an even
better resolution on the time scale.
ACCEPTED MANUSCRIPT

Figure 8: Influence of calender temperature on the crystallization of bonding points and


fibers of a nonwoven with a base weight of 20 g·m² calendered with a corresponding line
pressure of 190 N·mm-1.
Figure 8 shows the impact of the calendering temperature on the crystallinity of the bonding
point surface, the bonding point bulk and fiber surface crystallinity. The crystallinity of the
bonding point surface and bulk increases due to the introduction of heat, as the polymer can
recrystallize as a result of the input of thermal energy and through the onset of melting at
temperatures above 140 °C [40]. The higher the amount of introduced energy, the more it

PT
recrystallizes. Even though single fibers in between the calendering points are not bonded or
directly exposed to heat, the hot plates lead to a post-crystallization induced by emitting the
heat.

RI
These measurements underline the advantage of the spatial resolution of the method. As the
method is measuring crystallinity only on the contact surface of the crystal (circle with a

SC
diameter of approximately 2 mm). The spatial resolution is even higher when using micro-
ATR-FTIR.

U
Figure 9: Influence of the calender pressure on the crystallinity of bonding points and fiber
surface of nonwovens with a base weight of 20 g·m-2 and a bonding temperature of 140 °C.
AN
Figure 9 shows the impact of the calendering pressure on the crystallinity of the bonding
points and the fibers in between. As expected, there is no impact of the calendering pressure
M

on the crystallinity of the bonding point. The crystallisation behaviour is only dependent on
the calendering temperature. Higher calendering pressures do not affect the crystallization of
the bonding point. An impact on the crystallinity of the fiber surface can be excluded as it is
D

not exposed to any pressure and the temperature of the calendering plate is the same
independent from the applied pressure. Moreover, Figure 9 shows the excellent
TE

reproducibility of the method. Each point represents mean and standard deviation of three
separate measurements.
EP

Figure 10: Influence of haul-off velocity on the bulk and surface crystallinity of fibers for high
and low haul-off velocities.
Figure 10 shows the influence of the haul-off velocity on the crystallization profile of the
C

fibers. It can be seen that regions near to the surface (penetration depth = 0.7 µm) tend to have
a higher crystallinity compared to the regions more distant from the surface (penetration depth
AC

= 2.55 µm). The highest crystallinity can be found when measuring bulk properties. This
leads to the assumption that the spatial distribution of crystallinity along the cross section has
then two local maxima. The first maximum can be found directly at the surface and the
second maximum can be found in the core of the fiber. This finding is in discrepancy with the
current state of the art even though this crystallinity profile has been published and
conclusively explained in literature [27,28]. After extrusion the fiber cools faster on the
outside compared to the core leading to an already solid surface around a core which is still in
the molten state. The cold fiber surface is stretched much more compared to the core which is
leading to strong orientation of the polymer chains resulting in a highly crystalline skin. The
molten core is stretched separately from the already crystalline skin and therefore exhibited to
a strong shear induced crystallization [7,41]. Between surface and core the material is too cold
for a shear or strain induced crystallization in the melt and too far from the surface for a
stretching induced crystallinity which leads to a minimum in the degree of crystallinity.
ACCEPTED MANUSCRIPT
These experiments show the applicability of the determination of crystallinity via properly
calibrated FTIR to retain crystallinity profiles throughout the whole fiber / nonwoven cross
section. Even more points in dependence of the penetration depth could be measured with
other crystals made from different materials or other incident angles.

5. Conclusions

This study shows a novel approach to utilize Fourier Transformed Infrared Spectroscopy in
Attenuated Total Reflection mode (FTIR-ATR) as a highly sensitive, non-destructive and fast

PT
way to measure surface crystallinity of fibers in a penetration depth of 0.5 – 2.5 µm and in
transmission mode to measure bulk crystallinity. Samples without a crystallinity gradient
produced with different cooling rates from 1 – 3500 K/min were measured via XRD and

RI
FTIR-ATR to calibrate the method. Subsequently fibers were produced at different processing
conditions on a pilot spunbond line to verify the method and show its applicability to measure
bulk and surface crystallinity of fibers and nonwovens produced at different processing

SC
conditions.
Further research and optimization procedures which result in shortened measuring times can
lead to efficient tools for process monitoring. Another interesting field of research might open
when using polarizers to also track the orientation development in fibers on- and offline.

Acknowledgments:
U
AN
The authors would like to acknowledge Lisa Freund, Steffen Neumeier and Professor Mathias
Göken from the Department of Materials Science, Institute I: General Materials Properties
from the University of Erlangen-Nürnberg for the possibility to carry out the XRD
M

measurements and for their guidance throughout the experiments.

References
D

[1] E. Moghbelli, R.L. Browning, W.-J. Boo, S.F. Hahn, L.J.E. Feick, H.-J. Sue, Effects of
molecular weight and thermal history on scratch behavior of polypropylene thin sheets,
TE

Tribology International 41 (5) (2008) 425–433.


[2] N. Kawamoto, H. Mori, M. Terano, N. Yui, Blood compatibility of polypropylene
surfaces in relation to the crystalline-amorphous microstructure, Journal of Biomaterials
EP

Science, Polymer Edition 8 (11) (1997) 859–877.


[3] N. Dulal, R. Shanks, T. Gengenbach, H. Gill, D. Chalmers, B. Adhikari, I. Pardo
Martinez, Slip-additive migration, surface morphology, and performance on injection
moulded high-density polyethylene closures, Journal of colloid and interface science 505
C

(2017) 537–545.
AC

[4] D.G.M. Wright, R. Dunk, D. Bouvart, M. Autran, The effect of crystallinity on the
properties of injection moulded polypropylene and polyacetal, Polymer 29 (5) (1988)
793–796.
[5] G. Farrow, The measurement of crystallinity in polypropylene fibres by X-ray
diffraction, Polymer 2 (1961) 409–417.
[6] A. Fichera, R. Zannetti, Thermal properties of isotactic polypropylene quenched from the
melt and annealed, Makromol. Chem. 176 (6) (1975) 1885–1892.
[7] C. Jinan, T. Kikutani, A. Takaku, J. Shimizu, Nonisothermal orientation-induced
crystallization in melt spinning of polypropylene, J. Appl. Polym. Sci. 37 (9) (1989)
2683–2697.
[8] R.G. Quynn, J.L. Riley, D.A. Young, H.D. Noether, Density, crystallinity, and heptane
insolubility in isotactic polypropylene, J. Appl. Polym. Sci. 2 (5) (1959) 166–173.
ACCEPTED MANUSCRIPT
[9] A.S. Nielsen, D.N. Batchelder, R. Pyrz, Estimation of crystallinity of isotactic
polypropylene using Raman spectroscopy, Polymer 43 (9) (2002) 2671–2676.
[10] C. Minogianni, K.G. Gatos, C. Galiotis, Estimation of crystallinity in isotropic isotactic
polypropylene with Raman spectroscopy, Applied spectroscopy 59 (9) (2005) 1141–
1147.
[11] V. Bansal, R.L. Shambaugh, On-Line determination of density and crystallinity during
melt spinning, Polym. Eng. Sci. 36 (22) (1996) 2785–2798.
[12] S. Misra, F.-M. Lu, J.E. Spruiell, G.C. Richeson, Influence of molecular weight
distribution on the structure and properties of melt-spun polypropylene filaments, J.

PT
Appl. Polym. Sci. 56 (13) (1995) 1761–1779.
[13] J. Cao, I. Sbarski, Determination of the enthalpy of solid phase transition for isotactic
polypropylene using a modified DSC technique, Polymer 47 (1) (2006) 27–31.

RI
[14] J. Zhao, Z. Wang, Y. Niu, B.S. Hsiao, S. Piccarolo, Phase transitions in prequenched
mesomorphic isotactic polypropylene during heating and annealing processes as revealed
by simultaneous synchrotron SAXS and WAXD technique, The journal of physical

SC
chemistry. B 116 (1) (2012) 147–153.
[15] S.A. Arvidson, S.A. Khan, R.E. Gorga, Mesomorphic−α-Monoclinic Phase Transition in
Isotactic Polypropylene: A Study of Processing Effects on Structure and Mechanical
Properties, Macromolecules 43 (6) (2010) 2916–2924.

U
[16] W. Heinen, Infrared determination of the crystallinity of polypropylene, J. Polym. Sci. 38
(134) (1959) 545–547.
AN
[17] T.A. Huy, R. Adhikari, T. Lüpke, S. Henning, G.H. Michler, Molecular deformation
mechanisms of isotactic polypropylene in α- and β-crystal forms by FTIR spectroscopy,
J. Polym. Sci. B Polym. Phys. 42 (24) (2004) 4478–4488.
M

[18] R. Zannetti, G. Celotti, A. Fichera, R. Francesconi, The structural effects of annealing


time and temperature on the paracrystal-crystal transition in isotactic polypropylene,
Makromol. Chem. 128 (1) (1969) 137–142.
[19] G. Lamberti, V. Brucato, Real-time orientation and crystallinity measurements during the
D

isotactic polypropylene film-casting process, J. Polym. Sci. B Polym. Phys. 41 (9) (2003)
998–1008.
TE

[20] G. Lamberti, V. Brucato, G. Titomanlio, Orientation and crystallinity in film casting of


polypropylene, J. Appl. Polym. Sci. 84 (11) (2002) 1981–1992.
[21] X. Zhu, D. Yan, Y. Fang, In Situ FTIR Spectroscopic Study of the Conformational
EP

Change of Isotactic Polypropylene during the Crystallization Process, J. Phys. Chem. B


105 (50) (2001) 12461–12463.
[22] L. Li, T. Liu, L. Zhao, W.-k. Yuan, Effect of compressed CO2 on the melting behavior
and βα-recrystallization of β-form in isotactic polypropylene, The Journal of
C

Supercritical Fluids 60 (2011) 137–143.


AC

[23] R.K. Gupta, K.F. Auyeung, Crystallization in polymer melt spinning, J. Appl. Polym.
Sci. 34 (7) (1987) 2469–2484.
[24] H.P. Nadella, H.M. Henson, J.E. Spruiell, J.L. White, Melt spinning of isotactic
polypropylene: Structure development and relationship to mechanical properties, J. Appl.
Polym. Sci. 21 (11) (1977) 3003–3022.
[25] K. Katayama, K. Nakamura, T. Amano, Structural formation during melt spinning
process, Kolloid-Z.u.Z.Polymere 226 (2) (1968) 125–134.
[26] J.R. Dees, J.E. Spruiell, Structure development during melt spinning of linear
polyethylene fibers, J. Appl. Polym. Sci. 18 (4) (1974) 1053–1078.
[27] P.Y.-F. Fung, S.H. Carr, Morphology and deformation of melt-spun polyethylene fibers,
J. of Macromolecular Sc., Part B 6 (4) (1972) 621–633.
[28] P.Y.-F. Fung, E. Orlando, S.H. Carr, Development of stress-crystallized morphology
during melt-spinning of polypropylene fibers, Polym. Eng. Sci. 13 (4) (1973) 295–299.
ACCEPTED MANUSCRIPT
[29] A. Kilic, K. Jones, E. Shim, B. Pourdeyhimi, Surface crystallinity of meltspun isotactic
polypropylene filaments, Macromol. Res. 24 (1) (2016) 25–30.
[30] I. Karacan, H. Benli, The use of infrared-spectroscopy for the structural characterization
of isotactic polypropylene fibres, Tekstil ve Konfeksiyon 21 (2011) 116–123.
[31] D.W. van Krevelen, K.t. Nijenhuis, Properties of polymers: Their correlation with
chemical structure ; their numerical estimation and prediction from additive group
contributions / D.W. van Krevelen, 4th ed., Elsevier, Amsterdam, Boston, 2009.
[32] P. Böckh, T. Wetzel, Wärmeübertragung: Grundlagen und Praxis, 5th ed., Springer
Berlin Heidelberg; Imprint; Springer Vieweg, Berlin, Heidelberg, 2014.

PT
[33] C. Minogianni, K.G. Gatos, C. Galiotis, Estimation of crystallinity in isotropic isotactic
polypropylene with Raman spectroscopy, Applied spectroscopy 59 (9) (2005) 1141–
1147.

RI
[34] H. Hertz, Ueber die Berührung fester elastischer Körper, Journal für die reine und
angewandte Mathematik 92 (1882) 156–171.
[35] A. Weidinger, P.H. Hermans, On the determination of the crystalline fraction of isotactic

SC
polypropylene from x-ray diffraction, pp. 98–115.
[36] M. van Drongelen, T.B. van Erp, G.W.M. Peters, Quantification of non-isothermal,
multi-phase crystallization of isotactic polypropylene: The influence of cooling rate and
pressure, Polymer 53 (21) (2012) 4758–4769.

U
[37] V. La Carrubba, S. Piccarolo, V. Brucato, Crystallization kinetics of iPP: Influence of
operating conditions and molecular parameters, J. Appl. Polym. Sci. 104 (2) (2007)
AN
1358–1367.
[38] D.R. Burfield, P.S.T. Loi, The use of infrared spectroscopy for determination of
polypropylene stereoregularity, J. Appl. Polym. Sci. 36 (2) (1988) 279–293.
M

[39] H. Tadokoro, M. Kobayashi, M. Ukita, K. Yasufuku, S. Murahashi, T. Torii, Normal


Vibrations of the Polymer Molecules of Helical Conformation. V. Isotactic
Polypropylene and Its Deuteroderivatives, The Journal of Chemical Physics 42 (4)
(1965) 1432–1449.
D

[40] T. Nishino, T. Matsumoto, K. Nakamae, Surface structure of isotactic polypropylene by


X-ray diffraction, Polym. Eng. Sci. 40 (2) (2000) 336–343.
TE

[41] C. Zhang, H. Hu, D. Wang, S. Yan, C.C. Han, In situ optical microscope study of the
shear-induced crystallization of isotactic polypropylene, Polymer 46 (19) (2005) 8157–
8161.
C EP
AC
D
M
A
TE
D
M
AN
U
ED
M
AN
ED
M
AN
U
TE
D
M
AN
U
ED
M
AN
U
TE
D
M
AN
U
TE
D
M
AN
U
TE
D
M
AN
U
TE
D
M
AN
U
TE
D
M
AN
U
ACCEPTED MANUSCRIPT
A method to reveal bulk and surface crystallinity of Polypropylene by
FTIR spectroscopy - suitable for fibers and nonwovens
Highlights

- Development of a method, which is able to measure surface and bulk crystallinity of


Polypropylene
- Method is capable of measuring surface and bulk crystallinity of fibers in a non-destructive,
location-sensitive and fast way via FTIR spectroscopy

PT
- Surface and bulk crystallinity of fibers and nonwovens produced under different processing
conditions is determined
- Influence of the calendaring process on bonding points and fibers spun from the melt is

RI
shown
- Influence of storage time and temperature on fiber crystallinity and post-crystallization is
revealed

U SC
AN
M
D
TE
C EP
AC

Você também pode gostar