You are on page 1of 14

Polymer Foaming with Chemical Blowing Agents:

Experiment and Modeling

Jose´ Antonio Reglero Ruiz,1 Michel Vincent,1 Jean-Franc¸ois Agassant,1 Tarik Sadik,2,3 Caroline Pillon,2,3
Christian Carrot 2,3
1
MINES ParisTech—Centre de Mise en Forme des Mate´riaux (CEMEF), UMR CNRS 7635, 1, rue Claude
Daunesse, CS 10207, 06904—Sophia Antipolis Cedex, France
2
Inge´nierie des Mate´riaux Polyme` res (IMP), UMR CNRS 5223, Universite´ de Saint-Etienne, Jean Monnet, F-
42023, Saint-Etienne, France
3
Universite´ de Lyon, F-42023, Saint-Etienne, France

An experimental and theoretical analysis of the polypro- bubble growth process following nucleation, and the coalescence
pylene foaming process using three different chemical
blowing agents (CBA) was performed. A simple experi- during expansion.
ment was designed to analyze the foaming process of Different steps can be considered when processing such
polypropylene (PP)/CO2 system under two different pres- materials. First the polymer/gas solution formation, second the
sure conditions. The expansion ratio and final foam struc- microcellular nucleation, and finally the cell growth and the
ture was measured both by direct observation and from resulting density reduction. In the first stage, the polymer/gas
optical measurements and image analysis, showing a solution formation is accomplished by saturating a polymer
good agreement. A single bubble simulation based on rel-
under a high gas pressure, forming a single-phase supersaturated
evant differential scanning calorimetry and thermo-
gravimetrical analysis experiments, assuming each CBA solution governed by the gas dissolution in the polymer matrix
particles as a nucleation site and accounting for gas diffu- which is a function of pressure and, at a lesser extent, of tem-
sion in the surrounding polymer matrix has been built. perature. Numerous studies have been carried out analysing the
The sensitivity of the model to physical and processing dependence of the solubility of gas in several polymers, mainly
parameters has been tested. The calculation results are CO2 in polypropylene [1–3].
compared to the experiments and open the route to a In the second stage, it is necessary to submit the polymer/gas
simplified method for evaluating the efficiency of CBA.
POLYM. ENG. SCI., 00:000–000, 2014. VC 2014 Society of solution to a thermodynamic instability to nucleate microcells.
Plastics This nucleation can be achieved by lowering the solubility of
Engineers the solution through the temperature and the pressure of the sys-
tem. Usually, a rapid pressure drop produces a high nucleation
rate in the polymer matrix and in the ideal case this nucleation
INTRODUCTION occurs instantaneously.
The foaming process to produce microcellular thermoplastics In plastic foaming, nucleation refers to the process of generat-
has been widely analysed in the last decades. Microcellular plas- ing gas bubbles in a polymer melt through a reversible thermody-
namic process. In the classical nucleation theory, there is a
tics are generally formed by cell nucleation and growth of bub-
critical nucleus, which defines the minimum radius for a bubble
bles in the polymer matrix. Chemical blowing agents (CBA) or
to growth. Nucleated bubbles which size is larger than the critical
physical blowing agents (PBA) are used to introduce the gas
nucleus radius will survive, whereas those smaller will collapse.
that creates the cellular structure. This work is focused on CBA
Cell nucleation can occur homogenously or heterogeneously. The
foaming process. heterogeneous nucleation is usually 100 to 1000 times more
A typical polymer foaming process involves several steps: favourable than homogeneous nucleation. Several additives or
first, the dissolution under an elevated pressure of a gas created organic charges can be employed as nucleation sites in polymer
by a chemical reaction from (CBA), or gas dissolved in supersa- foaming processes [4, 5].
turated state in the molten polymer (PBA). Second, the nuclea- Once the cells are nucleated, they continue to expand by dif-
tion of a population of gas clusters in the supersaturated fusion of the dissolved gas from the polymer matrix into the
solution upon the release of pressure to the ambient pressure bubbles. In this stage of the process, a deep knowledge of the
and finally the growth of nucleated bubbles in the polymer to physical parameters that govern the diffusion properties of gas-
their ultimate equilibrium size. The final foam density depends polymer systems is necessary [6–9]. The cells grow reducing
on the original gas loading, the gas fraction which remains dis- the polymer density as the gas molecules diffuse into the
solved in the polymer matrix when it solidifies, the gas losses to nucleated cells. The rate at which the cells grow is limited by
the environment, and the depressurization rate. The cell size and the diffusion rate and the rheology of the polymer/gas solution.
cell size distribution depend on the kinetics of nucleation, the The cell growth process is controlled also by the time allowed
for the cells to grow before solidification, the temperature and
pressure of the system, the presence of other bubbles, and so
forth [10].
Correspondence to: Jos´e Antonio Reglero Ruiz; e-mail: jose-
antonio.reglero_ ruiz@mines-partistech.fr One of the main foaming processes involves CBA, which lib-
DOI 10.1002/pen.24044 erate gases under certain processing conditions either due to
Published online in Wiley Online Library (wileyonlinelibrary.com). chemical reaction or thermal decomposition. Most CBAs pro-
VC 2014 Society of Plastics Engineers duce nitrogen (N2) or carbon dioxide (CO2) after decomposition

POLYMER ENGINEERING AND SCIENCE—


2014
Thermal characterization was carried out to determine the
polypropylene fusion temperature and the decomposition tem-
peratures of each reaction of the reactive elements included in
the CBA pellets. Results are presented in Fig. 1. A heating rate
of 6○C/min was chosen to assure the same heating rate as in the
experimental foaming expansions.
The curve for PP shows the fusion between 160 and 170 ○C.
For the three CBAs, the small peak around 90○C corresponds to
the PE melting temperature. For the CBA-1 and CBA-3, the
peaks between 150 and 175○C correspond to sodium bicarbonate
decomposition reaction. A peak between 190 and 220○C caused
by the decomposition of citric acid is visible for CBA-2 and CBA-
1. Finally, a small peak about 240○C takes place in the CBA-1,
caused by the coupling reaction of the remaining sodium
FIG. 1. DSC curve of the chemical foaming agents and of the polypropyl- bicarbonate and citric acid. This third reaction is only important at
ene compound. Constant heating rate of 6○C/min, from ambient temperature higher heating rates (above 20○C/min). For low heating rates, the
to 250○C. importance of this reaction is negligible, as it can be seen in Fig.
1. All the decomposition reactions of the CBAs start after the
[11–13]. CBAs reactions can be endothermic or exothermic. polypropylene fusion, which assures that the gas obtained from
Azodicarbonamide is the most representative exothermic CBA, the CBA can be diluted in the melted matrix.
commonly having a high gas yield, with decomposition temper- TGA measurements were carried out to determine the quan-
atures between 170 and 200 ○C [14]. Sodium bicarbonate and tity of gas released by the reactive elements in the CBA par-
zinc bicarbonate are the most common endothermic blowing ticles (Fig. 2). The relative weight loss refers to the original
agents [15]. weight of the granule containing 30 wt% of polyethylene and 70
In this article, a simple experiment has been designed to ana- wt% of reactive elements. This loss is associated to the gas
lyse the foaming expansion as a function of time of a polypropyl- escaping the sample, assuming that the pressure conditions do
ene containing three types of CBA, in static conditions (no flow). not allow any gas dissolution in the polyethylene.
The expansion ratio has been measured by direct observation and The TGA curve of CBA-3 shows that the decomposition
from optical measurements and image analysis. A single bubble reaction of sodium bicarbonate begins at 150 ○C, reaching a
simulation based on differential scanning calorimetry (DSC) and weight percentage of created gas around 24% at the end of the
thermo-gravimetrical analysis (TGA) experiments, assuming each reaction, at about 210○C. In the case of CBA-2, the decomposi-
CBA particle as a nucleation site and accounting for gas diffusion tion reaction of citric acid begins at 215 and ends around
in the surrounding polymer matrix has been built. The sensitivity 300○C, with a weight percentage of gas created about 35%.
of the model to physical and processing parameters has been Finally, the CBA-1 presents a first decomposition reaction
tested and the results are compared to the experiments. which begins at 150○C. The citric acid decomposition reaction
is probably coupled with the sodium bicarbonate reaction, as it
can be observed in the small change in the slope about 200 ○C.
EXPERIMENTAL
The maximum quantity of gas generated at the end of the
Materials decomposition reactions is about 28 wt% at 300 ○C. In all the
The polypropylene compound was a 12% mineral filled (5% cases, the PE decomposition begins at 450○C.
wt. talc et 7% wt. fibers), elastomer modified polypropylene
(SUMIKA PP) with a melt flow index of 65 g/10 min (ISO
R1133), a Newtonian plateau viscosity of 500 Pa s· at 210 ○C,
determined from rotational rheological measurements, and a
density of 0.91 g/cm3. Three different endothermic CBA
referred as CBA-1, CBA-2, and CBA-3 were analysed. These
foaming agents are PE-based compounds with reactive elements
(typically citric acid, sodium bicarbonate, or a mix of both com-
ponents). The CBA-1 was composed of 35 wt% of citric acid
and 35 wt% of sodium bicarbonate. The CBA-2 was composed
of 70 wt% of citric acid and finally the CBA-3 was based in 70 wt
% of sodium bicarbonate. In all the cases, the percentages
represent the weight concentration respect to PE matrix. In the
following, CBA refers to the compound and not to the reactive
elements only [16–20].
Chemical foaming agents have been extensively employed in
the last few years [16–20]. These products, with decomposition
temperatures between 160 and 210○C, can be added directly into FIG. 2. TGA curve of the chemical foaming agents (heating ramp 10○C/
the hopper of an injection moulding machine in the form of pel- min from ambient temperature to 600○C).
lets in proportions from 1% to 4 wt%.

2 POLYMER ENGINEERING AND SCIENCE— DOI


FIG. 3. Quantity of gas created during the decomposition of the reactive FIG. 4. SEM micrograph of the precursor foaming material.
elements in the CBA particles.

It is possible to obtain the evolution of the total quantity of portion of 98 wt% of polypropylene and 2 wt% of CBA. Then,
moles of created gas from the stoichiometry of the decomposi- cylindrical samples of 9 mm height (hi) and 20 mm diameter (/)
tion reactions. The sodium bicarbonate decomposition is (Reac- were fabricated in a steel mold by compression under a
tion 1): pressure of 20 MPa at 60○C for 30 min. A total number of six
samples, three groups of two samples with the same CBA, were
2NaHCO3 ! Na2CO31H2O1CO2 (1) fabricated, with a volume Vi of 2.83 cm3. The density of all the
The citric acid decomposition is (Reactions 2 and 3): samples (qsample) was about 0.905 g/cm3, with a densification value
up to 99.9%, calculated as:
C6H8O7 ! C6H6O61H2O (2)
C6H6O6 ! C5H6O41CO2 (3)

Finally, a coupling reaction occurs when the sodium bicar-


bonate and citric acid react together to produce CO 2 and H2O
(Reaction 4):

C6H8O713 · NaHCO3 ! Na3C6H5O713 · H2O13 · CO2


(4)

In CBA-3 and CBA-2 only Reactions 1 and 2 occur, respec-


tively. As our experiments were carried out at T_ < 20○C/min,
the coupling Reaction (4) for CBA-1 is neglected.
One gram of CBA contains 0.7 g of reactive element. The
gas escaping the sample is composed of CO 2 (molar mass of
48 g) and H2O (molar mass of 18 g). From the stoichiometry of
the reactions it is then possible to obtain the evolution of the
number of moles of gas per gram of CBA generated during the
decomposition reaction (Fig. 3). It is expected that after foaming
and cooling of the samples, the H2O remains in the samples as
condensed water vapor.

Foaming Experiment
The objective is to analyze the foaming behavior of different
CBA/PP samples, obtained by mixing the three CBAs with PP
granules in the same proportions as in injected samples, usually
between 1 and 4% wt. These materials are foamed in a steel
mold and the expansion ratio and cellular structure will be
analyzed.

Sample Preparation and Characterization. The challenge is to


mix the components without activating the chemical reactions.
FIG. 5. Calculation of the CBA particle average size: (a) binarized image
A solid sample was fabricated starting from PP and CBA pow- and (b) resulting histogram.
ders. Powders were mixed at room temperature in a fixed pro-

DOI POLYMER ENGINEERING AND SCIENCE—2014


XN n l
0 i i
l 5 Xi (8)
n
0
N
i i
The determination of l from Eq. 8 gives a result of 74
0
mm. Another way for calculating l is to assume a uniform
distribu- tion of the reactive agent particles in the solid sample
using crystallographic considerations. Considering for example a
prim- itive cubic (PC) Bravais lattice, the number of reactive agent
particles in the cubic cell (Ncell) is 1. The determination of the
0 P
distance l is related to the volume of the solid sample Vi,
the number of reactive agent particles in each cubic cell Ncell and
P
the total number of reactive agent particles NP through the
expression:
FIG. 6. Resulting histogram of the calculation of distance between 03

particles. l NP
cell 5i (9)
. Σ NP
qsample
Densification 5 100 (5) Similar calculations can be carried out for the body-centered
3 qPP
cubic and face-centered cubic bravais lattices, in which the
taking qPP 5 0.91 g cm·
23
as the density of the solid PP. number of reactive particles in each cubic cell is 2 and 4,
Figure 4 presents a typical SEM micrograph of the sample respectively. Taking Np 5 7.45 3 106 cm23 and Vi 5 2.83 cm3,
surface showing the distribution of the foaming agent (white the average initial distance between reactive agent particles cal-
particles) in the polypropylene matrix. The average particle size culated ranges from 76 to 87 mm, which is similar to the value
0
can be calculated by image analysis, using the ImageJVR obtained from the binarized image. This distance l will increase
soft- ware. The minimal observable size is 1 mm. After during the expansion process.
binarization of SEM images (Fig. 5a), the apparent particle
radius distribu- tion is obtained and represented as a histogram Foaming Experiment Under Pressure. Figure 7 presents the
(Fig. 5b). scheme of the foaming expansion experiment designed to ana-
The number average radius R¯ and the number average lyze the behavior of the polypropylene/gas systems. The solid
vol- sample, with a number of reactive agent particles NP, has an ini-
ume V can be obtained assuming a spherical geometry of the tial height hi and an initial volume Vi. It is placed in a steel res-
CBA particles, which is far from reality (Eqs. 6 and 7):
R¯ 5 XN
i51 niRi (6) ervoir (height 200 mm) under an external pressure P ext . The
XN ni inner diameter is 20 mm. The external pressure is applied with
i51 a weight deposited on a circular steel cap, with a diameter
X N n R3 slightly lower than the inner cylinder diameter. The device is
i i 4p introduced in an oven at a temperature T during a time t. A type
V 5 i51 (7)
XN n 3 K thermocouple is introduced in the sample to monitor the evo-
i
i51
lution of the temperature. Another one is located inside the
where N represents the total number of particles (N 5 1800). to calculate the number average value l .
0

The average value of the reactive agent particle radius is


R¯ 5
4.46 mm and the average volume is V¯ 2.80 3 1029 cm3.
This calculation was performed in three different SEM
micrographs, with a dispersion value of 610%.
Knowing that each CBA particle contains 70 wt% of reactive
agent and 30 wt% of low-density PE, and using the mixing law
(with qreactive agent 2 g/cm
=
3
and qLDPE 5 0.9 g/cm3), the den- sity
of a CBA particle is 1.67 g cm 23, ·which leads to an aver- age
mass of a reactive agent particle of 4.70 3 1029 g. The total
mass of reactive agents in the sample (mass of 2.5 g) is obtained
knowing that each sample contains 2 wt% of CBA, in which the
reactive agents represent 70 wt% (0.035 g). The total number of
reactive agent particles (Np) in the solid sample is obtained simply
by the ratio between the total mass of reactive agent in the sample
and the average mass of a reactive agent particle: 0.035/4.70 3
1029 5 7.45 3 106.
It is also possible to obtain the distance between each parti-
cle (l), from the previous binarized image (Fig. 5a). The result-
ing histogram is presented in Fig. 6, together with the expression

4 POLYMER ENGINEERING AND SCIENCE— DOI


oven. At the end of the heating step, the device is removed
from the oven and cooled down at ambient air. When cooling is
finished the weight is removed. The foamed sample has a final
height hf, a final volume Vf, and a cellular structure with a cell
density Nc and an average cell radius R. The height increase
during the foaming process is measured optically with a record-
ing camera, with a precision of 60.5 mm. It is important to
mention that the temperature will vary differently as a function
of time at different points of the sample, due to heat conduction
from the oven. This means that the foaming develops heteroge-
neously inside the sample, starting in the sample edges and
propagating towards the core. The evolution of the temperature
with time is also represented. The sample and oven temperatures
are not equal until the last part of the experiment. The average
heating ramp of the sample temperature is about 6○C/min, con-
sidering a total heating time of 32 min from ambient tempera-
ture to 210○C.
Two external pressures were tested to evaluate their influence
on the expansion rate, gas diffusion and final pore radius: 0.25
MPa (Samples 1-1, 2-1, and 3-1) and 0.5 MPa (Samples 1–2, 2-
2, and 3-2).

4 POLYMER ENGINEERING AND SCIENCE— DOI


FIG. 7. Scheme of the foaming expansion experiment and evolution of temperature and pressure with time. [Color
figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

Results and Discussion. The measurement of the height varia- It is important to estimate the volume of gas retained in the sam-
tion started when the sample temperature reached 180 ○C, ple to analyze possible gas diffusion outside the polymer. Assuming
slightly above the melting point of the PP, which is located that the decomposition reaction is complete, the theoretical total vol-
between 160 and 170○C according to DSC data (see Fig. 1). ume of gas Vgenerated can be calculated from perfect gas equation.
g
Figure 8 presents the samples height evolution from the begin- total
ning of the expansion measurements (about 23 min after starting nðtÞ · R · T
the experience, see Fig. 7). V
generated
5 P (10)
gas ext
All samples show a remarkable volume expansion due to the
gas creation and the foaming process. The expansion process
takes about 4 min for CBA-1 and CBA-3, and 6 min for CBA-
2, leading to expansion ratios around 200% with Pext 5 0.25
MPa and 140% with Pext 5 0.5 MPa. After that time, the sample
height reduces because of cooling and thermal shrinkage.
The samples foamed with the CBA-2 showed a higher expan-
sion than samples foamed with the other two CBAs. This may
be explained by the larger number of moles of gas generated by
the chemical reactions, as shown on Fig. 3.
At the end of the process, shrinkage is clearly seen, espe-
cially in samples foamed with CBA-2. No shrinkage was
observed in the lateral direction. For CBA-2 at 0.25 MPa,
height reduces from 22.1 mm at 200 s to 20.1 mm at 260 s
after (both times measured after starting the expansion at 23
min), which corresponds to a volume reduction of 10%. For
the same CBA at 0.5 MPa, the height reduction begins at
170 s with a maximum value of 14.8 mm and a final value of
13.3 mm at the end of the experience, with a similar volume
reduction value. FIG. 8. Evolution of the height of each sample during the expansion process.

DOI POLYMER ENGINEERING AND SCIENCE—2014


TABLE 1. Physical parameters and theoretical total gas volume generated during the expansion process for each sample.

Sample CBA Pext (MPa) n(t)total (mol) Vgenerated (cm3) Vf (cm3) Expansion ratio q (g/cm3)
gas

1-1 1 0.25 3.22 3 1024 5.17 5.62 1.94 0.47


1–2 0.5 2.58 3.98 1.38 0.65
2-1 2 0.25 4.72 3 1024 7.58 6.31 2.18 0.41
2-2 0.5 3.79 4.17 1.44 0.63
3-1 3 0.25 3.22 3 1024 5.17 5.84 2.02 0.45
3-2 0.5 2.58 3.92 1.36 0.66

Pext is the external pressure (MPa), T is the foaming tempera-


Using the ImageJV
R
software previously presented, it is
ture, which is taken as a constant value of 210○C (483 K) in a first
possi- ble to quantify the average bubble radius Rbubble, the average
approximation, and is the gas constant (8.31
distance between bubbles, (namely lbubble), and the total volume of
J mol21 K21). The value of n(t)total<corresponds to the total number
gas in each sample Vgas. As an example, Fig. 10 presents the
of· moles of gas created in the solid sample for the dif- ferent
histogram of both bubble radius, assuming that each bubble can
CBAs at the end of the reaction.
be considered as a sphere (Fig. 10a), and the distance between
According to Fig. 3, the total number of moles of gas created
bubbles (Fig. 10b). This parameter is calculated for each bubble,
per gram of CBA is 0.0092 mol/g in the case of the CBA-2 and
taking the minimum distance between the edge of this bubble
CBA-3, and 0.0135 mol/g in the case of the CBA-1. The total
and the surroundings ones. Results presented have been calcu-
number of moles of gas generated in the sample n(t)total can be
lated from Test 1-1, but similar calculations have been per-
calculated by multiplying the previous values and the mass of
formed for the other samples.
CBA particles in each sample, in our case 2 wt% of 2.5 g, and
The average results are obtained from 10 micrographs of
assuming that only the 70% of each CBA particle contains reac-
each sample with a maximum error of 65%.The software
tive elements.
accounts for the number of bubble in each image and the aver-
The gas volume generated at the end of the expansion, before
age radius. The total cell number Nc in the sample was calcu-
any cooling and thermal shrinkage of the polymer is deduced
lated using Eq. 11 [21], which accounts for the 3D extrapolation
from Eq. 10, and represented on Table 1, together with the final
starting from a 2D image:
volume Vf, the expansion ratio after cooling and the foam den- . Σ
sity q. q
6 12 F
The fraction of gas in the foamed polymer can be also deter- N c5 qP (11)
mined from image analysis. Several optical micrographs of the Vf
pR3
fracture surface of the expanded samples are presented in Fig. 9. qF represents the foam density, and qP the solid polymer den-
The structure is homogenous throughout the whole sample vol- sity (0.91 g/cm3). Vf represents the final volume of the sample
ume, which indicates that the thermal gradients do not influence after shrinkage. The experimental gas volume can be also easily
the final foamed structure, even when the heating ramp may be calculated from expansion measurements, following the next
locally different in the sample. equation:
Figure 9 shows that CBA-2, containing citric acid as reactive
element, induces a finer bubble size and higher expansion rate Vdirect5Vf 2Vi (12)
than the other CBAs (see Fig. 9b). The addition of sodium g

Both results (indirect method V indirect , from ImageJV


R
bicarbonate (CBA-1) or CBA based only on sodium bicarbonate
g
(CBA-3), leads to larger bubble size but lower expansion rate. analysis and direct method Vdirect, from expansion measurements),
g

FIG. 9. Optical micrographs of the expanded samples: (a) Sample 1-1; (b) Sample 2-1; (c) Sample 3-1.

6 POLYMER ENGINEERING AND SCIENCE— DOI


the remaining dissolved gas, with negative results. This indicates
that a small proportion of the gas has probably diffused out of
the samples.
Another interesting parameter that can be analyzed from the
foaming experiment is the coalescence. It is possible to define
the ratio k between the final number of cells Nc and the initial
number of reactive agent particles NP. It lies between 0.2 and
12%, indicating that coalescence is a very important phenom-
enon that will be discussed lately.

THEORETICAL APPROACH
Many studies have been devoted to the development of
numerical models for the bubble nucleation and growth in poly-
meric foaming process. In the classical work presented by
Amon and Denson [22] a complete mathematical analysis of a
bubble growth in a Newtonian matrix is presented. Bikard et al.
[23] and Bruchon [24] solved the same problem with a 3D finite
element method which allows accounting for the simultaneous
growing of multiple bubbles. Koopmans et al. [25] introduced a
viscoelastic multimode Maxwell behavior for the polymer
matrix in a “bubble influence volume” surrounding the growing
bubble. They also account for non-isothermal phenomena occur-
ring at die exit in an extrusion process. Otsuki and Kanai [26]
introduced a more realistic Phan-Thien Tanner viscoelastic con-
stitutive equation which limits the dramatic increase of the Max-
well model elongation viscosity. Shafi et al. [27, 28] and Joshi
et al. [29] developed a homogeneous nucleation model that they
couple to the Newtonian Amon and Denson bubble growing
model. Taki [30, 31] compares these calculation results to
FIG. 10. Resulting histograms of the SEM micrograph presented in Fig. experiments performed under several pressure release rates.
9a. (a) Bubble radius and b) distance between bubbles. Feng and Bertelo [32] investigated bubble nucleation in a visco-
elastic polymer melt (Oldroyd B constitutive equation) contain-
together with the morphological determinations, are presented ing nucleating agents.
in Table 2. They are in the same range, and the slight differen- All these works assume that the polymer is saturated with
ces between both measurements can be due to specific errors gas that diffuses from the matrix to the bubble. In our case, the
associated to the software employed to analyze the optical diffusion process is in the opposite direction from the bubble,
micrographs. The last column of Table 2, Vremaining, represents the which is supposed to be nucleated around the CBA particle,
C
total volume of CO2 remaining. This value is calculated assuming towards the polymer matrix. Recently Emami et al. [33] ana-
than after cooling and shrinkage H 2O condensates into liquid lyzed the bubble nucleation in non-pressurized foaming CBA
water, and extracting the volume of water vapor generated from systems starting from solid materials, composed of PP and CBA
the total volume of gas generated calculated in Table 1. From the powders as in our experiments. It was observed that the nuclea-
chemical Reactions (1), (2), and (3) both citric acid and sodium tion process proceeded in two distinct stages, namely primary
bicarbonate decompose in one mole of H 2O and one mole of CO2. and secondary nucleation. Primary nucleation occurred in the
The ratio between the molar masses of both components is interstitial regions of the sintered plastic powder and the
18/(18 1 44) 5 0.29, indicating that the 29% of the gas agglomerated blowing agent particles acted as nucleation sites,
generated transforms into liquid water by condensation. and secondary nucleation occurs in the polymer melt. The visual
The proportion between the measured gas volume and the observations indicated that most of the first generation of bub-
theoretical gas volume varies between 50 and 75%. bles endured the entire foaming process, whereas most of the
The order of magnitude of the thermal shrinkage between bubbles generated during secondary nucleation disappeared over
210○C and room temperature for the polypropylene is around time. These results support the previous assumption which con-
20%. This cannot explain the measured difference. An incom- siders each reactive agent particle as a nucleation site, with no
plete chemical reaction is unlikely according to the DSC and further nucleation phenomenon.
TGA measurements. It could be speculated that a part of the gas In the following, a kinetic model for a single bubble expan-
generated has not been nucleated and does not produce any sion in a Newtonian fluid coupled with the gas diffusion in the
expansion. It could remain dissolved in the polymer matrix or surrounding polymer matrix is proposed. The nucleation phe-
have diffused outside the sample during the expansion process. nomenon will be simplified, assuming that each reactive agent
To test these hypotheses, the foamed samples were re-heated up particle can be considered as a nucleation site. The model will
to 210○C to analyze a possible second expansion produced by be applied to the experimental conditions presented in the previ-
ous section. The sensitivity of the model to several unknown

DOI POLYMER ENGINEERING AND SCIENCE—2014


TABLE 2. Gas volume and morphological parameters in the expanded samples.

Test CBA Bubble count Rbubble (mm) 3


Vbubble (mm ) Lbubble (mm) qF (g/cm3) Nc (cm23) Vindirect (cm3) Vdirect (cm3) Vremaining (cm3)
gas gas CO2

4
1-1 1 64 360 0.195 50 0.47 1.43·104 2.89 2.73 3.77
1–2 1 108 170 0.021 42 0.65 8.97·10 1.19 1.09 1.88
5
2-1 2 241 120 0.007 25 0.41 4.30·10 3.11 3.42 5.53
5
2-2 2 292 85 0.003 23 0.63 7.58·10 1.35 1.28 2.76
4
3-1 3 79 350 0.180 56 0.45 1.65·10 2.90 2.95 3.77
4
3-2 3 121 190 0.029 38 0.66 6.24·10 1.19 1.03 1.88

parameters has been tested. Comparison between calculation and later, in terms of the gas diffusion distances, bubble radius
experiments will be discussed. and sample dimensions.

The first step consists in the determination of the number of


Single Bubble Growth Model moles of created gas n(t)created. Three different CBA kinetics are
A schematic of the bubble growth model is shown in Fig. shown in Fig. 3. Considering the experimental foaming results,
11. A reactive agent particle creates n(t)created moles of gas by the two different CBAs, CBA-1, and CBA-3, present similar results
chemical decomposition reaction deduced from the TGA curve. in terms of expansion rate and morphology. This indicates that
The bubble growth is governed by the competition between the the coupling reaction between citric acid and sodium bicarbon-
gas which remains within the bubble and induces the growing ate has not occurred at the foaming temperature of the experi-
mechanism (number of moles n(t)) and the moles of gas ence (210○C), as indicated in the DSC curves Fig. 1. For this
n(t)diffused which diffuses in the surrounding polymer matrix at the reason, two different simulations will be presented, first for the
external pressure Pext. The gas concentration C(r,t) propagates PP 1 CBA-2 samples (2-1 and 2-2) with the decomposition of
concentrically in the surrounding polymer melt. The average citric acid, and then for the PP 1 CBA-3 samples (3-1 and 3-2)
concentration of gas at the bubble surface, C(r,t)/r5R, is related to with the sodium bicarbonate decomposition reaction.
the gas pressure inside the bubble Pgas through the solubility factor
K.
The following assumptions are made:

1. The bubble is spherically symmetric when it nucleates and


remains so for the entire period of growth.
2. The polymer matrix is Newtonian.
3. The growth process is considered isothermal. Latent heat of
reaction is neglected.
4. Inertia effects are neglected and the fluid is assumed to be
incompressible, which is reasonable in the pressure range
studied.
5. The gas released by the chemical reactions follows the ideal
gas law.
6. The matrix is considered as an infinite medium and one sin-
gle bubble is considered. This condition will be commented

FIG. 11. Single bubble growth model. [Color figure can be viewed in the
FIG. 12. Theoretical fitting of the number of moles created during the
online issue, which is available at wileyonlinelibrary.com.]
CBA decomposition: (a) CBA-2 and (b) CBA-3.

8 POLYMER ENGINEERING AND SCIENCE— DOI


TABLE 3. Fitting parameters of the number of moles of gas created during dnðtÞdiffused @C
the decomposition reaction of the 524pR2D (16)
r
j d @

CBA A1 (mol) A2 (mol) t0 (s) k (s) R2 D is the diffusion coefficient, and C is the gas concentration
of the diffused gas. The concentration profile is given by a dif-
2 4.17 3 1012 2.95 3 10214 63.12 8.59 0.9991 fusion equation around the bubble which writes:
3 3.92 3 10212 2.18 3 10214 57.87 9.20 0.9994 . Σ
dC @ C @ C DðT Þ @ @C
Figure 12 presents the theoretical fitting of the experimental 5 1uðr; tÞ 5 r 2
(17)

TGA data extrapolated to one single particle of reactive agent dt @t @r r2 @r @r


for both decomposition reactions, at a temperature of 210○C. u is the velocity field related to the bubble expansion, in spheri-
The low heating rate (about 6○C/min according to the tempera- cal coordinates. The ratio between the convection term and the
ture measurements during the foaming experiment), permits, as diffusion term is a Peclet number (Pc) given by (see Joshi et al.
a first approximation, the extrapolation of the non-isothermal [29]):
results derived from the TGA to isothermal kinetics at 210○C.
The total reaction time was between 100 and 200 s, much convection u=R
P c5 = (18)
lower than the foaming time employed during the expansion diffusion D=
experiment, which assures that the chemical reactions are com- R

plete. Different kinetic models can be found in the literature, The diffusion coefficient value can be taken from the litera-
such as the Kamal and Sourour model [34]. In our case, the ture [30], with a typical value, for PP and PE/CO 2 systems, of
best fitting correlation was found using Boltzmann’s exponen- D 5 1028 m2/s. The calculation without diffusion will show that
tial functions: the bubble radius goes from 5.10 26 to 2.1024 m and the growth
speed of the bubble radius is around 10 26 m/s, leading to a Pec-
A12A2 let number of 1024. Therefore, convection can be neglected.
ncreated .t 5AΣð21Þ t2t0 (13) The numerical implementation requires defining boundary
11exp k
and initial conditions. Three physical parameters must be
Fitting parameters, as well as the quality of the correlation defined at time t 5 0. First, the average size of the reactive
are presented in Table 3. agent particles can be considered as the initial bubble radius R0,
The differential equation for the bubble radius growth as a as a first approximation. This assumption is only an approxima-
function of time writes [30]: tion, and the sensitivity of the calculations when changing the
initial radius will be analyzed. The initial number of moles of
!
dR R nðtÞ<Tg 2c
dt
5 4g gas n0 can be directly calculated from the gas perfect law using the
2P 2 (14)
4 pR3 ext
initial radius R0 5 R¯ 5 4.46 mm, at the two different exter-
3 R nal pressures 0.25 MPa and 0.5 MPa. The boundary condition
R is the bubble radius, g is the polymer viscosity and c is the for the gas concentration at the bubble surface C(R, t) is
surface tension. As explained before, the gas temperature Tg is described by Henry’s law:
assumed to be constant (210○C) during the expansion process. 3
The variation of the number of moles of gas n(t) inside the CðR; tÞ5
nðtÞ<T g
k (19)

bubble is derived from Eq. 15: 4 pR3

dn dnðtÞcreated
5 2 (15)
dnðtÞdiffused

dt dt dt Numerical Implementation
The numerical implementation of the previous equations is
The number of moles of gas which diffuse outside the bubble
carried out by means of an incremental time marching approach
in the surrounding polymer matrix is obtained from the mass
during the decomposition reaction time (about 100 s, see Fig.
transfer of CBA at the gas-polymer interface and it can be
12). This approximation implies that the expansion process is
expressed as follows:
limited to the decomposition reaction time, and that no further

TABLE 4. Physical parameters and initial conditions employed for the foaming simulation.

Parameter CBA 2 CBA 3 Units

Initial number of moles n0 (3.56 or 7.13) 3 10 214


(3.56 or 7.13) 3 10 214
mol
Initial radius R0 4.46 3 1026 4.46 3 1026 m
Viscosity g 500 500 Pa·s
Surface tension ca 0.020 0.020 J·m
22

Diffusion coefficient Da 8 3 1029 8 3 10 29


m2·s21
Gas temperature Tg 483 483 K
DOI POLYMER ENGINEERING AND SCIENCE—2014
External pressure Pext (2.5 or 5) 3 105 (2.5 or 5) 3 105 Pa
Solubility parameter Ka 1.15 3 1024 1.15 3 1024 mol·m23 Pa2
a
Values taken from literature [30]

DOI POLYMER ENGINEERING AND SCIENCE—2014


FIG. 13. Gas diffusion profiles obtained from the diffusion equation (CBA- FIG. 15. Numerical prediction of the single bubble growth during the
2). expansion process. [Color figure can be viewed in the online issue, which is
available at wileyonlinelibrary.com.]
expansion occurs later. The variation of the number of moles of
gas created is obtained by differentiating Eq. 13: with bubble radius in a proportion C /R 23 (see Eq. 19). At
Two hundred and fifty iterations were used to solve the equa- high reaction times, the gas concentration at the interphase is
tions. We adjust the time step Dt in order to obtain an incremen- negligible.
tal radius variation less than 1%: Figure 14 compares the number of moles of gas created by
the decomposition reaction, diffused and retained in the bubble
during the expansion process for an external pressure of 0.5
RESULTS AND DISCUSSION MPa.
Table 4 presents the physical parameters and the initial con- These results show that, in the foaming conditions of Fig. 14,
ditions employed in the foaming simulation. the diffusion process becomes noticeable from 20 s. From that
Figure 13 presents the gas concentration profiles outside the time, the quantity of diffused gas increases, reaching a final
bubble for five different reaction times and CBA-2, for an exter- value about 1.35 3 10212 moles. On the other hand, the number
nal pressure of 0.25 MPa. Results for the concentration profiles of moles of gas created by the decomposition reaction reaches a
taking an external pressure of 0.5 MPa were almost equivalent, value around 4 3 10212 moles.
which indicates that the slight variation in the external pressure At the end of the reaction, the ratio between diffused and
does not affect significantly the gas diffusion. created gas is about 33%. This value is in reasonable agreement
The quantity of gas which diffuses outside the bubble with the experimental values found previously in Table 2
increases as the reaction develops. For reaction times close to (between 25 and 50% of diffused gas).
40 s, the gas diffusion penetration thickness is around 150 mm, Figure 15 presents the predicted radius evolution for both
whereas for reaction times of 2 s, this distance is only 50 mm. CBA-2 and CBA-3.
The gas concentration at the gas-polymer interphase decreases The final bubble radius depends strongly on the external
strongly with time, even if the quantity of gas generated inside pressure and is obviously limited by the gas diffusion. For the
the bubble n(t) increases according to the TGA kinetics. The CBA-2, when the external pressure is 0.25 MPa, the bubble
reason for this decrease is related to the dependence of C(R,t) radius reaches 530 mm when gas diffusion is neglected, and
about 130 mm when the diffusion process is considered. When
the external pressure is increased up to 0.5 MPa, these values
are 330 and 90 mm, respectively. The theoretical values obtained
(considering the diffusion process), are in qualitative agreement
with the foaming experiments values (120 and 85 mm, respec-
tively, see Table 2). A similar analysis can be performed for the
CBA-3. In this case, the theoretical values obtained considering
the diffusion process are about 300 and 180 mm, for both exter-
nal pressures of 0.25 and 0.5 MPa. These results are also in
qualitative agreement with the foaming experiments (350 and
190 mm, respectively, see Table 2).
Using the R(t) value allows to compute the velocity u at the
bubble/polymer interface. The evolution of u for the CBA-2 and
CBA-3 is similar, with values are around 10 26 m/s which justi- fies
neglecting the convection term in the diffusion Eq. 17.
The sensitivity of the model to the initial radius R0, and the
FIG. 14. Evolution of the number of moles of gas during the foaming diffusion coefficient D and the viscosity g to the bubble growth
expansion of a single bubble (CBA-2). [Color figure can be viewed in the
rate has been investigated (Fig. 16). The results are shown for
online issue, which is available at wileyonlinelibrary.com.]
CBA-3 for an external pressure of 0.25 MPa.

10 POLYMER ENGINEERING AND SCIENCE— DOI


The chosen diffusivity value was taken directly from the litera-
ture [30], and corresponds to the diffusivity value of CO 2 into a
PP matrix at 483 K. A more detailed analysis should include the
diffusion process of water vapor into a PP matrix. It is possible
to assume a value of the diffusion coefficient of water vapor
slightly higher than the value for CO 2 due to the lower molar
mass. However, the range of diffusion coefficients analyzed in
Fig. 16b is expected to cover both CO2 and water vapor diffu-
sion process.
It can be observed that increasing viscosity g from 200 Pa·s
to 1000 Pa·s induces a decrease of the final bubble radius from
600 mm to less than 100 mm (Fig. 16c). The viscosity was deter-
mined experimentally from shear rheological measurements and
presents a Newtonian plateau around 500 Pa.s at the strain rates
encountered during the foaming process. During the foaming
process, the flow around a single bubble is purely elongational,
which means that the chosen viscosity value determined by
shear measurements may be significantly underestimated. This
justifies testing higher viscosity values in the single bubble
growth model, as seen in Fig. 16c.
The influence of the initial number of moles n0 and the influ-
ence of the initial radius R0 are correlated from the gas perfect
law (n0 R0/3
).
Finally, as the number of bubbles is less important than the
number of CBA particles and as the final radius of the bubbles
is more important than the mean initial distance between the
particles, it is clear that bubble coalescence appears quite early
in the process. Due to the large important CBA particle size dis-
tribution (Fig. 5), one can believe that the biggest initial par-
ticles will induce the biggest bubbles in the first stage of their
development (see Fig. 16a), which will coalesce with the small
surrounding bubbles initiated by the smallest initial particles.
Considering, for example, one bubble with a diameter of 100
mm surrounded by eight bubbles with a diameter of 25 mm (see
Fig. 16a at t 5 5 s), the final diameter of the central bubble will
be only 108 mm after coalescence. As a consequence, if 88% of
the bubbles disappear during the foaming process, (which corre-
sponds to the scenario of one bubble surrounded by eight
smaller bubbles), the final bubble diameter will be only
enhanced by 8%. This may explain the good agreement between
the calculation which does not account for coalescence and the
experimental observation where coalescence takes place.

CONCLUSIONS
A simple polymer expansion experiment has been designed
FIG. 16. Influence of several parameters in the single bubble growth (Test 3- to test the foaming behavior of different CBA, submitted to dif-
1): (a) initial radius R0; (b) diffusion coefficient D; and (c) viscosity g. ferent pressure conditions. The bubble size and bubble size dis-
tribution depend on the blowing agent and on the applied
In Fig. 16a, three different values of the initial radius R0 pressure and careful bubble size measurements obtained by
were chosen (5 nm, 5 mm, and 100 mm). The lowest value corre- Image Analysis correlate well with the global macroscopic
sponds to the typical critical radius value that can be found in expansion of the foamed sample.
the literature for homogeneous nucleation [4]. The final pre- A single bubble expansion model assuming nucleation on
dicted bubble radius is only slightly influenced by the initial each CBA particle, accounting for the different chemical reac-
radius value (between 290 and 320 mm) but differences are tions and for the gas diffusion from the bubble to the surround-
obviously very important at intermediate time steps (till 30 s). ing polymer matrix agrees fairly well with the experimental
Varying the diffusion coefficient from 1028 m2/s to 10211 m2/s results, despite the strong hypothesis. This allows us to build a
leads to a final bubble radius around 100 and 600 mm, simple method to estimate the capability of a CBA to develop a
respectively (Fig. 16b). Typical diffusion values for a gas- foamed structure and the resulting mechanical properties. In this
polymer system are in the range between 1028 and 10210 m2/s. work, the expansion ratio varied between 1.4 at 0.5 MPa and 2

DOI POLYMER ENGINEERING AND SCIENCE—2014


at 0.25 MPa. The cell size obtained when using CBA based in 15. C.A. Villamizar, and C.F. Han. Polym. Eng. Sci., 18, 687
citric acid was much smaller than the obtained using sodium (1978).
bicarbonate CBA, due to the number of moles of gas released. 16. A.K. Bledzki, J. K€uhn, H. Kirschling and W. Pitscheneder.
Also, a simple calculation using the single bubble growth Cell. Polym., 27, 91 (2008).
model, with no dissolution or nucleation phenomena, represents
17. A.K. Bledzki and O. Faru. Macromol. Mater. Eng., 291, 16
well the experiments.
(2006).
Further studies analyzing injection molding experiments with
the same CBAs are in progress. 18. E. Bociaga and P. Palutkiewicz. Polym. Eng. Sci., 53, 780
(2013).
REFERENCES 19. A.K. Bledzki and O. Faruk J. of App. Polym. Sci., 97, 1090
(2005).
1. D. Li, T. Liu, L. Zhao, and W. Yuan. Ind. Eng. Chem. Res., 48,
7117 (2009). 20. Additives for Polymers., 2013, 1 (2013).
2. M. Hasan, Y.G. Li, G. Li, C.B. Park, and P. Chen. J. Chem. 21. R. Gosseling and D. Rodrigue. Polym. Test., 24, 1027 (2005).
Eng. Data., 55, 4885 (2010). 22. M. Amon and C.D. Denson. Polym. Eng. Sci., 24, 1026 (1984).
3. Z. Lei, H. Ohyabu, Y. Sato, H. Inomata, and R. Smith. J. 23. J. Bikard, J. Bruchon, T. Coupez and B. Vergnes. J. Mater.
Supercrit. Fluids, 40, 452 (2007). Sci., 40, 5875 (2005).
4. J.H. Han and C.D. Han. J. Polym. Sci. Part B, 28, 711 (1990). 24. J. Bikard, J. Bruchon, T. Coupez and L. Silva. Coll. and Surf.
5. W. Zhai, J. Yu, L. Wu, W. Ma, and J. He. Polymer, 47, 7580 A., 309, 49 (2007).
(2006). 25. R.J. Koopmans, C.F. Jaap, C. F. den Doelder and A.N. Paquet.
6. S. Areerat, E. Funami, Y. Hayata, D. Nakagawa, and M. Adv. Mater., 23, 49, (2000).
Ohshima, Polym. Eng. Sci., 44, 1915 (2004).
26. Y. Otsuki and T. Kanai. Polym. Eng. Sci., 45, 1277 (2005).
7. A. Arefmanesh, S.G. Advani, and E.E. Michaelides, Int. J. Heat
Mass Transf., 35, 1711 (2004). 27. M.A. Shafi and R.W. Flumerfelt. Chem. Eng. Sci., 52, 628
(1997).
8. D.C. Venerus. Polym. Eng. Sci., 41, 1390 (2001).
28. M.A. Shafi, J.G. Lee, and R.W. Flumerfelt. Polym. Eng. Sci.,
9. D.C. Venerus. Cell. Polym., 22, 9 (2003).
36, 1950 (1996).
10. D.F. Baldwin, C.B. Park, and N.P. Suh. Polym. Eng. Sci., 36,
1437 (1996). 29. K. Joshi, J.G. Lee, M.A. Shafi, and R.W. Flumerfelt J. Appl.
Polym. Sci., 67, 1353 (1998).
11. A.K. Nema, A.V. Deshmukh, K. Palanivelu, S.K. Sharma, and
T. Malik. J. Cell Plast., 44, 277 (2008). 30. K. Taki. Chem. Eng. Sci., 63, 3569 (2008).
12. S. Nakai, K. Taki, and J. Tsujimura. Polym. Eng. Sci., 48, 107 31. K. Taki, T. Yanagimoto, E. Funami, M. Okamoto and M.
(2008). Ohshima. Polym. Eng Sci., 44, 1004 (2004).
13. X. Qin, M.R. Thompson, and A.N. Hrymak. Polym. Eng. Sci., 32. J.J. Feng, and C.A. Bertelo. J. Rheol., 48, 439 (2004).
47, 522 (2007). 33. M. Emami, M.R. Thompson, and J. Vlachopoulos. Polym. Eng
14. D. Eaves. Handbook of Polymer Foams, Rapra Technology Sci., 54, 1201 (2014).
Limited, United Kingdom (2004). 34. S. Sourour and M.R. Kamal. Thermochim. Acta, 14, 41 (1976).

12 POLYMER ENGINEERING AND SCIENCE— DOI

You might also like