ArticlePDF Available

Density Modeling of Polyurethane Box Foam

Authors:

Abstract and Figures

A model based on about a dozen fundamental differential equations is used to evaluate and simulate the urethane reactions and physical processes of urethane box foaming. This work focuses on quantitative modeling of foam density for foams using water and physical blowing agents. The final densities of foams range from 30 to 90% of the density as projected with full utilization of the blowing agent. The primary sources of inefficient use of blowing agent are loss of the physical blowing during open-air mixing and degassing—basically, physical blowing agents with boiling points between 25 and 80°C will evaporate and experience cell rupture in box foams. This loss of blowing agent would not apply to in-line mixers used for commercial production and should be taken into account with scaling up box or cup foams commercial processes. POLYM. ENG. SCI., 54:1503–1511, 2014. © 2013 Society of Plastics Engineers
Content may be subject to copyright.
Density Modeling of Polyurethane Box Foam
Lu Shen,
1
Yusheng Zhao,
1
Ali Tekeei,
1
Fu-Hung Hsieh,
2
Galen J. Suppes
1
1
Department of Chemical Engineering, University of Missouri-Columbia, Columbia, Missouri 65211
2
Department of Biological Engineering, University of Missouri-Columbia, Columbia, Missouri 65211
A model based on about a dozen fundamental differen-
tial equations is used to evaluate and simulate the ure-
thane reactions and physical processes of urethane
box foaming. This work focuses on quantitative model-
ing of foam density for foams using water and physical
blowing agents. The final densities of foams range
from 30 to 90% of the density as projected with full uti-
lization of the blowing agent. The primary sources of
inefficient use of blowing agent are loss of the physical
blowing during open-air mixing and degassing—basi-
cally, physical blowing agents with boiling points
between 25 and 80C will evaporate and experience
cell rupture in box foams. This loss of blowing agent
would not apply to in-line mixers used for commercial
production and should be taken into account with
scaling up box or cup foams commercial processes.
POLYM. ENG. SCI., 54:1503–1511, 2014. V
C2013 Society of
Plastics Engineers
INTRODUCTION
Polyurethanes (PU) play a significant role in global
industry with over three-quarters of the consumption in
the form of foams [1]. They are extensively used in the
field such as rigid insulations in the walls of refrigerators
and buildings, high-resilience and flexible foam cushions,
and elastomeric wheels and tires.
Foams can expand from during polymerization if gases
or vapors are produced in concert with the polymeriza-
tion. A common mechanism for gas generation is the
chemical reaction between water and isocyanate forming
carbon dioxide. In addition, exothermic polymerization
reactions can provide heat to evaporate volatile blowing
agents such as methyl formate (MF) and pentane. For
rigid foams, the physical properties of most interest are
density, compressive strength, and thermal conductivity.
Both cell size and cell number contribute to change of
density and compressive strength [2]. For PU foam, to a
first approximation, there exists a linear relationship
between density and compressive strength [3]. Density
impacts the thermal conductivity due to the lower thermal
conductivity of the gas phase relative to the resin phase
[4]. Therefore, accurately predicting density during poly-
urethane foaming is both important in its own right and
as a critical step to ultimately predict other physical prop-
erties of foams [5–7].
Baser and Khakhar [8, 9] carried out a detailed experi-
mental study of chemical and physical blowing agents
with measurement of both temperature and density
changes during the foaming process. The theoretical
model was based on the hypothesis that the foam was
under a single pseudohomogeneous phase, and the varia-
tion of density with time was obtained by applying
energy and mass balances along with the kinetic and ther-
modynamic relations. Later, Tesser et al. [10] developed
a mathematical model to simulate the temperature and
density of the foams with the evaluation of kinetics of the
reaction between polyol and isocyanate during the foam-
ing process with an emphasis on Flory–Huggins modeling
of blowing agent activity.
Gibson and Ashby [11] came up with an elastic modu-
lus based on the relative density for closed-cell (cc) foam
taking into account of three components: the strut flexion,
the stretching of the cell walls, and the gas pressure
inside the cc. Saint Michel et al. [12] characterized the
influence of the density and compared the experimental
results with Gibson and Ashby [11] and Christensen et al.
[13] in the linear domain and then extended to the nonlin-
ear domain.
Avalle et al. [14] tested several types of foams and
performed a study based on the Gibson model and the
Rusch (of the phenomenological type) model [15–17]
together with two other models namely a modified Gib-
son model and a new phenomenological model to charac-
terize the mechanical properties. Additionally, available
experimental data were used to obtain the relationship
between material density and model parameters.
These previous works illustrate the importance of
quantifying foam density in foam formulation. The work
of this article differs in two ways from this other work:
(a) the current work is based on over a dozen fundamen-
tal ordinary differential models for the reactions and
Correspondence to: Lu Shen; e-mail: lsc38@mail.missouri.edu
Contract grant sponsor: United Soybean Board.
DOI 10.1002/pen.23694
Published online in Wiley Online Library (wileyonlinelibrary.com).
V
C2013 Society of Plastics Engineers
POLYMER ENGINEERING AND SCIENCE—2014
physical processes and (b) this work evaluates sources of
blowing agent loss in addition to solubility of blowing
agents in the resin phase.
This work builds on the work of Zhao et al. [18] on
modeling the foaming process using a series of ordinary
differential equations with Arrhenius constants and
enthalpy parameters. Although the emphasis of Zhao’s
work was the prediction of polyol mixture performance in
formulations, the emphasis of this work is the modeling
and simulation of foam height (density).
EXPERIMENTAL DESIGN
Blowing Agents
As a chemical blowing agent, water reacts with isocya-
nate (RNCO) and produces carbamic acid (RNHCOOH)
which further decomposes into carbon dioxide and thus
generates gas bubbles in the resin according to reaction
Schemes 1 and 2 forming an amine (RNH
2
).
RNCO 1H2O!RNHCOOH (1)
RNHCOOH !RNH21CO21HEAT (2)
MF serves as a physical blowing agent that does not
rely on any chemical reactions. Table 1 summarizes phys-
ical properties of MF. The MF is initially soluble in the
mixture of isocyanate and polyol at ambient temperature.
Equation 3 is used to estimate the vapor pressure of
MF as a function of time [19].
ln Psat
i
PC

512xðÞ
21Ai
ðÞx1Bi
ðÞx1:51Ci
ðÞx31Di
ðÞx6

(3)
x51—T/T
c
where T
c
is critical temperature in Kelvins,
Psat
iis vapor pressure, in bars, P
c
is critical pressure, in
bars, Eqs.3 and 4 are valid at the temperature range of
220–487.2 K.
The vapor pressure is used in the modified Raoult’s
Law (Eq.4) equation which is used in combination with
heat and energy balances to estimate the amount of MF
that evaporates from the resin phase and proceeds to form
bubbles/cells. It should be noted that because the resin
phase is comprised of macromolecules, mass fraction was
used instead of mole fraction.
xicPsat
i5yiP(4)
It is assumed that the carbon dioxide and vapor-phased
MF are ideal with fugacity equaling to 1. Deviation from
ideality is accommodated by modifying the activity coef-
ficient. Mass balance equations are solved under the con-
straint of Raoult’s law to determine the extent to which
MF evaporates. During analysis of the data, performance
at an activity coefficient of 1.0 is compared to 15 to eval-
uate the extent that actual foam density varies from the
ideal model.
Materials
The isocyanate and petroleum-based polyols used in
this study were RUBINATE M isocyanate, Poly G76–
635, Voranol 360, and Jeffol R315x from Huntsman
Company and Dow Chemical Co. and their specifications
are shown in Tables 2. N,N-dimethylcyclohexylamine and
N,N,N0,N000 ,N00-pentamethyldiethylenetriamine were
used as gelling catalyst and blowing catalyst, respectively.
Momentive L6900 was used as the surfactant for rigid
foams, Tris (1-chloro-2-propyl) Phosphate (TCPP) was
used as fire retardant, and distilled water was use as
chemical blowing agent.
Experimental Procedure
Gelling and Foaming Experiment. Experiments were
performed to create polyurethane gels as well as rigid
polyurethane foams. The gel reactions were used as a
control to assist in interpreting the data for the foam reac-
tions. Table 3 provides the control formulation used in
these studies. The amount of water and MF in this formu-
lation was used as parameters to better understand the
foaming process.
The following steps were used in both the gelling and
foam experiments.
1. Polyols, water, blowing catalyst, gelling catalysts, and sur-
factant (B-side components) were added into a plastic cup
successively.
2. These B-side components were mixed for 10–15 s.
3. The mixture was allowed to degas for 2 min.
TABLE 1. Specifications of methyl formate (MF).
MW 60
Physical state Liquid
Boiling point (C) 31.5
P
c
(bar) 60.075
T
c
(K) 487.2
A
i
26.99601
B
i
0.89328
C
i
22.52294
D
i
23.16636
TABLE 2. Specifications of RUBINATE M isocyanate and three dif-
ferent polyols.
Product: RUBINATE M
Fn 2.70
Sp. gravity @25C 1.23
% NCO 31.2
Eq. wt. 135
Viscosity cps @25C 190
Product OH number
Poly G76–635 635
Voranol 360 360
Jeffol R315x 315
1504 POLYMER ENGINEERING AND SCIENCE—2014 DOI 10.1002/pen
4. Thereafter, preweighed isocyanate (A-side material) was
added and mixed at the same speed for 7–10 s.
5. The reacting mixture was then quickly poured into a box
with aluminum foil lining, and the foam was allowed to
rise and sit at ambient conditions (25C) during curing.
All the B-side chemicals were added in the foam reac-
tion, whereas water and blowing catalyst were not added
in the gel reaction.
A high-speed mixer blade attached to a floor-model
drill press was used to mix the chemicals. LabVIEW soft-
ware was used to monitor the temperature of the gel or
foam reactions for the first 15 min with a type-k thermo-
couple attached through a National Instruments SCB-68
box to a National Instruments PCI 6024E data acquisition
card.
Mass Loss Test. Experiments were also performed to
evaluate the amount of mass loss during box foaming. A
cardboard box containing the reacting foam mixture was
placed on a digital balance accurate to 0.01 g. The fol-
lowing sequence describes the manner in which mass was
balanced and measured:
1. Polyols, water, blowing catalyst, gelling catalysts, and sur-
factant (B-side components) were added and stirred in a
plastic cup. The B-side component was weighed and set
aside in a second plastic cup. About 2 min were allowed
for degassing.
2. The B-side components were mixed and stirred with A-
side for 15 s.
3. The mixtures were quickly poured into the cardboard box
(designated at t50 s) located on the balance. The foam
was allowed to rise and sit at ambient conditions during
curing.
4. Residual masses in the mixing cup and stirrer were meas-
ured (by mass differences) and recorded for subsequent
mass balance calculations.
RESULTS AND DISCUSSION
Foaming Height Studies
Figure 1 demonstrates the effect of water content on
the height of PU foams. Increasing water content leads to
increasing height of the box foam and decreasing density.
A doubling of the water content resulted in an approxi-
mate doubling of the height.
As a physical blowing agent, MF does not react with
isocyanate; it provides an additional degree of freedom to
control foam volume. Because of a lower boiling point
and volatility, MF volatilizes and dissolves in polymer
phase and consequently bubbles inside the foams come
into being. Figure 2 illustrates how more MF produces
more vapor which contributes to the volume increase.
Based on inspection, water appears to be more effective
in forming cells; the subsequent discussion pursues a
quantification and insight into the differences.
Increased amounts of both water and MF provide
lower foam densities. These trends are qualitatively con-
sistent with the mechanisms of Eqs.2 and 4. Figure 3
compares performances to what the ordinary deferential
TABLE 3. Foaming formulation of rigid polyurethane foam.
Weight/g Equivalents
B-side materials
Poly G76–635 13.84 0.1569
Voranol 360 15.68 0.1008
Jeffol R315x 4 0.0225
Dimethylcyclohexylamine
(gelling catalyst)
0.12
Pentamethyldiethylenetriamine
(blowing catalyst)
0.32 (foam reaction only)
Momentive L6900 0.6
TCPP 2
Distilled water (blowing agent) 1.04 (foam reaction only) 0.08
A-side material
RUBINATE M 61.548 0.4559
FIG. 1. Foaming height with different water content. Symbols “w”, “”, and “D” represent two repeated
samples containing 2.4 g MF with 1.0, 0.75, and 0.5 g water, respectively.
DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—2014 1505
equation model predicts for extreme cases of the activity
coefficient equal to 1 and 15. Even in the extreme case of
high activity coefficients, the actual foam height is much
less than what is projected by the model. Temperatures in
the foam exceed 100C during curing, and as such, little
MF remains in the liquid even if high activity coefficients
impacted performance. The model trends show that the
activity coefficient predominantly has an impact during
foam rise when the temperature ranged from MF’s boiling
point to about 40C greater than MF’s boiling point.
Actual foam height only attained about 63% of what
was projected with full use of the blowing agents and
ideal gas law calculation of the gas volume. Possible fac-
tors leading to the inconsistency are low-efficiency water-
isocyanate reaction, the rupture of carbon dioxide and MF
gas cells, and higher internal pressure. Further studies
were conducted that evaluated the mass of the foam dur-
ing the foaming process.
Mass Loss Studies
The box used to contain the rigid foam during the
foaming process was placed on a digital balance accurate
to 0.01 g to monitor weight loss during the foaming reac-
tion. Possible sources for weight loss are (1) air’s buoyant
force applied by the increasing volume of the foam, (2)
rupture of foam cells (bubbles) with collapse of the cell
or replacement of the vapors by new vapors, and (3)
evaporation/escape of MF, water, or carbon dioxide along
the upper surface of the foam.
Figure 4 illustrates the mass loss for a sample that
used only MF blowing agent and distinguishes between
FIG. 2. Foaming height with different MF content. Symbols “w”, “”, and “D” represent two repeated
samples contain 1.04 g water and 2.4, 1.5, and 0.5 g MF, respectively.
FIG. 3. Experimental height versus modeling height. Symbols “” and “w” represent two repeated samples
with 0.75 g water and 2.4 g MF. The smooth and dash lines represent MATLAB Simulation with 100%
water and MF when cequals to 15 and 1, respectively.
1506 POLYMER ENGINEERING AND SCIENCE—2014 DOI 10.1002/pen
buoyant force and true weight loss. The volume caused
by buoyancy was calculated using Eq.5.
Vb5DVqa5Vo2Vexp

qa(5)
V
b
is volume of the buoyancy, q
a
is density of the air,
V
o
is initial volume before the growth of the foam, which
corresponds to the total volume of isocyanate, polyols,
water, and MF, V
exp
is experimental volume of the foam.
For the foam of Fig. 4, V
o
592.5 cm
3
and V
exp
5572.7
cm
3
resulting in an equivalency of 0.57 g of buoyancy. It
is presented from Fig. 4 that the total amount of loss is
2.13 g, which indicates 1.38 g of mass loss attributed to
the inefficiency of blowing agent (much of which is
attributed to the box-foaming technique rather than the
blowing agent).
Table 4 summarizes the buoyancy forces and mass
loss of several foams. Dmand Dhare total mass loss
from experiment and height decrease associated with the
release of the blowing agents. The foam in Fig. 4 stopped
expanding at 75 s because an increase of viscosity facili-
tated the rigidity of the foam that did not allow further
expansion. During the first 75 s, the mass loss was
resulted from the buoyancy applied by the increasing vol-
ume of displaced air and the rupture of blowing agent
characterized as Dm
2
and Dm
3
in Fig. 4. Afterwards, no
expansion occurred in the system, and all the mass loss
was attributed to the escape of gases in the cells.
Figures 5 and 6 compare actual to model heights
where the model heights are for no loss of blowing agent
versus losses as summarized by Table 4. The results sub-
stantiate that a primary mechanism for the inefficiency of
blowing agent is the loss of blowing agent through the
upper surface of the foam through cell rupture (convec-
tion) or through the surface evaporation. Loss from the
top surface could be through diffusion straight from the
resin to the air or it could be from cells/bubbles that rup-
ture through the top surface.
Modeling Loss of Blowing Agent Efficacy
Possible sources of inefficacy of blowing agents
include: (1) higher than atmospheric pressure in cells, (2)
solubility of blowing agent in the polymer phase, and (3)
release of blowing agent through the upper surface. The
impact of these is summarized by Eq.5in terms of gas
volume.
FIG. 4. Mass change during box foaming process. Symbols “w” and “” are corresponding to the mass
value prior to and during the foaming process. 1,2,3, and 4represent the stage of degassing, stirring, foam
rising, and after complete rise. Dm
1
is mass loss prior to expansion of foam cells, Dm
2
is buoyancy weight
adjustment, Dm
3
is mass loss during foam growth, and Dm
4
is mass loss after the maximum foam height is
attained.
TABLE 4. Results of mass loss of different samples.
Foam type
Dm
(g equivalent)
Dh
(cm)
Dmfrom
buoyancy
Dmfrom
blowing
agent
Efficiency
of blowing
agent (%)
Efficiency
calculated by
height (%)
Foam only with MF 1 2.13 8.83 0.57 1.38 42.4 40.0
Foam only with MF 2 2.12 8.48 0.52 1.42 41.0 33.2
Foam only with water 1 2.22 12.47 1.72 0.50 80.2 80.1
Foam only with water 2 2.26 11.61 1.78 0.48 81.1 82.7
Gel 1 0.38 2.02 0.10 0.28
Gel 2 0.29 1.54 0.11 0.18
DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—2014 1507
V5Vi2Vp2VS2Rb(6)
V
i
is ideal volume of the foam without the gas rupture,
V
p
is volume loss due to the pressure, V
S
is volume loss
due to the solubility of the MF, R
b
is release of the blow-
ing agent.
This approach does not distinguish between the sus-
ceptibility of MF versus carbon dioxide to these mecha-
nisms. Under the assumption that the pressure buildup
has a minor contribution, V
p
is cancelled out in Eq.6.
Closed cell content is a property commonly measured
in rigid foam. A reasonable hypothesis is: if the final
foam has a high cc content, the foam would have main-
tained a high cc content during the entire foaming pro-
cess; and ccs correlate this low release of blowing agent.
The Eq.7is an example correlation that can be tested
with data where cc is close cell content (%).
Rb5k212ccðÞ (7)
Two assumptions to simplify this model are assuming
ideal gas behavior and assuming that only MF partitions
in the resin phase. Substituting Eq.6to8, where n
1
and
n
2
are moles of MF and carbon dioxide, and S
1
is the
fraction of MF that is in the resin (soluble).
V5n11n2
ðÞRT
P
2n1RT
PS12k212ccðÞ (8)
By setting V
o
equals to the first two terms of the three
on the right-hand side of Eq.8and expressing volume in
term of density, Eq.8can be represented as a linear equa-
tion where the extent to which avaries from k
2
provides
a point of discussion. Equation 9 displays a relationship
between density and close cell content used to interpret
the data.
FIG. 5. Samples with 4.0 g MF. Symbols “” and “D” represent two repeated samples with 4.0-g MF
blowing agent. Smooth and dashed lines symbolize MATLAB Simulation with 30 and 100% efficiency of
MF.
FIG. 6. Samples with 1.0 g water. Symbols “” and “D” represent two repeated samples with 1.0 g water.
Smooth and dashed lines symbolize MATLAB Simulation with 80 and 100% efficiency of water.
1508 POLYMER ENGINEERING AND SCIENCE—2014 DOI 10.1002/pen
1
qo21
q
5k2cc1a(9)
Figure 7 summarizes this trend for density versus close
cell content data collected as summarized by Table 5. The
figure illustrates that while k
2
is close to a, the best fit is
not with them equal. Furthermore, the data illustrate that
the primary cause for them not being equal is the concen-
tration of data at a nonzero limit as cc approaches 1.
For foam cells to expand, they must have a pressure
greater than the surroundings. For foams with nearly
100% cc content, the strength of the resin to preserve cell
integrity is at its greatest relative to foam with an open
cell nature, weak cell walls, and respective low cc con-
tent. The conclusion is that the internal pressure is signifi-
cant, is a function of cc, and is likely near a value of
2.02 atm at 100% cc content.
The lower dashed line of Fig. 7 represents a linear rela-
tion for internal cell pressure. Within the standard deviation
of the data, the linear model (solid line) as represented by a
combination of the two dashed lines is substantiated.
Future work on this area would include several topics
of interest, including: (a) development of fundamentally
based (not simple linear equations) for the relations of
cell pressure and release to cc., (b) acquiring more accu-
rate data, and (c) developing fundamentally based models
that can predict cc based on the properties of the reagents
used to make the foam and the foaming formulation.
Figure 8 displays the comparison between modeling and
experimental density, which presents slight deviation
FIG. 7. Linear correlation for Eq.9. Symbols “” represent experi-
mental data between 1/q
o
21/qand close cell content. Smooth line rep-
resents the modeling line characterizing 1/q
o
21/q520.0846*
cc 10.0967. “--” line is displayed as trend of internal bubble pressure in
which pressure reaches highest at 2 atm when close cell content c is
close to 100%. “-.-” line stands for k
2
52a.
TABLE 5. Experimental and modeling density from samples displayed in Fig. 7 with comparison of efficiency, close cell content and quantities of
blowing agent.
Close cell content (%) Efficiency (%) Density exp (g/ml) Density model (g/ml) Quantities of MF (g) Quantities of Water (g)
87.55 54.29 33.00 21.43 2.40 1.04
90.30 53.64 33.90 21.76 2.41 1.02
91.66 69.16 33.36 26.55 2.42 1.06
94.06 69.12 28.50 24.75 2.45 1.01
94.63 58.05 42.22 26.76 2.39 0.76
91.96 57.27 52.50 34.24 2.40 0.50
95.46 50.80 57.20 33.12 2.48 0.50
94.40 66.21 44.85 35.17 2.47 0.51
94.31 71.16 33.31 26.83 1.51 1.01
98.20 58.72 42.09 28.81 1.53 1.05
99.66 67.48 39.80 27.31 1.50 1.04
93.99 62.89 44.79 29.96 0.52 1.01
89.10 56.32 43.63 27.01 0.50 1.07
98.80 63.79 40.03 27.12 0.50 1.04
94.84 62.00 48.60 30.14 0.00 1.05
97.76 71.20 39.59 30.58 0.00 1.04
83.22 54.48 27.05 18.32 2.35 1.53
73.16 50.21 24.59 16.26 2.29 1.97
93.39 59.68 32.32 21.83 2.20 1.03
71.80 50.67 31.5 17.19 2.40 1.50
56.83 44.32 25.72 12.63 2.30 2.00
94.96 55.33 34.76 18.60 2.32 1.02
90.27 60.88 41.32 26.67 2.33 1.03
74.62 50.00 36.87 28.42 2.25 1.00
78.70 50.00 32.76 27.01 2.20 0.98
91.36 55.33 36.32 21.14 2.42 1.05
97.63 63.79 36.10 23.59 2.40 1.06
97.67 67.26 32.91 24.58 2.42 1.05
98.53 64.16 33.37 22.39 2.41 1.05
98.53 64.37 30.99 21.10 2.41 1.06
96.78 68.61 33.36 24.92 2.41 1.05
DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—2014 1509
caused by higher internal bubble pressure and potential
experimental error. It is predictable to estimate the density
with the aid of MATLAB Simulation when we have the
foaming formulation.
CONCLUSIONS
The measuring of mass and height in combination with
computer-based simulation of the foaming process pro-
vided a closure of the mass balance of blowing agents
used to make rigid box foams. The results indicated that
evaporation losses during the mixing and degassing steps
leads to a loss of efficacy of the blowing agent. This con-
clusion would broadly apply to blowing agents that rely
on low boiling points which form gas cells in urethane
foams at temperatures between 30 and 80C.
The primary sources of inefficient use of blowing
agent are loss of the physical blowing during open-air
mixing and degassing. This loss of blowing agent would
not apply to in-line mixers used for commercial produc-
tion and should be taken into account with scaling up box
or cup foams commercial processes.
Although previous modeling work that was able to cor-
relate inefficiency with Flory–Huggins parameters
appeared to exaggerate the impact of activity coefficients,
the activity coefficients considered in this were reason-
ably close to 1.0. A correlation relates to lack of efficacy
of blowing agents to both closed cell (cc) content and the
buildup of pressure in the cells. The implication is that if
cc content is able to be simulated from a foam recipe,
increasingly accurate estimates/simulations of density can
also be attained. This was a major finding of this study as
it identifies that the modeling of cell formation and rup-
ture is critical to accurately model foam density.
ABBREVIATIONS
PU polyurethanes
cc closed-cell
MF methyl formate
ODE ordinary deferential equation
MW molecular weight
Fn functionality
Eq. Wt equivalent weight
csolubility
R ideal gas constant (L atm/K mol)
T temperature (K)
P pressure (atm)
t time (s)
V volume (ml)
qdensity (g/ml)
m mass weight (g)
h height (cm)
T
c
critical temperature (K)
P
c
critical pressure (bar)
P
sat
vapor pressure (bars)
V
b
volume of the buoyancy
q
a
density of the air
V
o
initial volume before the growth of the foam
V
exp
experimental volume of the foam
V
i
ideal volume of the foam without the gas rupture
V
p
volume loss due to the pressure
V
S
volume loss due to the solubility of the MF
R
b
release of the blowing agent
ACKNOWLEDGMENT
The authors thank the United Soybean Board for finan-
cial support of the experimental studies used to validate the
modeling work. Thanks to FSI Company providing foam
formulas and technology support.
REFERENCES
1. G. Avar, Polyurethanes (PU), Vol. 10, Kunststoffe Interna-
tional, 123 (2008).
2. H. Fan, A. Tekeei, G.J. Suppes, and F.-H. Hsieh, Physical
Properties of Soy-Polyol Based Polyurethane Foams
FIG. 8. Linear relationship between ideal and experimental density. Symbols “” represent experiment
density versus modeling density. Smooth line represents the modeling line.
1510 POLYMER ENGINEERING AND SCIENCE—2014 DOI 10.1002/pen
Reinforced with Microspheres and Nanoclay, American
Society of Agricultural and Biological Engineers Annual
International Meeting, 5, 3520 (2011).
3. H. Fan, A. Tekeei, G.J. Suppes, and F-H. Hsieh, J. Appl.
Polym. Sci.,127(3), 1623 (2013).
4. V. Yakushin, L. Bel’kova, and I. Sevastyanova, Properties
of Rigid Polyurethane Foams Filled with Glass Micro-
spheres, Mech. Compos. Mater.,48(5), 579 (2012).
5. G.S. Tay, L.N. Ong, and H.D. Rozman, J. Appl. Polym.
Sci.,125(1) 58 (2012).
6. M. Thirumal, D. Khastgir, N.K. Singha, B.S. Manjunath,
and Y.P. Naik, Cell. Polym.,28(2) 145 (2009).
7. L.T. Yang, C.S. Zhao, C.L. Dai, L.Y. Fu, and S.Q. Lin, J.
Polym. Environ.,20(1) 230 (2012).
8. S.A. Baser and D.V. Khakhar, Polym. Eng. Sci.,34(8) 642
(1994).
9. S.A. Baser and D.V. Khakhar, Polym. Eng. Sci.,34(8) 632
(1994).
10. R. Tesser, M. DiSerio, A. Sclafani, and E. Santacesaria, J.
Appl. Polym. Sci.,92(3) 1875 (2004).
11. L.J. Gibson and M.F. Ashby, Cellular Solids: Structure and
Properties, Cambridge University Press (1999).
12. F. Saint-Michel, L. Chazeau, J.Y. Cavaill
e, and E. Chabert,
Compos. Sci. Technol.,66(15) 2700 (2006).
13. R.M. Christensen, J. Mechan. Phys. Solids,34(6), 563
(1986).
14. M. Avalle, G. Belingardi, and A. Ibba, Int. J. Impact Eng.,
34(1) 3 (2007).
15. K.C. Rusch, J. Appl. Polym. Sci.,13(11) 2297 (1969).
16. K.C. Rusch, J. Appl. Polym. Sci.,14(5) 1263 (1970).
17. K.C. Rusch, J. Appl. Polym. Sci.,14(6) 1433 (1970).
18. Y. Zhao, M.J. Gordon, A. Tekeei, F.-H. Hsieh, and G.J.
Suppes, J. Appl. Polym. Sci. (in press).
19. R.C. Reid, J.M. Prausnitz, and B.E. Poling, The Properties
of Gases and Liquids, McGraw-Hill (1987).
DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—2014 1511
... In practical applications, the effects of different components, such as blowing agent and isocyanate index, on mechanical properties and functions of different structural foams have been extensively investigated [17,18]. Based on specific research objectives, many studies typically focused on the temporal and spatial changes in temperature, density, bubble scale, or foaming profile to assess the influencing factors [19][20][21][22][23]. Tu et al. [24] examined the influence of isocyanate index on density, compressive strength, and thermal conductivity of polyurethane foam, but the presence of physical blowing agent was not considered in their work. ...
... Recently, numerous numerical simulations have been conducted to investigate the foaming process of plastic foam. [19][20][21][22][23][24] As the time scales of foaming and polyurethane (PU) crosslinking were very short, differential scanning calorimetry and parallel plates rheometer were not suitable for measuring the material properties. Raimbault et al. 25 set up the FOAMAT device to record the PU foaming height, foaming temperature, and stress as a function of time, and then the analytical models of curing, foaming kinetics, and resin viscosity were identified. ...
Article
The evolution and control of the temperature field during the chemical foaming of epoxy resin are of paramount importance because the foam structure results from the competition of resin crosslinking and foaming, both of which are highly dependent on temperature distributions. Herein, the epoxy foams, consisting of diglycidyl ether of bisphenol A, glycidyl amine‐type epoxy resin, and 4,4′‐diamino diphenyl sulfone hardener, are prepared using azodicarbonamide and 4,4′‐oxydibenzenesulfonyl hydrazide as a chemical foaming agent (CFA). Kinetic models for heat release and CFA decomposition are established using the auto‐catalytic and n ‐th order models, respectively. By integrating the transient flow of resin during the foaming process, numerical simulations of the temperature field evolution within the self‐expanding geometry are conducted to investigate the effects of foaming temperature, heat transfer coefficient, and mold diameter on spatial temperature distributions. A comparison of the kinetic parameters of epoxy curing and CFA decomposition at various foaming temperatures (433, 443, and 453 K) reveals that the acceleration of the curing rate is consistent with that of the decomposition rate as the foaming temperature increases, then the foam structure remains largely unchanged across different foaming temperatures. However, local overheating is unavoidable for the foams at 443 and 453 K. This study offers a method for optimizing the processing parameters in preparing epoxy foams. Highlights Auto‐catalytic model for predicting heat released rate is well established. The decomposition kinetics of the chemical foaming agent is described by n ‐th model. Nonisothermal simulation is implemented to study the self‐expanded process of epoxy foam. The foam structure is predicted by comparing the kinetic between the curing and expansion processes.
... However, high-temperature baking can only cause partial loss of PBAs, and the decomposition loss of polyurethane foam at high temperatures will also bring significant errors to the experiment. Shen et al. [21] quantitatively modeled the foam density of the physical process of polyurethane box foaming. Gandhi et al. [22] developed a model forecast the density distribution of polyurethane foams blown by water as a chemical blowing agent. ...
Article
Full-text available
Polyurethane rigid foam is a widely used insulation material, and the behavior characteristics and heat absorption performance of the blowing agent used in the foaming process are key factors that affect the molding performance of this material. In this work, the behavior characteristics and heat absorption of the polyurethane physical blowing agent in the foaming process were studied; this is something which has not been comprehensively studied before. This study investigated the behavior characteristics of polyurethane physical blowing agents in the same formulation system, including the efficiency, dissolution, and loss rates of the physical blowing agents during the polyurethane foaming process. The research findings indicate that both the physical blowing agent mass efficiency rate and mass dissolution rate are influenced by the vaporization and condensation process of physical blowing agent. For the same type of physical blowing agent, the amount of heat absorbed per unit mass decreases gradually as the quantity of physical blowing agent increases. The relationship between the two shows a pattern of initial rapid decrease followed by a slower decrease. Under the same physical blowing agent content, the higher the heat absorbed per unit mass of physical blowing agent, the lower the internal temperature of the foam when the foam stops expanding. The heat absorbed per unit mass of the physical blowing agents is a key factor affecting the internal temperature of the foam when it stops expanding. From the perspective of heat control of the polyurethane reaction system, the effects of physical blowing agents on the foam quality were ranked in order from good to poor as follows: HFC-245fa, HFC-365mfc, HFCO-1233zd(E), HFO-1336mzzZ, and HCFC-141b.
... However, they did not investigate the low efficiency of the blowing agent. Shen et al. [23] studied the physical process of polyurethane box foaming and quantitatively modeled the foam density. They found that in their study, the final density of the foam was only 30-90% of the predicted density. ...
Article
Full-text available
This work developed a novel method for measuring the effective rate of a PBA (physical blowing agent) and solved the problem that the effective rate of a PBA could not be directly measured or calculated in previous studies. The results show that the effectiveness of different PBAs under the same experimental conditions varied widely, from approximately 50% to almost 90%. In this study, the overall average effective rates of the PBAs HFC-245fa, HFO-1336mzzZ, HFC-365mfc, HFCO-1233zd(E), and HCFC-141b are in descending order. In all experimental groups, the relationship between the effective rate of the PBA, rePBA, and the initial mass ratio of the PBA to other blending materials in the polyurethane rigid foam, w, demonstrated a trend of first decreasing and then gradually stabilizing or slightly increasing. This trend is caused by the interaction of PBA molecules among themselves and with other component molecules in the foamed material and the temperature of the foaming system. In general, the influence of system temperature dominated when w was less than 9.05 wt%, and the interaction of PBA molecules among themselves and with other component molecules in the foamed material dominated when w was greater than 9.05 wt%. The effective rate of the PBA is also related to the states of gasification and condensation when they reach equilibrium. The properties of the PBA itself determine the overall efficiency, while the balance between the gasification and condensation processes of the PBA further leads to a regular change in efficiency with respect to w around the overall average level.
... Formulation development was first aided by urethane-forming reaction modeling by obtaining temperature, concentration [26], and foam height [27] profiles as functions of time. Most urethane formulations require the adjustment of more than a dozen variables (including reagents, catalysts, initial temperature, blowing agents, and additives), so the design process begins with the construction of a base case simulation program that is subsequently (and continues to be) iteratively enhanced to increase accuracy and versatility. ...
Article
Full-text available
A MATLAB program was developed to simulate urethane-forming reactions by solving over a dozen differential equations, energy balance, mass balance, and constitutive equations simultaneously. The simulation program was developed for half a decade to simulate the basic kinetics of polyurethane reactions and more complex phenomena that cannot be obtained in laboratories. In the current investigation, the simulation is applied to determine the limits of the performance of polyurethane foam formation. n-pentane, cyclohexane, and methyl formate were used as physical blowing agents, and water was used as a chemical blowing agent. The simulation code increases the accuracy of the results and makes the foam performance process less time- and money-consuming. Specifically, the MATLAB code was developed to study the impact of physical and chemical blowing agents at different loadings on the performance of rigid polyurethane foams. Experimental data were used to validate the simulation results, including temperature profiles, height profiles, and the tack-free time of urethane foam reactions. The simulation results provide a window for the proper type and the optimum amount range of different physical and chemical blowing agents.
... Furthermore, Knowledge of the mass of the reactants is also important for ensuring reproducible experiments. Shen et al., [8] modelled PUF box foam density using height and mass loss data, for a low boiling point blowing agent and water. The mass loss during mixing and degassing explained the inefficiencies in the low boiling point blowing agent. ...
Article
Full-text available
Adiabatic temperature rise is an important method for determining isocyanate conversion in polyurethane foam reactions as well as many other exothermic chemical reactions. Adiabatic temperature rise can be used in conjunction with change in height and mass measurements to gain understanding into the blowing and gelling reactions that occur during polyurethane foaming as well as give important information on cell morphology. FoamPi is an open-source Raspberry Pi device for monitoring polyurethane foaming reactions. The device effectively monitors temperature rise, change in foam height as well as changes in the mass during the reaction. Three Python scripts are also presented. The first logs raw data during the reaction. The second corrects temperature data such that it can be used in adiabatic temperature rise reactions for calculating isocyanate conversion; additionally this script reduces noise in all the data and removes erroneous readings. The final script extracts important information from the corrected data such as maximum temperature change and maximum height change as well as the time to reach these points. Commercial examples of such equipment exist however the price (>£10000) of these equipment make these systems inaccessible for many research laboratories. The FoamPi build presented is inexpensive (£350) and test examples are shown here to indicate the reproducibility of results as well as precision of the FoamPi.
... Finally, X w is the conversion of water molecules, which react with isocyanate to form a urea bond and carbon dioxide. Most complete model approaches include a system of partial differential equations describing the reactions taking place in the polymer to calculate the gelation rate (Al-Moameri et al., 2015;Ghoreishi et al., 2014;Shen et al., 2014;Zhao et al., 2014Zhao et al., , 2013Zhao and Suppes, 2015). However, a perfect knowledge of the existing reactions is required, such as: ...
Article
This study presents the estimation of the parameters of the polymerization kinetics of a mimosa tannin-based thermosetting resin. A dual approach, experimental and numerical, was used. The numerical approach consisted in solving a time-dependent 0D numerical model of the polymerization kinetics and the heat equation. Thus, the parameters were estimated by minimizing the difference between the measured and the simulation values for four different polymerization kinetics. During the mixing phase, the temperature was recorded in order to observe the impact of the mixer and the introduction of the polymerization catalyst on the temperature. The modelling showed that the reaction starts directly with the introduction of the catalyst and that this data should not be neglected in order to achieve the minimization. A numerical study on the effect of the simulation time showed a very limited impact on the estimation of the parameters. A simulation time of 350 s was chosen in order to better take into account heat losses. The four polymerization kinetics were consistent with the experimental data and the fit improved from kinetics #1 to kinetics #4 as the number of fitting parameters increased, but the results were of the same order of magnitude. In conclusion, this work presents a simple method from an experimental point of view but very effective for estimating the reaction kinetic parameters of a thermosetting resin based on mimosa tannin. The method can probably be adapted to other polymer systems.
Article
Full-text available
The present work describes the synthesis of new versatile polyurea (PU) and polyurethane (PUR) matrices, including different chain extenders, which facilitate the design of distinct, tunable properties, and high-performance derivatives. These polymers can be used for various defense and security applications, such as coatings for ballistic protection, CBRN protection, binders for energetic formulations, etc. Combining aliphatic and aromatic molecules in PU or PUR structures enables the synthesis of polymers with improved and controllable thermo-mechanical properties. Thus, for polyurea synthesis, we utilized two types of polymeric aliphatic diamines and three types of aromatic chain extenders (1,1’-biphenyl-4,4’-diamine, benzene-1,2-diamine, and 1,2-diphenylhydrazine). An analogous method was used to synthesize polyurethane films by employing one polymeric aliphatic polyol and three types of aromatic chain extenders (benzene-1,3-diol, benzene-1,4-diol, and benzene-1,2,3-triol). Subsequently, various analytic techniques (Fourier transform infrared spectroscopy–attenuated total reflectance (FTIR-ATR), single cantilever dynamic mechanical analysis (DMA), thermogravimetric analysis (TGA), frequency-dependent shear modulus survey, tensile tests, water contact angle measurements, and scanning electron microscopy (SEM) with energy-dispersive X-ray analysis (EDX)) have been utilized to characterize the synthesized materials and to evaluate the influence of each chain extender on their final properties.
Article
Through systematical experiment design, the physical blowing agent (PBA) mass loss of bio-based polyurethane rigid foam (PURF) in the foaming process was measured and calculated in this study, and different eco-friendly PBA mass losses were measured quantitatively for the first time. The core of the proposed method is to add water to replace the difference, and this method has a high fault tolerance rate for different foaming forms of foams. The method was proved to be stable and reliable through the standard deviations σ1 and σ2 for R1 (ratio of the PBA mass loss to the material total mass except the PBA) and R2 (ratio of the PBA mass loss to the PBA mass in the material total mass) in parallel experiments. It can be used to measure and calculate the actual PBA mass loss in the foaming process of both bio-based and petroleum-based PURF. The results show that the PBA mass loss in PURF with different PBA systems is controlled by its initial mass content of PBA in PU materials ω. The main way for PBA to dissipate into the air is evaporation/escape along the upper surface of foam. This study further reveals the mechanism of PBA mass loss: the evaporation/escape of PBA along the upper surface of foam is a typical diffusion behavior. Its spread power comes from the difference between the chemical potential of PBA in the interface layer and that in the outside air. For a certain PURF system, R1 has approximately linear relationship with the initial mass content of PBA in PU materials ω, which can be expressed by the functional relationship R1 = kω, where k is a variable related to PBA’s own attributes.
Article
Full-text available
We describe experiments designed to inform computational models of the dynamic filling process of chemically blown, polyurethane foams, especially subgrid models to predict bubble size affecting foam properties. Three experimental methods are used to observe the evolution of bubble sizes during blowing. Magnified views of bubbles at a transparent wall of a channel are recorded during the foaming. The bubble sizes in the final frame after the expansion has stopped are compared to scanning electron microscope images of the interior of the cured samples to determine wall effects. In addition, diffusing wave spectroscopy is used to determine the average bubble sizes across the width of a similar channel during foam expansion. We conclude that the bubble size distribution is dependent on the formulation of foam being tested, temperature, the height in the foam bar, the proximity to a wall, and the degree of overpacking.
Article
The properties of rigid polyurethane foams can be modified over a wide range depending on the raw materials used for its synthesis. The polyols, chain extender and blowing agent have major impact on the properties of the PUF. This investigation reports the effect of polyol and its blends, chain extender such as 1,4-butanediol (1,4-BD) and blowing agents [chemical as well as physical; such as distilled water and n-pentane(n-C5),hydrochlorofluorocarbon(HCFC)] on the properties of rigid PUF. Addition of the chain extender improves the mechanical properties initially but then it decreases the properties on further addition of the same. The thermal conductivity of the chain extended and that of the physically blown PUF was lower than the water blown PUF. Higher loading of chain extender and blended polyol decreased the glass transition temperature (Tg) of the PUF due to the decrease in cross-link density because of low -OH functionality in the presence of linear dials like hydroxyl terminated polybutadiene (HTPB). It was observed that there is no significant change in degradation temperature of the chain extended, physically blown and blended system except that observed in the case of HTPB filled PUF.
Conference Paper
This study investigated the effect of incorporating microsphere and nanoclay fillers from 1-7% on the physical properties of polyurethane (PU) foams made from a polyol containing 15% soybean oil based polyol. With increasing filler percentage, the PU foam volume increased because these fillers provided surfaces for nucleation and more gas bubbles were generated during the foaming process. The compressive strength of PU foams decreased slightly when increasing the microsphere content from 1 to 3% and then increased when the filler was higher than 3% . At 7% microsphere content, the foams displayed the same compressive strength as the control foams made from 100% petroleum polyol. For PU foams reinforced with nanoclay, their compressive strength changed little from 1 to 5%, but decreased at 7% due to a lower density. Foams containing 5 to 7% microspheres or 3 to 7% nanoclay had density-compressive strength comparable or superior to the control. SEM was used to observe the morphology of reinforced foams. Foams reinforced with fillers had more cells and smaller cell size than foams made from 15% soy-polyol but without fillers. During the foaming process, the maximal temperatures reached by PU foams containing were not affected by the presence of 1 to 7% of microspheres or nanoclay, but slightly lower than the control. In addition, foams with fillers displayed roughly the same thermal conductivity as soy-polyol based foams without fillers.
Article
The objective of this study was to investigate the effects of isocyanate/hydroxyl ratio and ammonium polyphosphate (APP) content on the properties of polyurethane foam. Polyurethane (PU) foam was prepared from polymeric diphenylmethane diisocyanate and polyethylene glycol with molecular weight of 200, reinforced with oil palm empty fruit bunch (EFB) using one shot process. The effect of EFB content on the properties of PU foam was also studied. It was noticed that EFB enhanced the properties of the PU foam. This was due to EFB acting as hard segment in PU foam system. The NCO/OH ratio played an important role in determining the properties of the PU foam produced. However, since EFB is a highly flammable material, APP was introduced to the PU foam system. From the results, APP improved the fire retardant behavior of the PU foam. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012
Article
This study investigated the physical properties of water‐blown rigid polyurethane (PU) foams made from VORANOL®490 (petroleum‐based polyether polyol) mixed with 0–50% high viscosity (13,000–31,000 cP at 22°C) soy‐polyols. The density of these foams decreased as the soy‐polyol percentage increased. The compressive strength decreased, decreased and then increased, or remained unchanged and then increased with increasing soy‐polyol percentage depending on the viscosity of the soy‐polyol. Foams made from high viscosity (21,000–31,000 cP) soy‐polyols exhibited similar or superior density‐compressive strength properties to the control foam made from 100% VORNAOL® 490. The thermal conductivity of foams containing soy‐polyols was slightly higher than the control foam. The maximal foaming temperatures of foams slightly decreased with increasing soy‐polyol percentage. Micrographs of foams showed that they had many cells in the shape of sphere or polyhedra. With increasing soy‐polyol percentage, the cell size decreased, and the cell number increased. Based on the analysis of isocyanate content and compressive strength of foams, it was concluded that rigid PU foams could be made by replacing 50% petroleum‐based polyol with a high viscosity soy‐polyol resulting in a 30% reduction in the isocyanate content. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013
Article
The effect of hollow glass microspheres with a density of 125 kg/m3 on the properties of low-density (54-90 kg/m3) rigid polyurethane foams is investigated. The thermal expansion coefficient of the foams and their properties in tension and compression in relation to the content of the microspheres (0.5-5 wt.%) are determined. An increase in the characteristics of the material in compression in the foam rise direction with increasing content of filler is revealed. The limiting content of the microspheres above which the mechanical characteristics of the filled foams begin to decrease is found. The distribution of the microspheres in elements of the cellular structure of the polyurethane foams is examined.
Article
A soypolyol based on epoxidized soybean oil (ESO) was prepared in the presence of HBF4 and diethanolamine (DEA) was used as ring opener. A series of polyurethane rigid foam were prepared by mixing polyol with TDI using an isocyanate index of 1.1. The polyol used in this paper were a mixture of soypolyol and a commercial PL-5601 polyester polyol and the mass fraction of PL-5601 was in the range of 0–60%. The thermal properties of the resins were characterized by DSC and TG. The results showed that these rigid foams possess high thermal stability. There were two glass transition temperature of each foam and Tg1 was increasing with the increasing of OH value. The compression strength of the foam was also recorded, and the effect of mass ratio of soypolyol and PL-5601 polyester polyol on the compression strength was discussed.
Article
Cellular solids include engineering honeycombs and foams (which can now be made from polymers, metals, ceramics, and composites) as well as natural materials, such as wood, cork, and cancellous bone. This new edition of a classic work details current understanding of the structure and mechanical behavior of cellular materials, and the ways in which they can be exploited in engineering design. Gibson and Ashby have brought the book completely up to date, including new work on processing of metallic and ceramic foams and on the mechanical, electrical and acoustic properties of cellular solids. Data for commercially available foams are presented on material property charts; two new case studies show how the charts are used for selection of foams in engineering design. Over 150 references appearing in the literature since the publication of the first edition are cited. It will be of interest to graduate students and researchers in materials science and engineering. © Lorna J. Gibson and Michael F. Ashby, 1988 and Lorna J. Gibson and Michael F. Ashby, 1997.
Article
Mechanics analyses are used to derive the effective elastic moduli for low density materials. Both open cell and closed cell geometric models are employed in the case of isotropic media. The five independent effective moduli are derived for a low density transversely isotropic medium. Compressive strength, as defined by elastic stability, is also derived for open cell and closed cell isotropic materials. The theoretical results are compared with some experimental results, and also are assessed with respect to previous work.
Article
The energy-absorbing characteristics of a foam are determined by its load–compression response, and hence reflect the geometric structure and physical properties of the matrix material. In this report, the energy-absorbing characteristics are expressed in terms of three dimensionless quantities: (1) K, the energy-absorbing efficiency, (2) I, the impact energy per unit volume divided by Ef, and (3) I/K, the maximum decelerating force per unit area divided by Ef, where Ef is the apparent Young's modulus. Using the calculation procedures described in this report, it is now possible to delineate the geometric structure and physical properties a foam matrix must possess to meet a given energy absorption specification. This approach shows that: (1) the energy-absorbing characteristics of a brittle foam are superior to those of a ductile foam, (2) the optimum energy-absorbing foam has a large cell size, a narrow cell size distribution, and a minimum number of reinforcing membranes between the cells, (3) foam composites offer no significant advantage over a single foam, and (4) the optimum energy-absorbing region obtains over a tenfold change in impact velocity and can be extended in a given system only if the foam stiffness increases while the impact velocity is increased, as in a fluidfilled foam.