You are on page 1of 20

European Polymer Journal 73 (2015) 500–519

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Macromolecular Nanotechnology – Review

Microcellular to nanocellular polymer foams: Progress


(2004–2015) and future directions – A review
Chimezie Okolieocha, Daniel Raps, Kalaivani Subramaniam, Volker Altstädt ⇑
Department of Polymer Engineering, University of Bayreuth, Universitätsstrasse 30, 95447 Bayreuth, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The importance of low-density microcellular closed-cell thermoplastic foams over conven-
Received 17 August 2015 tional macrocellular ones has been demonstrated over the years in several areas of appli-
Received in revised form 24 October 2015 cations. But more advanced and future expectations in terms of material cost and superior
Accepted 1 November 2015
properties (thermal, mechanical, electrical, etc.) have necessitated the need to move away
Available online 11 November 2015
from fabricating microcellular to nanocellular foams. A lot of studies in this direction have
yielded results especially in terms of cell size reduction but their corresponding foam den-
Keywords:
sities remain either too high or quite close to the density of the non-foamed material or are
Polymer foam
Closed-cell
simply omitted in several published articles. This review article summarises the progress
Low-density made in the last decade by revealing polymer foams with the smallest mean cell sizes
Nanocellular foam and lowest foam densities using the conventional foam processing techniques (batch foam-
Microcellular foam ing, foam extrusion and foam injection moulding).
Expansion ratio Ó 2015 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 501
2. Foaming basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
2.1. Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
2.2. Polymer–filler interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
2.3. CO2 sorption and diffusivity in polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
2.4. Terminologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
2.5. Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
3. Preparation of microcellular and nanocellular thermoplastic polymer foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
3.1. Batch foaming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508
3.2. Foam extrusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510
3.3. Foam injection moulding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510
4. Thermoplastic polymers foams with the smallest mean cell sizes and densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511

Abbreviations: HDPE, high-density polyethylene; LDPE, low-density polyethylene; PC, polycarbonate; MAM, poly(methyl methacrylate)-b-poly(butyl
acrylate)-b-poly(methyl methacrylate); PCL, poly(caprolactone); POE, poly(ethylene-octene); PEN, poly(ethylene 2,6-naphthalate); PES, polyethersulfone;
PETG, poly(ethylene terephthalate glycol); PHVB, poly(hydroxybutyrate-co-hydroxyvalerate); EVA, ethylene vinylacetate copolymer; PEEK, poly(ether
ether ketone); PEI, polyetherimide; EPDM, ethylene–propylene-diene terpolymer; PFDA, poly(tetrahydroperfluorodecyl acrylate); COC, cyclic olefin
polymer; PBS, poly(butylene succinate).
⇑ Corresponding author.
E-mail address: altstaedt@uni-bayreuth.de (V. Altstädt).

http://dx.doi.org/10.1016/j.eurpolymj.2015.11.001
0014-3057/Ó 2015 Elsevier Ltd. All rights reserved.
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 501

4.1. Nanocellular and microcellular foams produced via batch foaming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
4.2. Extruded polymer foams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
4.2.1. Poly(lactic acid) (PLA) as a base matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
4.2.2. Polypropylene (PP) as a base matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
4.2.3. Polystyrene (PS) as a base matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
4.2.4. PE as a base matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
4.2.5. Copolymers as a base matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
4.3. Injection moulded foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
5. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
Words used interchangeably. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516

1. Introduction

The development of polymeric foams started in the 1930s with the first patented foam (1931) being macrocellular poly-
styrene foams (with cell size above 100 lm). Further development continued until early 1980s when solid-state batch foam-
ing (a technique that enables the dissolution of gas [chiefly CO2 or N2] into a solid plastic and by subsequent change of phase,
tiny bubbles of less than 100 lm in diameter [1], form within the plastic material and remain even after the plastic solidifies)
was pioneered by Professor Suh et al. [2] at Massachusetts Institute of Technology (MIT). The principal motive behind this
invention was to (i) reduce material density (save plastic) without compromising its properties and (ii) decrease cost. Since
then, plastic foams in comparison to non-foamed plastics have found several applications in automobile, aeronautic, pack-
aging and construction industries, etc. [3–6] owing to their light-weight, high ratio of load-bearing strength to weight, low
thermal conductivity and high dissipation of impact energy.
Advancements in foam processing technology from late 1980s to late 1990s led to the development of foams classified as:
fine-celled foams (cell size between 10 lm and 100 lm) and microcellular foams (cell size of 610 lm and cell density of
>109 cells/cm3). These classifications however failed to specify the foam density range or expansion ratio to be fulfilled.
Furthermore, microcellular foams compared to macrocellular foams at same density, offered improved properties partic-
ularly mechanically due to specifically, their smaller cell sizes and cell size uniformity.
Likewise, the fact that overall, better foam properties were obtained from microcellular foams triggered in the early twen-
tieth century (i.e. early 2000s) a paradigm shift towards nanocellular foams (cell size <1 lm and cell density >1015 cells/cm3).
It is expected that nano-sized foams will offer even more superior properties. In the light of this, most foam studies have
focused primarily on decreasing further the cell size at the expense of foam density. For instance, for thermal insulation pur-
poses, it is well known that when the cell size of a foam is less than the mean free path of gas molecules (70 nm) the so-called
‘‘Knudsen effect” [7–9] occurs which helps to decrease tremendously the thermal conductivity of the gas-contributing com-
ponent (accounts for about 50% of the overall thermal conductivity). In addition to that, the foams must have a closed-cell
structure (Fig. 1) and must also be of low-density i.e. they must have a high expansion ratio >40).
Closed-cell thermoplastic foams can be made by one of the three most widely used processing techniques namely: batch
foaming (a batch-wise process), foam extrusion (a continuous process) and foam injection moulding (a semi-continuous pro-
cess). Therefore, this review article, takes a look at published articles on foam studies covering the last decade (2004–2015),
highlighting the progress and limitations in fabricating low-density-nanocellular foams using the aforementioned
techniques.

(a) (b)

1 mm 100 µm

Fig. 1. (a) Open-cell, (b) closed-cell cellular structure of polymeric foams. Note: Closed-cell foams can have between 1% and 10% open-cell content.
502 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

In addition, graphical representations of the mean cell size versus their corresponding expansion ratio are shown in sub-
sequent sections according to their method of preparation in order to make comparisons and to identify foams with the low-
est cell size and foam densities.
Lastly, the following criteria were chosen to limit the scope of the study:

 Only thermoplastic closed-cell polymeric foams were considered i.e. open-cell foams were excluded.
 Published articles where the cell size and either one of the following (absolute density, expansion ratio or relative density)
were stated were considered to make the plots.
 Foams reported to have bimodal or multi-modal cell size distribution or were huge discrepancies in cell sizes were
reported were not taken into consideration.
 Foams with the following cell sizes were used as a benchmark for the respective processing techniques: cell size 610 lm
for batch foams; cell size 6200 lm for extruded foams; cell size 6450 lm for injection moulded foams.

2. Foaming basics

2.1. Nucleation

Nucleation is simply defined as the first random formation of a distinct thermodynamic new phase (daughter phase or
nucleus (an ensemble of atoms)) that have the ability to irreversibly grow into larger sized nucleus within the body of a
metastable parent phase. The resulting new phase has a free energy lower than the parent phase. In relation to polymer
foaming, the daughter phase and parent phase are referred to as the bubble (foam cell) and polymer melt respectively.
The basic driving force for such transformation is the difference in free energies between the two systems. The classical
nucleation theory (CNT) is to date; the most frequently used empirical theory to describe the mechanisms of cell nucleation
and growth in polymeric foams. The following are the postulations and approximations of CNT [10–12].

 The classical nucleation theory assumes that the nucleus (bubble) is a spherical droplet with distinct sharp boundary.
 That all nuclei have the same physical properties as the bulk.
 It takes into account the effect of pressure drop on cell nucleation and disregards the effect of pressure drop rates.
 It also assumes that instantaneous pressure drop is followed by instantaneous nucleation, a scenario which is not the real-
ity because, pressure drop occurs over a span of time.
 CNT also describes the bubble interface as an infinite flat planner surface.

Due to the assumptions and approximations adopted by the CNT, a handful of published works have shown some remark-
able deviations from the theory and these have called to question, the accuracy and validity of CNT [10,13]. However, in spite
of these shortcomings, the CNT provides a powerful conceptual understanding of the nucleation process and therefore shall
be adopted in this article to explain the mechanism of bubble nucleation and growth with respect to polymer foams.

(a) (b)
+ Nucleus formation Nucleus growth
Conical
Contact angle
Newly formed Bubble (nucleus)
angle
Free energy change (∆G)

interface Interface
Θ

β
∆G*hom

∆G*het
Additive
r*

Critical radius Radius


∆Ghet
If R < r* R > r*
Nuclei re-dissolves Nuclei grows Parent phase (melt)
∆Ghom

Fig. 2. Bubble nucleation and growth as a function of free energy (a), contact angle and semi-conical angle for heterogeneous bubble nucleation (b)
modified according to [11].
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 503

The CNT classifies cell nucleation into two types namely: homogenous and heterogeneous nucleation.
Homogenous nucleation involves the spontaneous and random formation of bubbles (cells) with no foreign bodies or addi-
tives (heterogeneous nucleating agents – NAs) or pre-existing cavities present in the embryo formation stage. It usually
occurs in response to random fluctuations or change in temperature or pressure. The free energy barrier reaches a high-
energy maximum when R equals the critical radius (r ) (the critical radius represents the minimum size an embryo should
have before it can grow to become a stable bubble. Below this radius, the embryo collapse into the melt and above this size, a
stable nucleus is formed and then continues to grow) and afterwards decreases (see Fig. 2a).
Furthermore, the stability of the bubble also depends on the competition between the pressure difference DP (the differ-
ence between the internal bubble pressure and the pressure surrounding the sample) and the interfacial energy (surface
energy). By taken these factors into account, an expression for the Gibb’s free energy is obtained as shown in Eq. (1) [14].
4 3
DGðRÞ ¼  pR DP þ pR3 cab  DV free v ol ð1Þ
3
The dependence of r the pressure difference (DP) and interfacial energy (cab ) is expressed in a relationship in Eq. (2) and the
homogenous Gibb’s free energy barrier DGHom is given in Eq. (3).
2cab
r ¼ ð2Þ
DP

16pc3ab
DGHom ¼ ð3Þ
3 DP 2
Since homogenous nucleation occurs in the absence additives, the DV term in Eq. (1), which accounts for free volume due
to the presence of additives can be ignored. Therefore, the rate of homogenous nucleation can be calculated from the expres-
sion in Eq. (4).

Table 1
Some commonly used NAs and BAs as reported in the literature (patents included) [16–41].

Heterogeneous nucleating agents Blowing agents (BAs)


Carbon black (CB),a,b furnace black,a channel black,a acetylene black,a ceramic Aliphatic hydrocarbons
microspheres,a powdered amorphous carbon,a graphite (natural or milled),a,b mica,a n-Pentane, isopentane, methane, ethane
granulated asphalt,a metal flakes, black Iron oxide, titanium dioxide (TiO2),a,b carbon Ethers
nanofibers (CNF),a activated carbon (AC),a carbon nanotubes (CNT),a chemically reduced Dimethyl ether, vinyl methyl ether
graphite oxide (CRGO),a,b thermally reduced graphite oxide (TRGO),a,b graphite oxide Ester
(GO), expanded graphite (EG), graphite nanoplatelets, talc,c ethyl vinyl acetate (EVA),c Methyl formate
ethylene methyl acrylate (EMA),c polyethylene glycol (PEG),c boron nitride, crystalline Inorganic BAs
silica (SiO2), silicates (calcium, sodium, lithium), modified layered silicates (MMT, Carbon dioxide, nitrogen, water
OMMT), clay, (organic and inorganic), azodicarboxylic acid diamide, nanoclays Argon, helium, xenon, ammonium
(cloisites), zeolites, supramolecular compounds, diatomaceous earth (inorganic), Halogenated hydrocarbons
sodium hydrogen tartrate (organic), sodium and potassium hydrogen succinate Chlorofluorocarbons (CFCs)
(organic), calcium carbonate (CaCO3), sodium bicarbonate (NaHCO3), stearates (barium, Hydrochlorofluorocarbons (HCFCs)
calcium, sodium), barium sulphate, wood fibres, flax fibres, cellulose fibres, pine Hydrofluorocarbons (HFCs)
(standard softwood), Alkali lignin, glass fibres, ethyl lactate, spherical ordered Difluoromethane, trichlorofluoromethane
mesoporous silica (s-OMS), magnesium oxide, magnesium hydroxide (Mg(OH)2), Dichlorofluoromethane
aluminium trihydroxide (ATH), nanowhiskers (chitin, cellulose), waste ground rubber Boron trifluoride
tire powder, expandable thermoplastic microspheres (ETM), silk fibroin powder Ethylchloride
Chlorotrifluoromethane
Methyl chloride
Aliphatic alcohols
Ethanol, methanol, ethyl-hexanol
Isopropanol, etc.
Aliphatic cyclohydrocarbons
Cyclobutane, cyclohexane, isopropanol, water,
isobutane, 4-methylbenzene-1-sulfonohydrazide
Chemical BAs (endo- and exothermic)
Azobisisobutyronitrile (AIBN)
Sulphohydrazide
Bicarbonates or carbonates
Azodicarbonamide (ADC)
2-methacryloyloxyethylhy-hexadecyldimethyl
ammonium bromide (MHAB)
Sodium bicarbonate in combination with citric acid
a
Infrared attenuating agents (IAA).
b
Heat reflectors.
c
Cell enlarging/anti-blocking agents.
504 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

 
DGhom
NHom ¼ f 0 C 0 exp ð4Þ
KBT

where f 0 is the frequency factor of the gas molecules joining the nucleus (it varies with the surface area of the nucleus); C 0 is
the concentration of gas molecules; K B is the Boltzmann constant; T is the temperature and DGhom is the energy barrier for
homogeneous nucleation.
Heterogeneous nucleation on the other hand, does not occur spontaneously. It is triggered by presence of additives (with
low activation energy) that act as nucleation centres. The introduction of additives (assumed to be plate-like and with planer
surface) lowers the Gibb’s free maximum but this depends on the particle type and topography of the surface (shape factor
(S)). These factors taken into account, yields the following expressions in Eqs. (5) and (6).

DGHet ¼ DGHom SðhÞ ð5aÞ

16pc3ab
DGHet ¼ SðhÞ ð5bÞ
3 DP 2

1
SðhÞ ¼ ð2 þ 3 cosðhÞ  cos3 ðhÞÞ ð6Þ
4
Also, the shape factor (S) depends on the wetting or contact angle (h) between the bubble of radius ðRÞ and the surface of the
additive. S decreases as h increases and therefore lead to further reduction of DGHet . Acknowledged that in reality, some stan-
dard heterogeneous nucleating agents (NAs) such as talc are not plate-like but particle-like, nucleation is assumed to occur in
a conically shaped pit (conical cavity) i.e. the preferred nucleation site defined by the semi-conical angle (b) (see Eq. (7)) [15]
as shown in Fig. 2b. The size of b will influence the work required to nucleate a bubble. Table 1 shows a collection of the types
of NAs) and blowing agents (BAs) that are commonly used in the literature.
 
1 cos h cos2 ðh  bÞ
Sðh; bÞ ¼ 2 sinðh  bÞ þ ð7Þ
4 Sin b

2.2. Polymer–filler interaction

Polymer–filler interaction is of utmost importance to assess nucleation efficiency (defined as the ratio of the measured
cell density to the potential nucleant density; see Eq. (16)). Interestingly, the expectation that smaller filler particles lead
to higher nucleation densities was found to be true only to a certain lower limit of particle size [42,43]. In another study
conducted by Goren et al. [44] they concluded that there is a competition between the number of nucleation sites (smaller
particles, more sites) and nucleation efficiency (smaller particles, lower nucleation efficiency).
In addition to particle size, the shape and topography of the surfaces as mentioned earlier have been shown to influence
nucleation. For instance, in an article by Fletcher [45], it was reported that particles with flat surfaces offer better nucleation
rate compared to tubular particles because of their curved edges but these arguments were countered in another work by
Chen et al. [46].

(a) (b)
20
PMMA PMMA
Amount of CO2 absorbed (wt %)

18 20
25 °C 25 °C
CO2 mass uptake (wt %)

16

14 15
12

10
10
8 y = -0.5873 + 19.548
6 R2 = 0.9968

4 5

0 0
0 5 10 15 20 25 30 35 40 0 2 4 6 8 10
t (h) t1/2(min1/2)

Fig. 3. (a) CO2 sorption curve, (b) CO2 desorption curve of poly(methyl methacrylate) (PMMA).
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 505

Leung et al. [47] studied the effect of wetting angle (h) and the angle of a conical cavity (b) of an NA. They found out that
the difference ðh  bÞ determines the growth state of a bubble. Both angles affect the free energy in such a way that increas-
ing h leads to a decrease of DGHet , whereas the effect of b is dependent on h. It was also shown that when h  90 , smaller b
improved heterogeneous nucleation. But when both angles are small, heterogeneous nucleation is not favourable anymore.
Conclusively, an ideal nucleating agent must fulfil the following criteria [48]: (i) They must not have very strong adhesion
to the matrix (equivalent to a large wetting angle). (ii) They should be easily dispersed and properly exfoliated within the
polymer matrix. (iii) They must be present in sufficient amount to foster the creation of more nucleation centres. (iv) The
nucleants should be characterized by uniform size, shape and surface properties.

2.3. CO2 sorption and diffusivity in polymers

One of the advantages of foam processing is that it can be done at relatively low temperatures. This is made possible by
the plasticizing effect of the blowing agents (BAs) on the chains of the polymer into which they are incorporated. The sorp-
tion capacity (i.e. how much BA the polymer matrix can absorb) especially of an amorphous polymer under certain pressure
and temperature, determines the extent to which the glass transition temperatures (Tg) of the polymer can be depressed.
Goel and Beckman [49] demonstrated that the Tg (105 °C) of poly(methyl methacrylate) (PMMA) was reduced to ambient
temperature due to its high CO2 absorption capacity which is attributed to the Lewis acid–base interaction between its car-
bonyl oxygen and CO2 [50].
There are basically two widely used methods to estimate or calculate the total amount of CO2 absorbed by a polymer. The
first method involves saturating the polymer sample in a high-pressure autoclave just like in the first step in Fig. 4b. Peri-
odically, the autoclave is depressurized at relatively low temperature and the sample is taken out and the weight is mea-
sured. It goes on until no change in weight is observed which means the sample has been fully saturated (i.e. it has
attained its equilibrium gas concentration). The mass uptake of CO2 at any given time is calculated using Eq. (8) [51,52]
where Mgas t is the gas concentration at time t, W after time t is the mass of the sample after time t and W initial is the initial mass
of the sample.
 
W after time t  W initial
Mgas t ¼  100% ð8Þ
W initial

Mt =M 0 ¼ 4ðDd =pÞ1=2 ðt1=2 =LÞ ð9Þ


Fig. 3a shows as an example of how a plot of the sorption time versus the mass uptake (wt%) for PMMA looks like. The
limitation of this method is that it does not provide any data about gas diffusivity of the system and does not estimate accu-
rately the amount of gas uptake by the sample. There is usually a loss of gas (less than 10% of its original value) as the auto-
clave is depressurized and during the time the sample is removed from the autoclave to be weighed. A more accurate
method is to use a magnetic suspension balance which gives a more accurate measure of gas solubility but they have pres-
sure limitations of about 25 MPa.
The second method involves saturating the sample to equilibrium (values are usually known through the first method);
removing the sample from the autoclave and subsequently measuring the weight loss of the sample at certain times as CO2
diffuses out of the sample. In order to obtain the equilibrium concentration of CO2 in the polymer (M0) without any desorp-
tion i.e. estimation of the total amount of CO2 at time, t = 0, a plot of Mt (the amount of CO2 absorbed by the polymer at time
t) versus the square root of the of desorption times recorded (Fig. 3b). By fitting the desorption curve linearly and extrapo-
lating to zero, M0 is obtained. Assuming that the sample thickness L ðcmÞ remains unchanged, Fick’s second law in Eq. (9)
[53] can be applied and from the slope ð4=LÞðDd =pÞ1=2 of the curve, the diffusion coefficient Dd ðcm2 =min or cm2 =sÞ is
obtained.

2.4. Terminologies

In this section, we intend to describe certain properties that are commonly used to characterize polymer foams.
Foam density: is a foam property evaluated by water-displacement method based on Archimedes principle. What is basi-
cally measured is the mass of the sample in air (A) divided by the mass of the sample in water (B) and by using the expression
in Eq. (10), the foam density (qfoam ) is obtained.

A
qfoam ¼ ð10Þ
AB
Relative foam density (qrel ): is the qfoam divided by the density of the solid polymer (unfoamed polymer) and it is the
inverse of the volumetric foam expansion ratio.
qfoam
qrel ¼ ð11Þ
qpolymer
506 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

A minimum of three scanning electron microscope images of the foam samples are usually employed for cell size, cell size
distribution and cell density. Typically, about 100–200 cells are selected in a defined area of the 2 dimensional SEM pictures
and the following estimations are made:
Cell density: The cell density (N o ) is expressed as the number of cells per unit volume with respect to either the foamed
polymer (Eq. (12)) or with respect to the unfoamed solid polymer (Eq. (13a)).
3=2
No ¼ ðnM 2 =AÞ ð12Þ

3=2
No ¼ ðnM 2 =AÞ ð1=1  V f Þ ð13aÞ
!
qfoam
Vf ¼ 1  ð13bÞ
qpolymer
where n is the number of cells selected for evaluation, A is the total area occupied by the selected cells, M is the magnification
factor. V f is the void fraction of the foam (Eq. (13b)). The purposes these two different equations serve are explained in
greater details elsewhere [54,55].
Cell size: The cell size is usually the average value of all foam cell sizes calculated from various scanning electron micro-
graphs. The cell distribution is also another important characteristic to consider as it gives indication of how homogeneous
the foam morphology is. It can also be calculated according to Eq. (14) [56] if all cells are assumed to be spherical in shape.
" !#1=3
1 6 qfoam
£3D ¼   1 ð14Þ
N0 p qpolymer
Nucleation density (qnucl ): Gives an indication of the nucleation efficiency of the NAs added into the foam systems. It is
expressed as the number of nuclei or nucleated cells per cubic centimetre of the unfoamed polymer assuming the absence
of cell coalescence and is given by Eq. (15) [56,57]. This calculation is also based on the assumption that the cells are
spherical.
N0
qnucl ¼ ð15Þ
1  Vf
A detailed knowledge about the nucleation efficacy of the NAs can be estimated from Eq. (16) [58], which reveals the
potential nucleation density of the NAs (N Nul ). Note, it is different from the actually cell density of the foam.
Nucleants wp q
NNul ¼ ¼  c ð16Þ
cm3 qpolymer V p
where wp is the weight fraction of the NA in the polymer composite, qpolymer and qc are the densities particle and the polymer
and V p is the volume an individual NA particle.

2.5. Classification

Polymer foams are generally classified according to of the following:

 Cellular structure: This refers to how the cells are connected to each other within the foam. i.e. if it has an open-cell or
closed-cell structure. When the cells are interconnected (each cell is connected to other cells through holes), they are
referred to as having open-cell structure (Fig. 1). But if the cells exist as a separate, discrete entity, they are described
as closed-cell.
 Wall stiffness: Foams with cell walls that collapse when pressed are said to be flexible while, those with stiff cell walls are
referred to as rigid foams.
 Volume expansion ratio (VER): (inverse of relative density): Is the ratio of the bulk density of the pure or compact polymer
material to the bulk density of the foamed sample. Based of this definition foams can be classified into: high-density (if
VER 6 4), medium-density (if VER P 4–10) and low-density (if VER P 10–50).
 Cell size: Foams characterized by a mean cell size of less than 10 lm and cell density of between 109 and 1012 cells/cm3
are microcellular foams. Nanocellular foams on the other hand, are foams that have a mean cell size of less than 1 lm and
corresponding cell densities greater than 1015 cells/cm3 [59]. This classification type does not take into account the foam
density, a property, which happens to be one of the most important properties next to cell size.

3. Preparation of microcellular and nanocellular thermoplastic polymer foams

Since the main focus of this article is on low-density closed-cell foams, we shall discuss briefly the three widely used
foam-processing methods with each having its unique features and modus operandi. Table 2 describes the basic similarities
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 507

Table 2
Similarities and differences between the three different foam processing methods.

Bases of comparison Batch foaming Foam extrusion Foam injection Moulding


Process type Batch-wise Continuous Semi-continuous
Purpose Fundamental foam studies of new For economically feasible and large- For fabricating more complicated
materials scale foam production foam parts
Material amount Small amount (in grams) Large amount is needed (usually in Medium to large amount (in
required kilograms) but depends on the kilograms) is needed, but depends on
machine throughput the machine capacity
Pre-moulding step Pre-moulding step such as injection Not required, as it is a moulding tool Not needed just as in foam extrusion
moulding, compression moulding or in itself
casting is normally done to mould the
samples in the necessary form as needed
Screw type Not needed Long single or double (usually co- Shorter single reciprocating screws
rotating) screws compared to screws used in foam
extrusion
Addition of nucleating Foaming composition is fixed from the Composition can be changed at any Nucleating agents can be introduced
agents/process onset. Must be done in the pre-moulding time. i.e. Nucleating agents can be also at anytime during processing
flexibility step using either injection moulding or introduced at anytime during
extrusion, etc. processing
Filler dispersion/mixing Determined from the result of the pre- Better mixing quality compared to Adequate
quality moulding step foam injection moulding
Sample state during gas Solid-state Melt state Melt state
loading/saturation Done conventionally below Tg for Metered above Tg for amorphous Done above Tg for amorphous
temperature Amorphous polymers and above Tm for polymers and above Tm for semi- polymers and above Tm for semi-
semi-crystalline polymers crystalline polymers crystalline polymers
Blowing agent supply Polymer sample is saturated with the Blowing agent is metered (i.e. a Blowing agent is metered but not
blowing agent until equilibrium is specific amount of gas is supplied) more than the melt can absorb at a
reached but not more than the melt can take certain processing condition
Blowing agent uptake Moderate to high Low to high Low to moderate
Maximum achievable Up to 400 bar (40 MPa) Up to 250 bar (25 MPa) Up to 350 bar (35 MPa)
pressure presently
(from literature)
Temperature Experimental temperature is fixed but Offers more opportunities to use It allows for different temperatures
adjustment and can be changed from one level to another. different temperatures due to more at each zone to be set independently
control This method offers only one temperature available temperature zones
zone
Driving force for Pressure drop or change in temperature Pressure drop Pressure drop
nucleation
Shear impact on No shear, therefore sensitive materials High shear Higher shear
polymer can be processed
Cell density range 106–1016 cells/cm3 104–1011 cells/cm3 104–108 cells/cm3
Cell size distribution Uniform cell size distribution can be Although uniform cell size It is difficult to obtain foams with
achieved distribution can be achieved, uniform cell morphology because the
sometimes the diameter of the cells filling and holding time are coupled
found at the core are different in size and this result to cells being
from those found at the edges. This nucleated at different locations at
maybe due to faster surface cooling different times (Fig. 8)
been achieved than cooling in the
middle
Foam density reduction Up to 80% Up to 80% Up to 60%
achievable
Foam volume Moderate to high expansion Low to very high expansion Low to moderate expansion
expansion ratio
Surface quality Good Good and often glossy Usually has poor surface finish but
can be improved using gas-counter-
pressure
Foam skin Light skin Light skinned foams Thicker skinned foams
Size of foam parts that Small sample parts that can be used to Simple but large foam parts such as From simple to complicated foam
can be produced make some very basic foam foam slabs parts like foam sandwiches
measurements
Shaping element/ Foams take the shape of the preformed The geometry of the die determines The mould cavity fixes the shape of
flexibility sample but are often irregular. Shape of the shape of the foam. Foams exiting the resulting foam. Foams cannot be
foam cannot be controlled the die can be calibrated. Allows for calibrated. The moulds can also be
more flexibility of changing dies changed if needed
Total experimental Varies from 30 min to 72 h depending on Shorter times (2–15 min) Shorter time (cycle time) in
time required to polymer type and sample thickness comparison to batch process
obtain a foam
sample
Evaluation of multiple Possible with many sample holders in the Limited/not possible Limited but possible with a multi-
specimens pressure autoclave complex tool

(continued on next page)


508 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

Table 2 (continued)

Bases of comparison Batch foaming Foam extrusion Foam injection Moulding


Tooling cost Relatively very cheap compared to the Expensive depending on machine size Expensive and depends on machine
other two methods and capacity. Additional cost is capacity. Cost of additional part e.g.
incurred from other peripherals like mould may escalates the price
melt pump, calibrator, etc.

Pressure is released
usually at higher
Pressure temperature
release valve

Pressure
Autoclave

1a 1b

CO2 Sorption Depressurization

Polymer sample Foam cell


Polymer sample
impregnated with nucleation and
CO2 growth
CO2 Cylinder

One-Step

Fig. 4a. Pressure induced batch foaming process ðDP=DtÞ.

and differences between the aforesaid techniques. However, irrespective of the foam processing method, the principle of
foaming is the same as explained in the following steps:

I. The polymer is saturated (in the case of batch-foaming) or impregnated (foam extrusion and injection moulding) with
a blowing agent (e.g. non reactive gas, solvent or supercritical fluid) at sufficiently high pressures and defined temper-
atures (depending on the glass transition temperature for amorphous polymers and melting temperature for semi-
crystalline polymers).
II. Thermodynamic instability is introduced into the super saturated polymer–gas mixture either by a sudden increase of
temperature or a rapid decrease of pressure. At this point, phase separation of the mixture occurs, prompting cell
nucleation to occur due to decrease of gas solubility.
III. Cell growth and stabilization evolve as the gas diffuses from the polymer matrix into the already nucleated cells while
the foam structure gradually forms. Afterwards, the cell structure is stabilized via cooling. Furthermore, the growth
rate of the already nucleated cells are basically controlled by gas diffusion rate, viscoelastic behaviour of the polymer
at that temperature and state of supersaturation, etc. and cell stabilization depends on the cell growth stresses and
melt-strength.

3.1. Batch foaming

Batch foaming as the name implies, is a batch-wise process conducted in an autoclave. i.e. a closed system. It involves the
basic aforementioned sequential foaming steps but can also be conducted as a one-step process also known as pressure-
quench method or pressure-induced method. If a two-step process is followed, it is referred to as the temperature-
induced method. The slight differences in these methods are explained shortly. Note that plastic samples used for such
experiments are usually in the form of a circular disc, rectangular or square in shaped compact materials. The thickness
of the sample which is usually between 0.5 mm and 3 mm is a very important dimensional factor as it can be used to extract
information about gas diffusivity of the system.
In the pressure-induced method (Fig. 4a), the polymer is saturated in an autoclave with a blowing agent as described in
step I and after equilibrium saturation time has been reached (i.e. the time it takes for the polymer sample to be completely
saturated with gas), cell nucleation and growth is induced by rapid depressurization of the system to atmospheric pressure.
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 509

Pressure is released at
relatively low
temperatures to prevent
Pressure
sample from foaming
release valve

Pressure
Autoclave

1b Step 1
1a

Polymer sample Polymer sample


impregnated with
CO2 Cylinder CO2 2a

2b
Step 2

Sample immersed in Sample immersed


water to cool and in oil bath to
stabilise the cells trigger foaming

Fig. 4b. Temperature-induced batch foaming process ðDT=DtÞ.

CO2 gas supply


Pressure sensors
Polymer
pellets
Pump
controller Heat
Foam calibrating system
Feed exchanger
hopper
Die element

Extruded
foam showing
cell
nucleation,
growth &
Screw stabilization
Mixer
Melt pump
Die land

Melt extruder Cooling extruder

Melt Melt + CO2 Homogeneous melt mixture Cell Cell Cell


(two phase system) (a phase system) nucleation expansion stabilization

Fig. 5. Schematic representation of foam extrusion on a tandem-line.

Post foaming is followed either by cooling the sample in a solvent of mixture of solvents or by cooling with air in other to fix
the final cell microstructure. The saturation process usually takes place at higher temperatures compared to the
temperature-induced method.
In the temperature-induced method (Fig. 4b), step I is usually conducted at moderately low temperatures. After the equi-
librium saturation time has been reached, the sample is taken out of the autoclave. Step two here, involves dipping the
510 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

supersaturated sample in hot oil bath set at a fixed temperature of choice (usually between 80 °C and 150 °C) for a specific
time period (typically between 1 s and 600 s) to induce cell nucleation and growth. The sample is taken out of the hot oil
bath and is immersed in a cooling bath filled with water or some other solvent.

3.2. Foam extrusion

The term extrusion originates from Greek roots that simply mean to ‘‘push out” [1]. This phrase rightly describes the plas-
tic extrusion process. Extrusion is a continuous process, which consists of an extruding unit and die-shaping unit. Foam
extrusion also follows the same operational principle as in plastic extrusion but the basic difference is the introduction of
BA at a certain point on the extrusion line.
Most foam extrusion lines used industrially are in tandem (Fig. 5) although single lines are also in use. Polymers in pellet
form are introduced into the first barrel (temperature-controlled) of the melt-extruder through the feed hopper. The pellets
are melted under high pressure at relatively high temperatures. Through an injecting unit located on the barrel, BA such as
CO2 in supercritical condition (Tc > 31 °C, Pc > 7.38 MPa) [60] is injected into the melt and a one-phase homogenous system is
formed from a two-phase systems (polymer–gas) due to high pressure.
The gas-loaded melt is advanced into the second barrel by the motion of the screw. Here, the melt is cooled down (initial
cooling) at a temperature lower than in the melt extruder but still provides adequate mixing. The cells are also prevented
from being nucleated due to high pressure encountered in this barrel. The melt-pump controls the polymer melt flow rate,
independent of the temperature and pressure changes. The heat exchanger provides further cooling for the polymer melt to
suppress cell coalescence. As the polymer–gas melt leaves the extruder through the die, foam cells are nucleated by rapid
pressure drop generated as the mixture exit the die. The cells keep growing until vitrification occurs. It is important to note
that cell nucleation can occur either in the die or after the die as shown in Fig. 6. In most cases, the final part of this process is
calibration and cutting of the extruded foams. An excellent review article by Sauceau et al. [61] is suggested for further
reading.

3.3. Foam injection moulding

Basically, foam injection moulding is an extension of the conventional injection moulding process but with additional gas
injection unit. It comprises of the Injection unit and the clamping unit (Fig. 7). While the former is concerned basically with
melting of polymer, melt accumulation and maintaining the injection pressure during mould filling, the function of the latter
is for precise opening and closing of the mould with the right clamping force, cooling and ejection of final part.
A major difference in comparison to foam extrusion lies in the motion of the screw. While in foam extrusion, the screw
rotation continuously pushes melt forward and out of the extruder through the die, here, the screw rotates and moves back-
wards due to the collection of a pool of gas-loaded melt at the tip end of the screw. Afterwards, the melt loaded with gas is
then injected through the nozzle into the atmospheric-pressured-high-temperature mould cavity. After the melt injection
step is over, a holding step is observed to further fill-in the mould to compensate for shrinkage. The mould is then opened
almost immediately or after a slight delay and a rapid drop in pressure occurs with respect to atmospheric pressure. Cell
nucleation phenomena occur and subsequently, the cells grow, expand and stabilize, forming the cellular foam with a usu-
ally thick conspicuous integral skin which varies with the degree of fast cooling of the mould.
There are currently three widely known foam injection-moulding technologies available to produce microcellular foams
using CO2 as a physical blowing agent. They are MuCell by Trexel Inc. (USA), Optifoam by Sulzer Chemtech AG (Switzerland),
and ErgoCell by Demag (Germany) [62].
Undoubtedly, there are several other modified foam injection moulding technologies with regard to nozzle design, mould
design, etc. that have been developed but these topics fall outside the scope covered by this review article. The following
article [63] is recommended for more elaborate knowledge about foam injection moulding.
There are also some other existing supplemental technologies of the foam injection moulding process such as gas-
assisted-injection moulding (GAIM) [11] that are mainly focused to resolve moulding problems such as foam cells nucleating

(i) (ii)
Extrusion*die
Extrusion*die

Fig. 6. Cell nucleation occurring inside the die (i) and outside the die (ii).
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 511

CO2 gas supply Gas-loaded Cooling Ejected foam part


melt injected at with integral skin
Polymer high pressure
pellets
Pressure
Pump
sensor
controller
Feed
hopper

Screw
Nozzle

Closed mould Mould opened

Injection unit Clamping


unit

Fig. 7. Schematic representation of foam injection moulding.

Melt flow direction

Foaming during cooling Foaming during filling

Skin layer
Skin layer thickness

Skin layer

Far way to filling ends Close to filling ends

Fig. 8. Cell morphology of injection moulded foams during filling and cooling [64]. The problem of cells nucleating at different times due to filling and
holding time can be resolved by gas-assisted injection moulding as reported elsewhere [11].

at different times (see Fig. 8) due to variations of pressure drop in the melt during filling process filling and holding pressures
associated with low-pressure and high-pressure foam injection moulding. Furthermore, this technology has led to improved
quality in part, reduced cycle time, foaming at much lower pressure and lower clamping force and cost reduction as well.

4. Thermoplastic polymers foams with the smallest mean cell sizes and densities

As mentioned earlier, the objective of this review is not to examine the influence of several additive characteristics (sur-
face area, interaction with polymer, particle size, particle geometry or particle dispersion) and processing parameters (die
gap, die land, blowing agent plasticizing power, processing temperatures, etc.) on the foam cell microstructure. Rather, it
presents graphically, a collection of foams with the smallest mean cell size versus volume expansion ratio (VER) published
in the literature in the last 10 years.

4.1. Nanocellular and microcellular foams produced via batch foaming

Based on all the published articles surveyed, including articles where either the foam densities, expansion ratio, relative
densities or cell sizes were omitted, the pressure-induced method was mostly employed by researchers compared to the
temperature-induced method.
Fig. 9 shows a graphical representation of nanocellular and microcellular foams produced and reported in the literature
[65–75,3,76–82,44,9,83–85,14,86,51,87–93] over the last decade and are classified into high density, medium-density and
low-density based on their expansion ratio. It can be clearly seen that the base polymers cut across the entire range of ther-
moplastics (commodity, engineering, high-performance).
512 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

Z. Ma_2013_PC
High density foams Medium density foams J. B. Bao_2012_PS
L.Chen_2011_PMMA
10 D. Miller_2009_PEI
J. E. Weller_PC
M. H. N. Famili_2010_PS
Microcellular foams C. Li_2013_PS
J. A. Reglero-Ruiz_2011_MAM
Y. W. Chang_2006_POE
L. Sorrentino_2011_PES
Y. W. Chang_2014_EPDM
J. Yang_2013_PS
J. A. Reglero-Ruiz_2012_PS-b-PFDA
W. Zhai_2010_PEOc
J. A. Reglero-Ruiz_2010_PMMA
1 J. A. Reglero-Ruiz_2011_PMMA-MAM
Cell size (µm)

Y. M. Corre_PLA
K. Gorren_2010_PMMA
B. Notaro_2015_PMMA
B. Notaro_2015_PMMA-MAM
H. B Zhang_2011_PMMA
C. Okolieocha_2015_PMMA
S. K. Yeh_2013_TPU
S. Costeux_2015_PMMA-co-EMA
Nanocellular foams

S. Costeux_2013_PMMA-co-EMA
J. Pinto_2014_PMMA/MAM
J. Pinto_2014_PMMA
0.1 H. Guo_2015_PC
H. Ruckdäschel_2008_PPE/SAN
M. M. Khorasani_2010_HDPE
Y. H. Lee_2005_HDPE
Y. Ito_2006_PC
S. Costeux_2013_SAN/MMA-EA9
J. A. Reglero_2011_PMMA-MAM
M. P. Tran_2013_PMMA
N. J. Hossieny_2014_TPU

0.01
0 1 2 3 4 5 6 7 8 9 10
Volume expansion ratio

Fig. 9. Illustrates a plot of mean cell sizes versus the volume expansion ratios (VER) of various thermoplastic foams.

Costeux and Zhu [86] used the pressure quench method to foam PMMA-co-EMA (a random copolymer with 50 wt% ethyl
methacrylate – EMA) with a pressure of 30 MPa and a temperature of 40 °C. They reported that with 0.5 wt% of silica, no
considerable change was observed from the amount of CO2 uptake by the composite which was about 30 wt%. With a
post-foaming treatment in water for 3 min, they produced medium-density nanocellular foams with a cell size of 95 nm, cell
density of 8.6  1015 cells/cm3 and volume expansion ratio (VER) of 4.88 (absolute density of 242 kg/m3).
Also, in a similar study, Costeux et al. [94] used the same method to foam PMMA-co-EMA but with 51 wt% of EMA. No NA
was used and the foaming temperature was 5 °C lower than before. The CO2 uptake by the copolymer was up to 40 wt% pre-
sumable caused by the additional 1 wt% EMA in the copolymer. The cell size decreased from 90 nm to 80 nm, while the cell
density and VER increased to about 15  1015 cells/cm3 and 5 respectively. Note: the cell size achieved here is only 10 nm
larger than the theoretical cell size of 70 nm for Knudsen effect but nonetheless they are still far from been classified as
low-density foams.
In another study, Costeux and Bunker [89] employed a 50/50 wt% miscible blend of SAN/MMA-EA (poly(styrene acryloni-
trile/methyl methacrylate-co-ethyl acrylate)) to fabricate foams having a mean cell size of 92 nm, cell density of 4.7  1015
cells/cm3 and VER of 3.5 which by classification are high-density nanocellular foams. The used SAN contains 28 wt% acry-
lonitrile while MMA-EA contains 9 wt% of EA (ethyl acrylate). The foams were fabricated also by pressure-induced method
and were saturated at 35 °C with a pressure of 30 MPa for 6 h at pressure drop rate of 1 GPa/s. The presence of the-MMA-EA
in the blend which has a higher affinity for CO2 (36.6 wt%) compared to SAN (21.9 wt%) was probably exploited to increase
the cell densities.
Guo and Kumar [87] foamed polycarbonate (PC) by saturating the samples at 30 °C for 72 h. The samples absorbed
about 20.4 wt% of CO2. Foaming of the samples were carried out at 70 °C in a hot silicon oil bath (temperature-induced
method) and resulted to high-density foams (670 kg/m3) with a mean cell size of 21 nm, cell density of 4.1  1014 cells/
cm3 and VER of 0.56. This was achieved without having to use any NA.
Miller et al. [70] in a temperature induced batch foam process, saturated PEI sheets (1.5 mm in thickness) at 21 °C with a
pressure of 5 MPa for 280 h. The samples absorbed 40 wt% of CO2 and were foamed in a hot silicon oil bath to obtain foams
with the following characteristics: cell size: 30 nm and VER: 1.43 which by classification are still high-density nanocellular
foams.
Yang et al. [77] via a pressure-induced batch process, saturated PS filled with 5 wt% of spherical ordered mesoporous sil-
ica (OMS). They showed that the CO2 absorption capacity of neat PS and PS/5 wt% OMS were relatively similar absorbing
between 10.33 and 10.53 wt% respectively. They foamed the PS-OMS samples at 120 °C for 18 h under high pressure which
was released under 0.5 s. The samples were cooled afterwards with ice water to fix the morphology. The foams were char-
acterized by a mean cell size of 7.8 lm, cell density of 3.55  109 and a VER of 8.6. These foams are still medium-density but
has the highest VER of all batch foams presented in the plot in Fig. 8.
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 513

A one to one comparison between the mean cell sizes and VER of foams prepared with same polymer by different authors
cannot be made due to slight differences in the amount of BA sorbed, filler type/filler content and variation in processing
parameters. However, the following points could be deduced from our findings:

 The saturation and foaming temperatures account for the major difference in cell size and VER. It was also found that
foams fabricated at pressures higher than 13 MPa and at temperatures above 100 °C or at temperatures higher than
the normal Tg of the base polymer, resulted in foams of medium-densities (VER P 4–10) and bigger cell sizes (>6 lm)
as shown in Fig. 9. The effect of high foaming temperature on lowering the polymer density is more pronounced com-
pared to the effect higher pressure would have [49,95].
 Foam with cell size 61 lm were mostly foams prepared either at relatively low temperatures or at temperatures well
below their normal Tg) and with relatively high pressures of up to 35 MPa.
 Foams with cell sizes as low as 150 nm were fabricated with homopolymers, copolymers or blends that are CO2-philic (i.e.
polymers with high affinity for CO2) or with one of the blocks having high affinity for CO2. These polymers were shown to
absorb at least 20 wt% of CO2 [87,89].
 From all indications, high pressure drop rates and relatively low temperatures are necessary to obtain nanocellular foams
but processing at such low temperatures could produce foams with densities closer to the density of the non-foamed
material as seen with the copolymer (PMMA-MAM) [66].
 Fillers added in moderate amount (between 0.01 and 5 wt%) and having high surface area and affinity for BAs decreased
cell size significantly.

4.2. Extruded polymer foams

Via extrusion, it is comparatively easier in comparison to solid-state foaming to fabricate low-density foams (i.e. foams
with VER P 10) because BAs such as CO2 are metered into the polymers in the melt state under relatively high pressures and
temperatures and the various temperature zones in the extruder makes it possible to control and fine-tune the viscoelastic
properties of the polymer charged with CO2.
Fig. 10 shows a graphical representation of extruded foams reported in the literature [96–99,36,100–105,16,106–115]
showing their mean cell size plotted as a function of VER.
A first look at the graphical plot shows that commodity plastics were predominantly used as base materials for extruded
foams. In the subsequent sub-sections, only a few of the base materials, which yielded foams with smaller mean cell sizes
and high VERs will be compared and discussed.

D. Szegda_2014_PHBV
High Medium Low density foams Wei Liu_2013_PLA
W. Michaeli_2011_EPDM
G. M. Rizvi_2008_HDPE
R. Gendron_2006_PS
100 W. Kaewmesri_2006_PP (HMS)
fine-celled foam

E. H. Tejeda_2005_PE/PP (50/50)
Z. Xu_2005_PP (HMS)
A. P. Ranade_2004_PP
A. Rizvi_2013_PP/PTFE
J. M. Julien_2012_PLA
C. Zhang_2011_PS
C. Okolieocha_2015_PS
10 A. Larson_2013_PLA
Microcellular foam

A. K. Chaudhary_PP
M. Mihai_2010_PLA
Cell size (µm)

S-K Yeh_2010_PS
P. Spitael_2004_PP
A. Ayoub_2011_Corn starch
Y. H. Lee_2007_LDPE
Z. X. Zin_2009_PP
S. Costeux_2015_MMA-co-EA
1
Nanocellular foam

0.1

0.01
0 10 20 30 40 50
Volume expansion ratio

Fig. 10. Plot showing the mean cell sizes versus volume expansion ratio of extruded foams reported in the literature from 2004 to 2015.
514 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

4.2.1. Poly(lactic acid) (PLA) as a base matrix


PLA has been receiving huge attention compared to other thermoplastic polyesters like poly(ethylene terephthalate)
(PET), poly(butylene terephthalate) (PBT), etc. because its green, biodegradable and can be foamed with environmentally
friendly blowing agents such as CO2. Another reason is attributed to the low cost of the resin itself. The chiral nature of
the lactic acid (i.e. the proportion of D and L) influences the crystallization behaviour of PLA and therefore, narrow down
the foam-processing window.
From all the PLA studies examined, it shows that single-line extruders are mainly used for foam investigations because;
PLA has the tendency to undergo thermal degradation and chain scission while in molten state, owing to long residence time
and high processing temperatures. Chemical chain extenders such as 1,4-butane diisocyanate (BDI) are also been used in
many studies to increase the melt-strength and elasticity just like is done with PP.
Larson and Neldin [107] foamed PLA on a single-line extruder using 2% epoxy-functionalized chain extenders in the form
of a masterbatch. The chain extenders were added to enhance the melt-strength and elasticity of the polymer. By metering
8 wt% of CO2 with a fixed die temperature of 125 °C, the authors obtained fine-celled-low-density foams a very high VER of
(42.76) and cell size (22 lm).
Mihai et al. [109] in their study (conducted also on a single-line extruder), incorporated 0.5 wt% of chain extender
(epoxy–styrene–acrylic oligomer) into PLA. They metered almost similar amount (9 wt%) of CO2 and their die temperature
was set at 120 °C. This resulted to fine-celled-low-density foams also with a very high VER of 41.9 and cell size of 100 lm.
One peculiarity found in both studies, is the use of specially designed co-rotating screws. Emphases were placed on the screw
configuration and number of zones on the barrel for specific purposes such as adequate mixing, pressure build-up and
sustenance.

4.2.2. Polypropylene (PP) as a base matrix


PP is a semi-crystalline material generally characterized by poor solubility for CO2 and low melt-strength. Our findings
show that, chemical blowing agents (CBAs) that release CO2 were mostly used to carry out various the foam studies as
opposed to using physical blowing agent (PBA).
Xu [102] studied the foaming behaviour of HMS-PP with SiO2 as NA. They carried out their studies on a single-screw-
single-line extruder by fixing the die temperature at 160 °C. A combination of CBAs: exothermic (Azodicarbonamide) and
an endothermic (EPTcor 882) were used to obtain fine-celled-medium-density foams that have a low VER of 7.5 with a cor-
responding mean cell size of 60 lm.
Chaudhary and Jayaraman [108] in a similar study with PP, incorporated 3 wt% of organoclay as NA and an unstated
amount of maleic-anhydride-grafted-polypropylene (PP-g-MA) as compatibilizer. Using a single-screw extruder with a fixed
die temperature of 145 °C, the authors generated fine-celled-high-density foams with a very low VER of 3 and average cell
size of 87 lm.
Both investigations imply that the temperature of the die (the foam shaping element) is always set some few degrees
higher than the Tm of PP. This is due to the narrow processing window and for this reason, very cold or hot die element,
can trigger early or delayed crystallization, which influences the foaming behaviour to a large extent.

4.2.3. Polystyrene (PS) as a base matrix


PS is an amorphous polymer and foams made out of them are mostly used for thermal insulation applications owing to its
price, ease of processing and wide foam processing window. Unlike PP, PS is not characterized by low melt-strength so there
isn’t a need for chain extenders.
Zhang et al. [16] produced extruded polystyrene foams (XPS) by using CO2 and water as a co-blowing agent. For their
experiment, a throughput of 4 kg/h, a slit die (0.4 mm) and a die temperature of 120 °C were used. The authors observed that
0.5 wt% of graphite (GR) as NA resulted to fine-celled-low-density foams (VER: 37.14) having a mean cell diameter of about
75 lm.
Okolieocha et al. [106] carried out on a tandem foam extrusion line, similar studies on XPS. They used a combination of
CO2 and ethanol as PBA. A slit die (0.5 mm) set at a temperature of 126 °C and a throughput of 4.5 kg/h was used in their
experiment. With as little as 1 wt% of TRGO (thermally reduced graphite oxide), they obtained fine-celled-low-density foams
with a cell size of 25 lm having VER of 17.47.

4.2.4. PE as a base matrix


Lee and Wang [113] foam extruded LDPE nanocomposites with as little as 0.05 wt% layered silicate (Clay) as NA in addi-
tion to 15 wt% of LLDPE-g-maleic anhydride as coupling agent. By metering 8 wt% of CO2, and fixing the die temperature to
95 °C, (a relatively low temperature for foam extrusion), they generated high-density-microcellular foams with a mean cell
size of 5 lm and a very low VER of 3.6.

4.2.5. Copolymers as a base matrix


Costeux and Foether [114] used a single line extruder equipped with a long residence time (RT) cooling device fixed
before the die element to produce the only extruded nanocellular foam available to date in the literature. With MMA-EA
(poly(methyl methacrylate-co-ethyl acrylate)) copolymer filled with 0.25 wt% of silica particles, they were able to meter
20 wt% of CO2 into the melt which was then cooled by the RT cooler in such a way that more CO2 can be dissolved without
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 515

1000 A. Ahmadzai_2014_PS
High density foams Medium density foams F. J. Gomez-Gomez_2013_PETG
M. R. Barzegari_2009_HDPE
C. Tovar-Cisneros_2008_HDPE
G. Kotzev_2007_PP
M. R. Barzegari_2009_LDPE
G. H. Motlagh_2008_COC
Jingli Li_2013_PEI
100 J. Zhou_2014_PLA/PBS

Fine-celled foam
T. Kuboki_2014_PP
A. Ameli_2013_PLA
M. R. Barzegari_2009_HDPE
A. J. N. Spörrer_2007_PP

10
Cell size (µm)

Microcellular foam
1

Nanocellular foam
0.1

0.01
0 1 2 3 4 5 6 7 8 9 10
Volume expansion ratio

Fig. 11. Plot showing the mean cell sizes versus volume expansion ratio of injection moulded foams reported in the literature from 2004 to 2015.

having to freeze the die. Although, they successfully produced nanocellular foams (cell size 389 nm), the very low VER of
3.24 means that they are still high-density foams.
Conclusively, it is obvious to see from the plots (Figs. 9 and 10) that some of the frequently used polymers in batch foam-
ing studies are not been considered in many foam extrusion studies or that studies already conducted are yet to be pub-
lished. A good example is PMMA despite its high affinity for CO2 that could be taken advantage of. Another reason might
be attributed to high cost of material which is for example, about (2.60–3.20) € per kg for PMMA compared to (1.20–1.40) €
per kg for PS (note: the price stated here were estimated based the price survey we conducted).

4.3. Injection moulded foams

Based on our literature survey, it is somewhat difficult to calculate the mean cell size of foams produced by foam injection
moulding. The reason being that foams produced by this process are plagued by a wide cell size distribution (often multi-
modal) running from the edges down to the core.
The lack of clear understanding of the underlying mechanisms of cell nucleation and growth during mould filling and
after the application of packing pressure in both low-pressure and high-pressure foam injection moulding processes, result
in the aforesaid problems.
Furthermore, it appears that there are very less articles in the literature in comparison to studies on foam extrusion and
batch-foaming. A possible reason might be the high cost machinery and the fact that the process is best suited to fabricate
usually small and very complicated parts for industrial purposes, where slight reduction of weight is crucial. Fig. 11 shows a
plot of the average cell sizes versus VER of foams fabricated over the last decade [116–123] via foam injection moulding.
Li et al. [122] used supercritical nitrogen (ScN2) to foam PEI reinforced with 1 wt% of MWCNTs. By uniformly dispersing
the MWCNTs in the polymer, they obtained fine-celled-high-density foams with an average cell size of 16.7 lm and a VER of
1.03. These foams have the lowest cell size but also the least VER judging from the plot in Fig. 11.
Zhou et al. [123] in their study, investigated the influence of organophilic montmorillonite (OMMT) on the cell morphol-
ogy of PLA/PBS composite blends. With 5 wt% of Azodicarbonamide (AC) as chemical blowing agent (CBA) and 3 wt% of
OMMT, they produced high-density-macrocellular foams having an average cell size of 261 lm and VER of 2.98.

5. Conclusions and outlook

Due to the benefits and multi-functionality of polymer foams in many facets of lives, there is a paradigm shift from micro-
cellular to nanocellular foam fabrication especially in construction, automobile, packaging and wind industries. But, the bulk
516 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

of investigations carried out in these directions have focused more on cell size reduction at the expense of foam density,
which beyond question, is one of the most important foam properties.
Furthermore, in order to advance the foam processing field, the following should be stated in any article published in the
future:

 The bulk density of the pure polymer material or unfoamed polymer.


 The foam mean cell size with the corresponding absolute density or volume expansion ratio or the relative density which
are missing in some of the published articles.
 The amount of blowing agent (BA) absorbed by the samples at the chosen saturated temperature, the diffusion coefficient
of the gas and the pressure drop rates observed in foaming.
 Most importantly and if possible, the viscosity of the gas-loaded melt at a certain BA content.

And based on our findings, we make the following conclusions:

 There are to date, no low-density-nanocellular closed-cell foams either made by any of the three foam processing meth-
ods covered by this article. Rather, progress has been made at best in fabricating medium-density-nanocellular foams
using copolymers of PMMA-co-EMA [86,94] via batch foaming and high-density-nanocellular foams from MMA-EA
(poly(methyl methacrylate-co-ethyl acrylate)) copolymer [114] via foam extrusion.
 By classification, there are also to date no microcellular foams made by foam injection moulding.
 CO2 (a physical blowing agent) appears to be the sole candidate for making foams via solid-state foaming and foam extru-
sion while, chemical blowing agents (CBAs) are frequently used in foam injection moulding. The use of physical blowing
agents may be the way out to obtain nanofoams as it has well been demonstrated using the other two technologies.
 It has also been well established that nanocellular foams can be made by using polymers (homo and copolymers and
blends) with high CO2 absorption capacities (i.e. polymers that can absorb above 20 wt% of CO2 e.g. PMMA, PC, MMA,
MAM, EA, etc.) in the case of batch foaming and by extruding at relatively low temperature below 95 °C (cold extrusion).
Most importantly, these feat were accomplished without the use of any nucleating agent.
 It took over 20 years (1980–2000) to move completely from macrocellular to microcellular foams and a period of 15 years
(2001–2015) to move to nanocellular foams.
 Foam injection moulding appears to be the sole business of companies; therefore more studies are needed in this field to
help broaden the horizon of knowledge available in the public domain or can be advanced by other investigators inter-
ested in this field.
 Polymers with Tg > 130 °C usually yield foams with smaller cell sizes while polymer with Tg below the stated value are
able to yield foams with low densities.
 However, the tendency to produce closed-cell nanocellular low-density foams (VER P 50) using the aforementioned
technologies is conceivable in the nearest future through the development of new materials (NAs, copolymers, blends,
combination of BAs) and supplementary advanced processing techniques.

Words used interchangeably

Polymer/plastic/polymer matrix/base polymer.


Bubble/nucleus/cell.
Batch foaming/Batch-wise process.
Additive/filler/particle/heterogeneous nucleating agent.

Conflict of interest

None declared.

Acknowledgements

This research received no specific grant from any finding agency in the public, commercial or non-profit sectors. The
authors would like to thank Philipp Beutler, Sebastian Erlwein and Kevin Fehling for assisting in the collection of data.

References

[1] A.B. Strong, Plastics: Materials and Processing, Pearson Prentice Hall, New Jersey, 2005.
[2] J.E. Martini, F.A. Waldman, N.P. Suh, The Production and Analysis of Microcellular Thermoplastic Foams, vol. 43, SPE ANTEC Tech. Pap., 1982, pp. 674–
676.
[3] L. Sorrentino, M. Aurilia, S. Iannace, Polymeric foams from high performance thermoplastics, Adv. Polym. Technol. 30 (2011) 234–243.
[4] R.D. Roberts, J.C. Kwok, Styrene-maleic anhydride copolymer foam for heat resistant packaging, J. Cell. Plast. 43 (2007) 135–143.
[5] S. Subramonian, P. Filiccia, J. Alcott, Novel soft touch, low abrasion, fine cell polyolefin foams, J. Cell. Plast. 43 (2007) 331–343.
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 517

[6] H. Singh, A.K. Jain, Ignition, combustion, toxicity, and fire retardancy of polyurethane foams: a comprehensive review, J. Appl. Polym. Sci. 111 (2009)
1115–1143.
[7] S. Liu, J. Duvigneau, G.J. Vancso, Nanocellular polymer foams as promising high performance thermal insulation materials, Eur. Polym. J. 65 (2015) 33–
45.
[8] P. Ferkl, R. Pokorný, M. Bobák, J. Kosek, Heat transfer in one-dimensional micro- and nano-cellular foams, Chem. Eng. Sci. 97 (2013) 50–58.
[9] B. Notario, J. Pinto, E. Solorzano, J.a. de Saja, M. Dumon, M.a. Rodríguez-Pérez, Experimental validation of the Knudsen effect in nanocellular polymeric
foams, Polymer 56 (2014) 57–67.
[10] Y. Kim, C.B. Park, P. Chen, R.B. Thompson, Origins of the failure of classical nucleation theory for nanocellular polymer foams, Soft Matter 7 (2011)
7351–7358.
[11] P.U. Jung, Development of Innovative Gas-Assisted Foam Injection Molding Technology, 2013, pp. 1–196.
[12] D.W. Oxtoby, Nucleation of first-order phase transitions, Acc. Chem. Res. 31 (1998) 91–97.
[13] X. Xu, D.E. Cristancho, S. Costeux, Z.-G. Wang, Bubble nucleation in polymer–CO2 mixtures, Soft Matter 9 (2013) 9675–9683.
[14] S. Costeux, I. Khan, S.P. Bunker, H.K. Jeon, Experimental study and modeling of nanofoams formation from single phase acrylic copolymers, J. Cell.
Plast. (2014).
[15] C.B. Park, L.K. Cheung, S.W. Song, The effect of talc on cell nucleation in extrusion foam processing of polypropylene with CO2 and isopentane, Cell.
Polym. 17 (1998) 221–251.
[16] C. Zhang, B. Zhu, L.J. Lee, Extrusion foaming of polystyrene/carbon particles using carbon dioxide and water as co-blowing agents, Polymer 52 (2011)
1847–1855.
[17] A.N. Paquet, K.W. Suh, Carbon Black-Containing Bimodal Foam Structures and Process for Making, 5,210,105, 1993.
[18] M. Mihai, M.A. Huneault, B.D. Favis, Foaming of polystyrene/thermoplastic starch blends, J. Cell. Plast. 43 (2007) 215–236.
[19] C. Zhang, B. Zhu, D. Li, L.J. Lee, Extruded polystyrene foams with bimodal cell morphology, Polymer 53 (2012) 2435–2442.
[20] H. Weber, Grave I. De, E. Röhrl, Foamed plastics, Ulmann’s Encycl. 15 (2012) 540–570.
[21] B. Patel, Y. Delaviz, R. Breindel, Mitchell Z. Weekly, R. Loh, Manoj K. Choudhary, Cell Size Enlargers for Polystyrene Foams, WO 2007/149538 A1, 2006.
[22] M. Kurte-Jardin, Method and Device for Producing Thermoplastics Foam Panels, Panel Manufactured Using Same Method and Device and Panel Use,
EP 2 397 303 A1, 2011.
[23] K.W. Suh, C.P. Park, M.J. Maurer, M.H. Tusim, R. De Genova, R. Broos, et al, Lightweight cellular plastics, Adv. Mater. 12 (2000) 1779–1789.
[24] Z. Guo, J. Yang, M.J. Wingert, D.L. Tomasko, L.J. Lee, T. Daniel, Comparison of carbon nanofibers and activated carbon on carbon dioxide foaming of
polystyrene, J. Cell. Plast. 44 (2008) 453–468.
[25] D. Scherzer, K. Hahn, G. Alicke, G. Turznik, F.-J. Dietzen, Foam Panel having a Reduced Thermal Conductivity, EP000778309B1, 2002.
[26] R.M. Breindel, R.R. Loh, J.P. Rynd, Y. Delaviz, M.E. Polasky, Thermoplastic Foams and Method of Forming them Using Nano-Graphite, US 2011/0064938
A1, 2011.
[27] J. Gordon-Duffy, Infrared Attenuated Polymeric Foam Insulation with Flame Retardant Performance, US 2012/0074347 A1, 2012.
[28] Y. Delaviz, R.M. Breindel, M.Z. Weekly, Extruded Polystyrene Foam Containing Propylene Carbonate, Ethylene Carbonate or Butylene Carbonate as a
Process Aids, US 2011/0144221 A1, 2011.
[29] C.V. Vo, R.T. Fox, Assessment of hydrofluoropropenes as insulating blowing agents for extruded polystyrene foams, J. Cell. Plast. 49 (2013) 423–438.
[30] C. Jacobs, S.K. Dey, Inert gases as alternative blowing agents for extruded low-density polystyrene foam, J. Cell. Plast. 31 (1995) 38–47.
[31] L.M. Miller, T.E. Cisar, R.M. Breindel, M.Z. Weekley, Process for Producing Extruded Foam Products having Polystyrene Blends with High Levels of CO2
as Blowing Agent, US 6,268,046 B1, 2001.
[32] L.E. Daigneault, R. Gendron, Blends of CO2 and 2-ethyl hexanol as replacement foaming agents for extruded polystyrene, J. Cell. Plast. 37 (2001) 262–
272.
[33] K.-M. Lee, E.K. Lee, S.G. Kim, C.B. Park, H.E. Naguib, Bi-cellular foam structure of polystyrene from extrusion foaming process, J. Cell. Plast. 45 (2009)
539–553.
[34] M. Dondero, J. Carella, J. Borrajo, (CO2 + 2-propanol) mixture as a foaming agent for polystyrene: a simple thermodynamic model for the high VLE-
phase diagrams taking into account foam vitrification, J. Appl. Polym. Sci. 104 (2007) 2663–2671.
[35] C.V. Vo, A.N. Paquet, An evaluation of the thermal conductivity of extruded polystyrene foam blown with HFC-134a or HCFC-142b, J. Cell. Plast. 40
(2004) 205–228.
[36] R. Gendron, Foaming polystyrene with a mixture of CO2 and ethanol, J. Cell. Plast. 42 (2006) 127–138.
[37] V. Kan, J.J. Hubert, Process for Producing a Thermoplastic Foam Using Water and Ether, EP000937741B1, 2004.
[38] Y.P. Handa, Expanded and Extruded Biodegradable and Reduced Emission Foams Made with Methyl Formate-based Blowing Agent, WO 2008/076755
A1, 2008.
[39] Y.P. Handa, G.A. Francis, G.C. Castner, M. Zafar, Reduced-VOC and Non-VOC Blowing Agents for Making Expanded and Extruded Thermoplastic Foams,
US 2013/0011648, 2013.
[40] T. Hayashi, O. Kobayashi, J. Fukuzawa, H. Fujiwara, Extruded Styrene Resin Foams and Methods for Producing the Same, US 6,696,504 B1, 2004.
[41] L.M. Michael, R.M. Breindel, M.Z. Weekley, T.E. Cisar, K.J. Prince, Extruded Foam Product with 134A and Alcohol Blowing Agent, US 6,350,789 B1, 2002.
[42] L. Chen, K. Blizard, R. Straff, X. Wang, Effect of filler size on cell nucleation during foaming process, J. Cell. Plast. 38 (2002) 139–148.
[43] N.S. Ramesh, S.T. Lee, Do nanoparticles really assist in nucleation of fine cells in polyolefin foams?, Cell Polym. 24 (2005) 269–277.
[44] K. Goren, L. Chen, L.S. Schadler, R. Ozisik, Influence of nanoparticle surface chemistry and size on supercritical carbon dioxide processed
nanocomposite foam morphology, J. Supercrit. Fluids 51 (2010) 420–427.
[45] N.H. Fletcher, Size effect in heterogeneous nucleation, J. Chem. Phys. 29 (1958) 572–576.
[46] L. Chen, R. Ozisik, L.S. Schadler, The influence of carbon nanotube aspect ratio on the foam morphology of MWNT/PMMA nanocomposite foams,
Polymer 51 (2010) 2368–2375.
[47] S.N. Leung, C.B. Park, H. Li, Effects of nucleating agents’ shapes and interfacial properties on cell nucleation, J. Cell. Plast. 46 (2010) 441–460.
[48] R.B. McClurg, Design criteria for ideal foam nucleating agents, Chem. Eng. Sci. 59 (2004) 5779–5786.
[49] S.K. Goel, E.J. Beckman, Generation of microcellular polymeric foams using supercritical carbon dioxide. I: effect of pressure and temperature on
nucleation, Polym. Eng. Sci. 34 (1994) 1137–1147.
[50] M. Nelson, R. Borkman, Ab initio calculations on CO2 binding to carbonyl groups, J. Phys. Chem. A 102 (1998) 7860–7863.
[51] J. Pinto, J.A. Reglero-Ruiz, M. Dumon, M.A. Rodriguez-Perez, Temperature influence and CO2 transport in foaming processes of poly(methyl
methacrylate)-block copolymer nanocellular and microcellular foams, J. Supercrit. Fluids 94 (2014) 198–205.
[52] O. Muth, T. Hirth, H. Vogel, Investigation of sorption and diffusion of supercritical carbon dioxide into poly(vinyl chloride), J. Supercrit. Fluids 19
(2001) 299–306.
[53] D. Hu, J. Chen, S. Sun, T. Liu, L. Zhao, Solubility and diffusivity of CO2 in isotactic polypropylene/nanomontmorillonite composites in melt and solid
states, Ind. Eng. Chem. Res. 53 (2014) 2673–2683.
[54] X. Xu, C.B. Park, D. Xu, R. Pop-Iliev, Effects of die geometry on cell nucleation of PS foams blown with CO2, Polym. Eng. Sci. 43 (2003) 1378–1390.
[55] V. Kumar, J. Weller, Production of microcellular polycarbonate using carbon dioxide for bubble nucleation, J. Eng. Ind. 116 (1994) 413–420.
[56] J. Pinto, E. Solorzano, M.A. Rodriguez-Perez, J.A. de Saja, Characterization of the cellular structure based on user-interactive image analysis procedures,
J. Cell. Plast. 49 (2013) 555–575.
[57] P. Liu, D. Liu, H. Zou, P. Fan, W. Xu, Structure and properties of closed-cell foam prepared from irradiation crosslinked silicone rubber, J. Appl. Polym.
Sci. 113 (2009) 3590–3595.
518 C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519

[58] C. Saiz-Arroyo, M.a. Rodríguez-Pérez, J. Tirado, A. López-Gil, J.A. de Saja, Structure–property relationships of medium-density polypropylene foams,
Polym. Int. 62 (2013) 1324–1333.
[59] S. Costeux, CO2-blown nanocellular foams, J. Appl. Polym. Sci. 131 (2014) 41293.
[60] S.P. Nalawade, F. Picchioni, L.P.B.M. Janssen, Supercritical carbon dioxide as a green solvent for processing polymer melts: processing aspects and
applications, Prog. Polym. Sci. 31 (2006) 19–43.
[61] M. Sauceau, J. Fages, A. Common, C. Nikitine, E. Rodier, New challenges in polymer foaming: a review of extrusion processes assisted by supercritical
carbon dioxide, Prog. Polym. Sci. 36 (2011) 749–766.
[62] D.L. Tomasko, A. Burley, L. Feng, S.-K. Yeh, K. Miyazono, S. Nirmal-Kumar, et al, Development of CO2 for polymer foam applications, J. Supercrit. Fluids
47 (2009) 493–499.
[63] H. Guanghong, W. Yue, Microcellular Foam Injection Molding Process, Some Critical Issues for Injection Molding, Intech, 2012.
[64] G. Zhao, G. Wang, J. Wang, L. Zhang, Foaming mechanism and control in microcellular injection molding, SPE Plast. Res. (2014), http://dx.doi.org/
10.2417/spepro.005571.
[65] Z. Ma, G. Zhang, Q. Yang, X. Shi, A. Shi, Fabrication of microcellular polycarbonate foams with unimodal or bimodal cell-size distributions using
supercritical carbon dioxide as a blowing agent, J. Cell. Plast. 50 (2013) 55–79.
[66] J.A. Reglero, P. Viot, M. Dumon, Foaming of amorphous polymers and blends in supercritical CO2: solubility versus block copolymers addition, J. Cell.
Plast. 47 (2011) 535–548.
[67] M.P. Tran, C. Detrembleur, M. Alexandre, C. Jerome, J.M. Thomassin, The influence of foam morphology of multi-walled carbon nanotubes/poly
(methyl methacrylate) nanocomposites on electrical conductivity, Polymer 54 (2013) 3261–3270.
[68] J.-B. Bao, T. Liu, L. Zhao, G.-H. Hu, X. Miao, X. Li, Oriented foaming of polystyrene with supercritical carbon dioxide for toughening, Polymer 53 (2012)
5982–5993.
[69] L. Chen, L.S. Schadler, R. Ozisik, An experimental and theoretical investigation of the compressive properties of multi-walled carbon nanotube/poly
(methyl methacrylate) nanocomposite foams, Polymer 52 (2011) 2899–2909.
[70] D. Miller, P. Chatchaisucha, V. Kumar, Microcellular and nanocellular solid-state polyetherimide (PEI) foams using sub-critical carbon dioxide I.
Processing and structure, Polymer 50 (2009) 5576–5584.
[71] J.E. Weller, V. Kumar, Solid-state microcellular polycarbonate foams. II. The effect of cell size on tensile properties, Polym. Eng. Sci. 50 (2010) 2170–
2175.
[72] H. Janani, M.H.N. Famili, Investigation of a strategy for well controlled inducement of microcellular and nanocellular morphologies in polymers,
Polym. Eng. Sci. 50 (2010) 1558–1570.
[73] C. Li, G. Yang, H. Deng, K. Wang, Q. Zhang, F. Chen, et al, The preparation and properties of polystyrene/functionalized graphene nanocomposite foams
using supercritical carbon dioxide, Polym. Int. 62 (2013) 1077–1084.
[74] J.A. Reglero Ruiz, C. Saiz-Arroyo, M. Dumon, M.A. Rodríguez-Perez, L. Gonzalez, Production, cellular structure and thermal conductivity of
microcellular (methyl methacrylate)-(butyl acrylate)-(methyl methacrylate) triblock copolymers, Polym. Int. 60 (2011) 146–152.
[75] Y.-W. Chang, D. Lee, S.-Y. Bae, Preparation of polyethylene-octene elastomer/clay nanocomposite and microcellular foam processed in supercritical
carbon dioxide, Polym. Int. 55 (2006) 184–189.
[76] K. Lee, Y.-W. Chang, S.W. Kim, Ethylene-propylene-diene terpolymer/halloysite nanocomposites: thermal, mechanical properties, and foam
processing, J. Appl. Polym. Sci. 131 (2014).
[77] J. Yang, L. Huang, Y. Zhang, F. Chen, M. Zhong, Mesoporous silica particles grafted with polystyrene brushes as a nucleation agent for polystyrene
supercritical carbon dioxide foaming, J. Appl. Polym. Sci. 130 (2013) 4308–4317.
[78] J. Antonio, R. Ruiz, E. Cloutet, M. Dumon, Investigation of the nanocellular foaming of polystyrene in supercritical CO2 by adding a CO2 philic
perfluorinated block copolymer, J. Appl. Polym. Sci. 126 (2012) 38–45.
[79] W. Zhai, S.N. Leung, L. Wang, H.E. Naguib, C.B. Park, Preparation of microcellular poly(ethylene-co-octene) rubber foam with supercritical carbon
dioxide, J. Appl. Polym. Sci. (2010).
[80] J.A.R. Ruiz, P. Viot, M. Dumon, Microcellular foaming of polymethylmethacrylate in a batch supercritical CO2 process: effect of microstructure on
compression behavior, J. Appl. Polym. Sci. 118 (2010) 320–331.
[81] J.A.R. Ruiz, J. Marc-Tallon, M. Pedros, M. Dumon, Two-step micro cellular foaming of amorphous polymers in supercritical CO2, J. Supercrit. Fluids 57
(2011) 87–94.
[82] Y.-M. Corre, A. Maazouz, J. Duchet, J. Reignier, Batch foaming of chain extended PLA with supercritical CO2: influence of the rheological properties and
the process parameters on the cellular structure, J. Supercrit. Fluids 58 (2011) 177–188.
[83] H. Zhang, Q. Yan, W. Zheng, Z. He, Z. Yu, Tough graphene–polymer microcellular foams for electromagnetic interference shielding, Appl. Mater.
Interfaces 3 (2011) 918–924.
[84] C. Okolieocha, F. Beckert, M. Herling, J. Breu, R. Mülhaupt, V. Altstädt, Preparation of microcellular low-density PMMA nanocomposite foams:
influence of different fillers on the mechanical, rheological and cell morphological properties, Compos. Sci. Technol. 118 (2015) 108–116.
[85] S.-K. Yeh, Y.-C. Liu, W.-Z. Wu, K.-C. Chang, W.-J. Guo, S.-F. Wang, Thermoplastic polyurethane/clay nanocomposite foam made by batch foaming, J.
Cell. Plast. 49 (2013) 119–130.
[86] S. Costeux, L. Zhu, Low density thermoplastic nanofoams nucleated by nanoparticles, Polymer 54 (2013) 2785–2795.
[87] H. Guo, V. Kumar, Some thermodynamic and kinetic low-temperature properties of the PC-CO2 system and morphological characteristics of solid-
state PC nanofoams produced with liquid CO2, Polymer 56 (2015) 46–56.
[88] H. Ruckdäschel, Micro- and Nanostructured Polymer Blends – Processing, Properties and Foaming Behaviour, 2008, pp. 242–270.
[89] S. Costeux, S.P. Bunker, Homogenous nanocellular foams from styrenic-acrylic polymer blends, J. Mater. Res. 28 (2013) 2351–2365.
[90] M.M. Khorasani, S.R. Ghaffarian, A. Babaie, N. Mohammadi, Foaming behavior and cellular structure of microcellular HDPE nanocomposites prepared
by a high temperature process, J. Cell. Plast. 46 (2010) 173–190.
[91] Y.H. Lee, HDPE-clay nanocomposite foams blown with supercritical CO2, J. Cell. Plast. 41 (2005) 487–502.
[92] N.J. Hossieny, M.R. Barzegari, M. Nofar, S.H. Mahmood, C.B. Park, Crystallization of hard segment domains with the presence of butane for
microcellular thermoplastic polyurethane foams, Polymer 55 (2014) 651–662.
[93] Y. Ito, M. Yamashita, M. Okamoto, Foam processing and cellular structure of polycarbonate-based nanocomposites, Macromol. Mater. Eng. 291 (2006)
773–783.
[94] S. Costeux, I. Khan, S.P. Bunker, H.K. Jeon, Experimental study and modeling of nanofoams formation from single phase acrylic copolymers, J. Cell.
Plast. 51 (2015) 197–221.
[95] K. Satish, J. Eric, S.K. Goel, E.J. Beckman, Generation of microcellular polymeric foams using supercritical carbon dioxide. II: cell growth and skin
formation, Polym. Eng. Sci. 34 (1994) 1148–1156.
[96] D. Szegda, S. Duangphet, J. Song, K. Tarverdi, Extrusion foaming of PHBV, J. Cell. Plast. 50 (2014) 145–162.
[97] W. Liu, X. Wang, H. Li, Z. Du, C. Zhang, Study on rheological and extrusion foaming behaviors of chain-extended poly(lactic acid)/clay nanocomposites,
J. Cell. Plast. 49 (2013) 535–554.
[98] W. Michaeli, K. Westermann, S. Sitz, Extrusion of physically foamed rubber profiles, J. Cell. Plast. 47 (2011) 483–495.
[99] G.M. Rizvi, C.B. Park, G. Guo, Strategies for processing wood plastic composites with chemical blowing agents, J. Cell. Plast. 44 (2008) 125–137.
[100] W. Kaewmesri, Effects of CO2 and talc contents on foaming behavior of recyclable high-melt-strength PP, J. Cell. Plast. 42 (2006) 405–428.
[101] E.H. Tejeda, Morphology and mechanical properties of foamed polyethylene–polypropylene blends, J. Cell. Plast. 41 (2005) 417–435.
[102] Z. Xu, Effects of formulations and processing parameters on foam morphologies in the direct extrusion foaming of polypropylene using a single-screw
extruder, J. Cell. Plast. 41 (2005) 169–185.
C. Okolieocha et al. / European Polymer Journal 73 (2015) 500–519 519

[103] A.P. Ranade, Structure–property relationships in coextruded foam/film microlayers, J. Cell. Plast. 40 (2004) 497–507.
[104] A. Rizvi, A. Tabatabaei, M.R. Barzegari, S.H. Mahmood, C.B. Park, In situ fibrillation of CO2-philic polymers: sustainable route to polymer foams in a
continuous process, Polymer 54 (2013) 4645–4652.
[105] J.-M. Julien, J.-C. Bénézet, E. Lafranche, J.-C. Quantin, A. Bergeret, M.-F. Lacrampe, et al, Development of poly(lactic acid) cellular materials: physical
and morphological characterizations, Polymer 53 (2012) 5885–5895.
[106] C. Okolieocha, T. Koppl, S. Kerling, F.J. Tolle, A. Fathi, R. Mülhaupt, V. Altstädt, Influence of graphene on the cell morphology and mechanical properties
of extruded polystyrene foam, J. Cell. Plast. 51 (2015) 413–426.
[107] Å. Larsen, C. Neldin, Physical extruder foaming of poly(lactic acid)-processing and foam properties, Polym. Eng. Sci. 53 (2013) 941–949.
[108] A.K. Chaudhary, K. Jayaraman, Extrusion of linear polypropylene-clay nanocomposite foams, Polym. Eng. Sci. 51 (2011) 1749–1756.
[109] M. Mihai, M.A. Huneault, B.D. Favis, Rheology and extrusion foaming of chain-branched poly(lactic acid), Polym. Eng. Sci. 50 (2010) 629–642.
[110] S.-K. Yeh, J. Yang, N.-R. Chiou, T. Daniel, L.J. Lee, Introducing water as a coblowing agent in the carbon dioxide extrusion foaming process for
polystyrene thermal insulation foams, Polym. Eng. Sci. 50 (2010) 1577–1584.
[111] P. Spitael, C.W. Macosko, Strain hardening in polypropylenes and its role in extrusion foaming, Polym. Eng. Sci. 44 (2004) 2090–2100.
[112] A. Ayoub, S.S.H. Rizvi, Reactive supercritical fluid extrusion for development of moisture resistant starch-based foams, J. Appl. Polym. Sci. 120 (2011)
2242–2250.
[113] Y. Lee, K. Wang, Effects of clay dispersion on the foam morphology of LDPE/clay nanocomposites, J. Appl. Polym. Sci. 103 (2007) 2129–2134.
[114] S. Costeux, D. Foether, Continuous extrusion of nanocellular foam, SPE ANTEC Tech. Pap. 100 (2015) 1–6.
[115] Z.X. Xin, Z.X. Zhang, K. Pal, K.-J. Kim, D.J. Kang, J.K. Kim, et al, Microcellular structure of PP/waste rubber powder blends with supercritical CO2 by foam
extrusion process, J. Cell. Plast. 45 (2009) 499–514.
[116] F.J. Gomez-Gomez, D. Arencon, M.a. Sanchez-Soto, a.B. Martinez, Influence of the injection moulding parameters on the microstructure and thermal
properties of microcellular polyethylene terephthalate glycol foams, J. Cell. Plast. 49 (2012) 47–63.
[117] M. Reza Barzegari, D. Rodrigue, Prediction of the shear modulus of polymer structural foams, J. Cell. Plast. 45 (2009) 555–576.
[118] C. Tovar-Cisneros, R. Gonzalez-Nunez, D. Rodrigue, Effect of mold temperature on morphology and mechanical properties of injection molded HDPE
structural foams, J. Cell. Plast. 44 (2008) 223–237.
[119] G. Kotzev, S. Djoumaliisky, M. Krasteva, M. Iliev, E. Pérez, M.L. Cerrada, Effect of sample configuration on the morphology of foamed LDPE/PP blends
injection molded by a gas counterpressure process, Macromol. Mater. Eng. 292 (2007) 769–779.
[120] M.R. Barzegari, D. Rodrigue, The effect of injection molding conditions on the morphology of polymer structural foams, Polym. Eng. Sci. 49 (2009)
949–959.
[121] G.H. Motlagh, a.N. Hrymak, M.R. Thompson, Improved through-plane electrical conductivity in a carbon-filled thermoplastic via foaming, Polym. Eng.
Sci. 48 (2008) 687–696.
[122] J. Li, Z. Chen, X. Wang, T. Liu, Y. Zhou, S. Luo, Cell morphology and mechanical properties of microcellular mucell V injection molded polyetherimide
and polyetherimide/fillers composite foams, J. Appl. Polym. Sci. (2013) 4171–4181.
[123] J. Zhou, Z. Yao, C. Zhou, D. Wei, S. Li, Mechanical properties of PLA/PBS foamed composites reinforced by organophilic montmorillonite, J. Appl. Polym.
Sci. 131 (2014) 40773.

You might also like