Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

Accepted Manuscript

Title: Poly (lactic acid) Foaming: A Review

Author: Mohammadreza Nofar Chul B. Park

PII: S0079-6700(14)00039-2
DOI: http://dx.doi.org/doi:10.1016/j.progpolymsci.2014.04.001
Reference: JPPS 869

To appear in: Progress in Polymer Science

Received date: 10-10-2013


Revised date: 17-3-2014
Accepted date: 25-3-2014

Please cite this article as: Nofar M, Park CB, Poly (lactic
acid) Foaming: A Review, Progress in Polymer Science (2014),
http://dx.doi.org/10.1016/j.progpolymsci.2014.04.001

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Poly (lactic acid) Foaming: A Review

Mohammadreza Nofar and Chul B. Park*

t
ip
Microcellular Plastics Manufacturing Laboratory, Department of Mechanical and Industrial

cr
Engineering, University of Toronto, 5 King’s College Road, Toronto, Ontario, Canada M5S 3G8

us
* Corresponding Author: park@mie.utoronto.ca 

Abstract an
Poly (lactic acid) or polylactide (PLA) is an aliphatic thermoplastic polyester produced
M
from renewable resources and is compostable in the environment. Because of the massive use of
foamed products of petroleum-based polymers, PLA foams have been considered as substitutes
for some of these products. Specifically, because of PLA’s competitive material and processing
d

costs, and its comparable mechanical properties, PLA foams could potentially replace
e

polystyrene (PS) foam products in a wide array of applications such as packaging, cushioning,
pt

construction, thermal and sound insulation, and plastic utensils. Due to their biocompatibility,
PLA foams can also be used in such biomedical applications as scaffolding and tissue
ce

engineering. But PLA has several inherent drawbacks, which inhibit the production of low-
density foams with uniform cell morphology. These drawbacks are mainly the PLA’s low melt
strength and its slow crystallization kinetics. During the last two decades, researchers have
Ac

investigated the fundamentals of PLA/gas mixtures, PLA foaming mechanisms, and the effects
of material modification on PLA’s foaming behavior through various manufacturing
technologies. This article reviews these investigations and compares the developments made thus
far in PLA foaming.

Keywords: Poly (lactic acid), polylactide, PLA, blowing agent, foaming, review


 

Page 1 of 63
Table of Contents
1- Introduction ................................................................................................................................  
2- Fundamental Studies of PLA/Gas Mixtures ............................................................................  
2-1- Solubility, Diffusivity, Pressure-Volume-Temperature (PVT), and Surface Tension........... 
2-2- Crystallization Behaviors ....................................................................................................... 

t
ip
3- Foaming Mechanisms of PLA ...................................................................................................  
3-1- Effects of Crystallization........................................................................................................ 
3-2- Effects of Molecular Architecture and Configuration............................................................ 

cr
3-2-1- Effects of Chain Branching..........................................................................................................  
3-2-2- Effects of D-Lactide Content .......................................................................................................  

us
3-3- PLA Compounds .................................................................................................................... 
3-3-1- PLA Blends ..................................................................................................................................  
3-3-2- PLA Composites ..........................................................................................................................  
an
3-3-3- PLA Nanocomposites ..................................................................................................................  
4- PLA Foam Processing Technologies .........................................................................................  
4-1- PLA Extrusion Foaming......................................................................................................... 
M
4-2- PLA Foam Injection Molding ................................................................................................ 
4-3- PLA Bead Foaming................................................................................................................ 
5- Conclusions..................................................................................................................................  
Acknowledgments ...........................................................................................................................  
d

References
e
pt
ce
Ac


 

Page 2 of 63
1- Introduction
At present, most polymeric products are derived from fossil fuels, and they become non-

t
ip
degradable waste materials in the environment. Consequently, global efforts are being made to

create green polymers from renewable resources which are also biodegradable and compostable.

cr
Poly (lactide acid) or polylactide (PLA) is a biodegradable and biocompatible polymer produced

us
from such renewable resources as cornstarch and sugarcane [1-5]. It is a thermoplastic aliphatic

polyester that is synthesized through the ring-opening polymerization of lactide and lactic acid
an
monomers [6-8], as shown in Figure 1. Over the past decade, PLA has attracted extensive

industrial and academic interest as a potential substitute for petroleum-based polymers in both
M
commodity and biomedical applications. This is not only due to its green and biodegradable

features, but also because it releases no toxic components during the manufacture.
d

Due to its competitive material and processing costs, and comparable mechanical
e

properties, this environmentally-friendly biopolymer is considered as a promising replacement


pt

for polystyrene (PS), especially for PS foam products in such commodity applications as
ce

packaging, cushioning, construction, thermal and sound insulation, and plastic utensils [9-15]. It

would be environmentally very attractive to replace the commodity PS foam products with PLA
Ac

foams because the amount of landfill for the large volume PS foam waste has been a serious

global concern.

Figure 1.

PLA foaming has mostly been conducted by dissolving a physical blowing agent in the

PLA matrix and blowing it. Cell nucleation and growth occurs through the thermodynamic


 

Page 3 of 63
instability generated from the super-saturation of the blowing agent (i.e., a pressure drop or a

temperature increase). The foam structure is then produced by expelling the dissolved gas from

the PLA/gas mixture. Foam products are made with cell stabilization as the temperature reaches

below the PLA’s Tg, which is around 60oC [15-17].

t
Currently, the mass production of low-density PLA foams with a uniform cell morphology

ip
using supercritical carbon dioxide and nitrogen as the physical blowing agent is still quite

cr
challenging. This is mainly due to PLA’s low melt strength [15, 18-22]. Introducing a chain

us
extender to create a branched structure [16, 22-24]; modifying the L/D ratio configuration of the

PLA molecules [19-20]; varying the PLA’s molecular weight [18-20, 22]; and compounding the
an
PLA with different types of additives [25-34] have been recognized as efficient methods to

improve PLA’s poor melt strength, and consequently, its foamability. Enhancing PLA’s slow
M
crystallization kinetics has also been proven to significantly improve its inherently low melt

strength and to extend its applications [35-36]. The enhanced crystallization compensates for
d

PLA’s low melt strength during processing, and thereby improves its foamability [37-40].
e

Enhanced crystallization can further improve the final product’s mechanical properties and can
pt

compensate for PLA’s low heat deflection temperature (i.e., service temperature) [41-42].
ce

This article reviews the fundamental properties, foaming mechanisms, and processing

technologies for PLA foams. First, we address fundamental studies made of PLA/gas mixtures,
Ac

which are required for the foaming of PLA. We discuss the development of PLA foams through

various mechanisms, and examine the efforts that have been made to achieve low-density

microcellular and nanocellular PLA foams. The article also introduces the manufacturing

technologies that have been applied in PLA foam production.


 

Page 4 of 63
2- Fundamental Studies of PLA/Gas Mixtures
2-1- Solubility, Diffusivity, Pressure-Volume-Temperature (PVT),
and Surface Tension
The determination of solubility of a physical blowing agent in a polymer is important

t
because when gas encounters polymer under a high-pressure condition, it impregnates the

ip
polymer and thereby the polymer starts to swell. The amount of swelling depends on how much

cr
gas dissolves in the polymer. It also depends on the polymer’s molecular configuration and

us
architecture. The polymer’s swelling behavior can be analyzed through a pressure-volume-

temperature (PVT) apparatus (i.e., a dilatometer) to evaluate the polymer/gas mixture’s PVT
an
behavior [43]. The foaming behavior of a polymer (i.e., cell nucleation and growth) is generally

governed by the following thermodynamic properties: solubility [44], diffusivity [45], and
M
surface tension [46]. All of these heavily depend on the PVT properties of the polymer/gas

mixtures [45, 47]. Li et al. [45] and Mahmood et al. [48] investigated the PVT and the solubility
d

behaviors of PLA/gas mixtures by using a PVT apparatus and a magnetic suspension balance
e

(MSB), respectively. Two different equations of state (EOS) were used to predict and compare
pt

the volume swelling that would be caused by gas dissolution in the PLA: the Sanchez-Lacombe
ce

[SL] EOS and the Simha-Somocynsky [SS] EOS, see references [45] and [47] for details.

Li et al. [45] investigated the solubility and diffusivity of CO2 and N2 in PLA melts. The
Ac

supplied PLA was Ingeo 3001D from NatureWorks, with a low D-lactide content (1.4%). On the

other hand, Mahmood et al. [48] only explored the solubility of CO2 using three different PLA

materials with varying D-lactide content as follows: PLA 3001D (1.4% D-content), PLA 8051D

(4.6% D-content), and PLA 4060 D (12% D-content). The SS-corrected PVT behaviors and the

solubility of the PLA samples when exposed to various CO2 and N2 pressures [45, 48] are shown

in Figures 2a-2b and 3a-3b, respectively. The PLA samples with varying D-content are also


 

Page 5 of 63
shown under changing CO2 pressures [48]. With increased pressure and decreased temperature,

the volume swelling ratio and the gas solubility increased in the PLA/CO2 mixtures. As the CO2

pressure increased, more CO2 molecules dissolved in the PLA matrix and increased the volume

of the PLA/CO2 mixtures. In other words, when a polymer was solubilized by being exposed to a

t
pressurized gas, the dissolved gas increased the intermolecular distance and thereby caused a

ip
swelling, despite the hydraulic pressure effect. Consequently, the free volume and the volume of

cr
the polymer/gas mixture both increased. In addition, when the temperature increased, the PLA’s

us
molecular movement further increased. Thus, the free volume and the specific volume increased

[48-49]. But this increase was insufficient to enhance the CO2`s solubility, and thereby the CO2`s
an
solubility in the PLA was reduced. As seen, the solubility results of Li et al. and Mahmood et al.

were slightly different, because they had each used a different Tait’s equation when making the
M
PVT measurements.

Moreover, Li et al. reported a significant difference in the CO2’s swelling and solubility in
d

PLA melts other than those under N2 pressures. The maximum solubility achieved in a PLA melt
e

at 27.8 MPa supercritical CO2 and N2 pressures was reported as 20 wt% and 2 wt%, respectively
pt

[45]. On the other hand, Mahmood et al. observed that the swelling and solubility of CO2 in PLA
ce

melts did not change significantly with varied D-content [48].


Ac

Figure 2.

Figure 3.

Li et al. further investigated the diffusivity behaviors of CO2 and N2 in PLA melts at three

different pressures at 180oC [45]. Figure 4 shows that the N2 had a higher diffusivity than the


 

Page 6 of 63
CO2. However, with increased pressure, the diffusivity decreased due to the increased hydraulic

pressure of the gas, which reduced the free volume of the PLA/gas mixture [45].

Figure 4.

The solubility of a gas in semi-crystalline polymers becomes more complex when

t
ip
crystallization occurs [50]. As the crystallization occurs, the solubility of a gas in a polymer

should decrease. This is because the crystallization will expel the dissolved gas, and therefore gas

cr
can hardly dissolve in the crystalline structure [50]. Using a quartz crystal microbalance at high

us
pressures, Oliveira et al. [51] studied the solubility of CO2 in amorphous and in semi-crystalline

PLA samples. This was done at temperatures below the PLA’s Tg (~60oC) at which an existing
an
crystal’s impact on solubility could be considered. The amorphous and semi-crystalline PLA

grades had a D-lactide content of 20% and 2%, respectively. Figure 5 shows the CO2’s solubility
M
in PLA at various CO2 pressures at 30.9oC. CO2’s solubility in the amorphous PLA was higher

beyond 2 MPa than it was in the semi-crystalline PLA. Oliveira et al. [51] reported that this must
e d

have been due to the presence of crystallites, which reduced the PLA/gas mixture’s swelling and
pt

thereby lessened the dissolved gas content.

Figure 5.
ce

Aionicesei et al. [52] also studied the solubility of supercritical CO2 in poly (L-lactide)
Ac

(PLLA) at temperatures below the PLLA’s Tg by using an MSB. Figure 6 shows that a

temperature decrease increased the CO2’s solubility in PLLA. Other literature also confirmed this

finding [45, 48]. Moreover, a solubility of up to 43 wt% was achieved in PLLA at around 30

MPa CO2 pressure, although the existing crystal could also have negatively affected the CO2’s

solubility.


 

Page 7 of 63
Figure 6.

Cell nucleation during foaming can also be significantly influenced by a polymer’s

interfacial tension. Due to the polymer/gas mixture’s lower interfacial tension, the Gibb’s free

energy for cell nucleation is less than it would be for neat polymer. Therefore, an examination of

t
ip
the interfacial tension of a polymer/gas mixture is crucial. In another study, Mahmood et al. [53]

investigated the interfacial tension of PLA/CO2 mixtures (Figure 7). They showed that the

cr
pressure increase reduced PLA interfacial tension. However, the temperature effect depended on

us
the CO2 pressure level. At lower pressures, the PLA interfacial tension decreased with

temperature while at a higher CO2 pressure the opposite occurred. They explained that although
an
the interfacial tension of a polymer melt decreased with increased temperature, at higher CO2

pressures the influence of dissolved gas dominated the interfacial tension. Thus, PLA interfacial
M
tension was increased by increased temperature. They also showed that the interfacial tension of

PLA samples with varying D-lactide content did not significantly change.
e d

Figure 7.
pt

2-2- Crystallization Behaviors


ce

The dissolved physical blowing agent is one of the parameters that can affect a polymer’s

molecular mobility and, consequently, its crystallization kinetics and glass transition temperature
Ac

(Tg). Some research has demonstrated how dissolved gas can influence PLA’s isothermal [54,

57-62] and non-isothermal crystallization kinetics [55-58, 59-62]. Takada et al. [54], Yu et al.

[57], and Reignier et al. [58] showed that during isothermal crystallization the dissolved CO2 gas

enhanced PLA’s crystallization rate in the crystal-growth controlled region, whereas it reduced

the crystallization rate in the nucleation controlled region. Nofar et al. [59-62] also made an in-


 

Page 8 of 63
depth study of the effects of dissolved CO2 on PLA’s isothermal melt crystallization. They found

that the increased dissolved CO2 content enhanced the PLA’s crystallization rate, regardless of

the region. They also observed that the plasticizing effect of CO2 dissolved gas suppressed the

temperature ranges required for isothermal crystallization in PLA. In other words, at elevated

t
pressures, the corresponding crystal-growth and nucleation controlled regions and the Tcritical, at

ip
which the crystallization rate is the fastest, were depressed to lower temperatures by the

cr
lubricating effect of the CO2 gas molecules [59, 62]. Therefore, at higher CO2 pressures, the

us
crystallization rate was increased in the corresponding crystal-growth and nucleation controlled

regions. Figure 8 shows the crystallization rate variations in relation to different isothermal
an
temperatures in the PLA samples with a D-lactide content of 4.5% and 1.5%. The isothermal

melt crystallization was implemented under atmospheric pressure and at 45 bar of CO2 pressure.
M
Figure 8.
d

The majority of the literature [55-62] confirmed that during the non-isothermal melt
e

crystallization process, the increased CO2 gas pressure (i.e., the dissolved CO2 content) enhanced
pt

the crystallization rate. This was a result of the CO2’s plasticization impact, which reduced the
ce

dissipation energy required for molecular retraction [56, 59-62]. Using the Kissinger method, Li

et al. [56] showed that the activation energy required for the PLA’s non-isothermal melt
Ac

crystallization decreased with increased CO2 pressures (Figure 9a). They demonstrated that the

plasticization effect of dissolved CO2 gas reduced the viscosity of the PLA/gas mixture. This

meant that enhanced molecular mobility would require lower energy dissipation for molecular

retraction and hence, for crystallization, to occur. On the other hand, the PLA/gas mixture’s

reduced interfacial tension would facilitate the creation of a new interface in the amorphous

structure [56]. Yu et al. [55, 57], Li et al. [56], and Nofar et al. [59-62] reported that during the


 

Page 9 of 63
cooling process at a rate of 2oC/min., the crystallization half-time was reduced as the CO2

pressure increased. Figure 9b shows that the crystallization half-time was reduced at increased

CO2 pressures. But depending on the PLA type and its crystallization capability, the half-time

reduction with the CO2 pressure was different.

t
ip
Figure 9.

Some of the studies [54, 57, 59-60] also investigated the crystallization kinetics of the

cr
PLA/CO2 mixtures during isothermal crystallization through Avrami analysis [63], where the

us
Avrami n value exponent reflected the crystallization mechanisms. When the n value is around 3

and above, three-dimensional growth occurs. However, when the n value decreases to 2 and
an
below, crystallization corresponds to two-dimensional spherulitic growth. As Figure 10 shows,

with the exception of Yu et al. [57], other researchers [54, 59-60] have found that an increase in
M
the CO2 pressure decreased the Avrami exponent, n, close to or below 2, suggesting that two-

dimensional spherulitic crystal growth had dominated the crystallization kinetics. This was most
d

likely due to the CO2’s plasticization effect, which could reorient imperfect crystals into a planar
e

structure [60].
pt

Figure 10.
ce

The effects of dissolved gas on the PLA’s final crystallinity were also explored during non-
Ac

isothermal and isothermal crystallization processes [54-57, 59-62]. Takada et al. [54], Li et al.

[56], and Liao et al. [64] showed that dissolved CO2 increased the PLA’s final crystallinity, but

Yu et al. [55, 57] found an opposite trend. Figure 11 shows the final crystallinity results obtained

at various CO2 gas pressures by Li et al. [56]. Overall at different cooling rates, the PLA’s final

crystallinity was seen to increase. This was explained as the dissolved CO2 gas’s plasticization

effect having reduced the viscosity of the PLA/gas mixture. It meant that enhanced PLA

10 
 

Page 10 of 63
molecular mobility and lower dissipation energy would be needed for molecular retraction and

crystallization to occur [65]. Takada et al. [54] also showed that the isothermally treated PLA

samples under high pressure CO2 had a higher final crystallinity, although the maximum

implemented CO2 pressure was only 20 bar. In contrast, Yu et al. [55, 57] showed that the

t
increased CO2 pressure reduced the crystallization heat of fusion during cooling at a rate of

ip
2oC/min. They explained that at high CO2 pressures, crystallization occurred in the nucleation-

cr
controlled region, and that the dissolved CO2 retarded nucleation by dissolving and diluting the

us
crystal nuclei. Therefore, the crystallization was suppressed and the final crystallinity was

decreased.
an
Figure 11.

Nofar et al. [59-62] demonstrated that during isothermal and non-isothermal crystallization
M
the final crystallinity could have different mechanisms under various CO2 pressures. This was so

because at various gas pressures, different crystallization kinetics (i.e., nucleation and growth)
d

govern PLA’s final crystallinity. It was confirmed that, while the CO2’s plasticization effect can
e

increase PLA’s molecular mobility and thereby facilitate crystallization, it is not the sole
pt

parameter affecting it. In fact, both the impact of the plasticizing CO2 molecules and the density
ce

of the crystal nuclei can simultaneously influence PLA’s molecular mobility at various pressures.

Consequently, depending on both the effect of the number of crystal nuclei and on the
Ac

plasticization of the dissolved CO2 gas at various pressures, the crystal growth rate, and

subsequently the final crystallinity, can be different.

Figure 12 shows the final crystallinity of a wide range of PLA materials versus CO2

pressures during non-isothermal melt crystallization at a cooling rate of 2oC/min. The PLA

materials were as follows: linear PLAs with varying D-lactide content (i.e., D: 1.5%, 4.5%, and

10%), short and long chain branched PLAs (i.e., SCB and LCB), and PLA micro/nanocomposites,
11 
 

Page 11 of 63
with 1 wt% of talc, nanosilica, and nanoclay additives. Nofar et al. confirmed that during

isothermal and low-cooling-rate non-isothermal crystallization, the PLA’s final crystallinity

increased at low CO2 pressures, and that the crystals had a larger and more perfect structure.

However, at higher CO2 pressures, the final crystallinity decreased, and there were a larger

t
number of small-size imperfect crystals [59-62]. They demonstrated that at low CO2 pressures,

ip
the plasticizing CO2 molecules facilitated the molecular mobility of the PLA matrix. Thus, the

cr
crystal growth rate of a limited number of nucleated crystals was promoted. Consequently, the

us
final crystallinity was increased and had a more perfect crystalline structure. On the other hand,

at higher CO2 pressures, where the plasticization effect of the gas molecules improved the PLA’s
an
molecular movement, the enhanced crystal nucleation rate suppressed molecular mobility.

Eventually, crystal growth was retarded by the dominant molecular entanglement that was caused
M
by the network of a large number of nucleated crystals. Thus, the final crystallinity was reduced

with less perfect crystalline domain [59-62].


d

Figure 12.
e
pt

Increasing the dissolved CO2 content can also influence the crystallization temperature (Tc)
ce

and the glass transition temperature (Tg). All of the literature reported that both of these

temperatures were consistently depressed due to the CO2‘s plasticization effect [54-62]. Thus,
Ac

dissolved CO2 can provide much lower processing temperatures during such processes as

extrusion foaming, foam injection molding, and bead foaming. Figure 13 shows that with

increased CO2 pressure these temperatures decreased almost linearly, although trends differed

with the type of PLA material used.

Figure 13.

12 
 

Page 12 of 63
As discussed earlier, the plasticizing CO2 molecules can influence crystallization kinetics at

various pressures. Therefore, PLA’s crystal melting temperature (Tm) would also be affected by

the varying crystalline structure and perfection under a pressurized gas. Takeda et al. [54], Yu et

al. [55], Li et al. [56], and Nofar et al. [61] observed the Tm depression of PLA when it was

t
exposed to various CO2 gas pressures. As Figure 14 shows, the Tm depression occurred in

ip
different types of PLA materials when they were exposed to a high CO2 pressure. In fact, with

cr
increased CO2 pressure the large number of small-sized imperfect crystals that had been induced

us
had a lower melting temperature than that of the original PLA under atmospheric pressure,

because of the less perfect crystals at an elevated pressure [61].


an
Figure 14.
M
3- Foaming Mechanisms of PLA
d

3-1- Effects of Crystallization


e

During PLA foam processes, PLA’s inherently low melt strength [15] leads to cell
pt

coalescence and cell rupture during cell growth. Moreover, its low melt strength causes gas loss
ce

during foam expansion, which results in severe shrinkage [66]. Enhancing PLA’s crystallization

kinetics during foaming has been recognized as an effective way to overcome its weak
Ac

viscoelastic properties and to improve its foaming behaviors (i.e., cell nucleation and expansion)

[37-40]. Crystallization during foaming can improve PLA’s low melt strength through the

network of nucleated crystals [37-40]. This will consequently increase PLA’s ability to expand

by minimizing gas loss and cell coalescence [38-40]. It should be also noted that too high a

crystallinity would also suppress foam expansion due to excessive stiffness in the matrix and less

gas dissolution [66-67]. On the other hand, according to heterogeneous cell nucleation theory

13 
 

Page 13 of 63
[68-69], cell nucleation can be promoted around the nucleated crystal [64, 67, 70-71] through

local pressure variations [72], thereby significantly improving the foamed samples’ final cell

density. Based on this theory, cells can be nucleated at the interface of a polymer and hard phases

such as additives (i.e., cell nucleating agents) and crystals. The heterogeneous cell nucleation rate

t
is expressed as follows [72-75]:

ip
cr
(1)

us
where f is the frequency factor representing the frequency through which gas molecules are

joining the embryo nucleus. C is the concentration of heterogeneous nucleation sites, which is

nucleation is given as follows:


an
directly related to the hard-phase concentration. The Gibbs free energy for heterogeneous
M
(2)
d

is the geometrical factor that relates to the surface geometry of the hard phases, is the
e
pt

contact angle between the bubble surface and the solid surface measured in the liquid phase, and

is the semi-conical angle of conical cavities that models the irregular surfaces of nucleating
ce

agents. Hence, the presence of an interface greatly reduces the free energy for nucleation and

leads to a significantly increased nucleation rate.


Ac

Taki et al. [71] visualized the cell nucleation behavior of a PLLA/CO2 mixture in a batch

process. They showed that the presence of spherulites in the PLLA matrix promoted the number

of cell nuclei. Figure 15 shows that as depressurization occurred, the cells were nucleated around

the growing PLLA spherulites. Taki et al. [71] also showed that the number of nucleated cells

was significantly promoted as the spherulites density increased. They explained that the growing

14 
 

Page 14 of 63
spherulites expelled the dissolved CO2 from the interface phase between the amorphous and

spherulites sections of the PLLA. Therefore, an increased CO2 concentration in the interface

caused cell nucleation around these growing spherulites. Moreover, they demonstrated that the

cell density increase was a function of the spherulites area. Li et al. [56] also proposed that by

t
controlling the amount of isothermal crystallization under the compressed CO2, a varying range

ip
of cell sizes (80-270µm) and expansion ratios (15-30) could be achieved.

cr
Figure 15.

us
Liao et al. [64] also explored the effects of crystallization on the foaming behavior of

an
PLLA samples with various crystallinities and crystallite sizes. The PLLA samples were first

saturated with various CO2 pressures at 25oC to induce different degrees of crystallinity in the
M
PLLA samples. Then, all of the saturated samples were quenched and re-saturated with 2.8 MPa

CO2 at 0oC. Before the re-saturation, five PLLA samples were achieved with the crystallinity and
d

crystallite size ranging from 12%-30% and 2.2 nm-12.7 nm, respectively. The samples were
e

subsequently foamed at temperatures ranging between 50oC-100oC. The foam density trends
pt

showed that a higher amount of larger size pre-existing crystals created a higher foam density

(i.e., close to the neat PLLA’s density), although finer cells were achieved in the foamed regions.
ce

This means that the increased stiffness throughout the large crystalline domain restricted cell
Ac

growth and further expansion. On the other hand, in a lower amount of pre-existing crystallinity

with smaller crystallite sizes, both cell nucleation and cell growth were promoted, although the

cell growth rate dominated the cell nucleation rate. In other words, a lower degree of crystallinity

facilitated both cell growth and the eventual foam expansion ratio.

Several researchers have also investigated the effect of crystallization on PLA foaming

behavior during the foam extrusion process [37-40]. It is well known that isothermal melt

15 
 

Page 15 of 63
crystallization occurs at temperature ranges between the PLA’s Tm and Tg [59, 76-80]. As

explained earlier, during extrusion foaming, where the PLA/gas mixture can encounter

isothermal melt crystallization, a certain amount of nucleated crystals can affect the PLA’s

foaming behavior by enhancing the cell density and the expansion ratio. Using a twin-screw

t
extruder, Mihai et al. [39] showed that the presence of nucleated crystals in the flowing PLA/CO2

ip
mixture inside the extruder barrel acted as cell nucleation sites during the foaming step. This also

cr
enhanced the PLA’s melt strength and thereby its expansion ratio. Specifically, with a higher gas

us
content, the CO2-induced crystallization created larger cell nucleation sites. This eventually

resulted in higher expansion and increased cell density in the PLA foams. As Figure 16 shows,
an
with a lower CO2 content, the cell coalescence was severe due to the PLA’s low melt strength,

and the final foam expansion was not considerable. However, as the CO2 content increased, the
M
foaming behavior improved, and the expedited crystal nucleation rate caused by the gas increase

also promoted cell density and melt strength (i.e., the expansion ratio). Similar findings have also
d

been confirmed in work done by Mihai et al. [37], where they foamed various PLA materials
e

with different branching degrees using a twin-screw extruder. They demonstrated that the
pt

crystallizable PLA had a higher expansion ratio and a finer cell density due to the presence of
ce

nucleated crystals during extrusion foaming. The nucleated crystals, specifically at a higher CO2

content, must have enhanced the PLA’s melt strength and provided more cell nucleation sites
Ac

before the foaming action took place (Figure 17).

Figure 16.

Figure 17.

16 
 

Page 16 of 63
Wang et al. [38] also conducted extrusion foaming of linear and two-branched PLA

materials with a D-lactide content of 4.5%. The experiments were done in a tandem extrusion

system consisting of two single extrusion lines. The second extruder (i.e., the tandem line) was

used to further mix and cool the melt. A resistance die was used after the tandem line and prior to

t
the foaming step to induce further isothermal melt crystallization. Crystallization was controlled

ip
by varying the length of the die reservoir. Figure 18 shows the expansion ratio of the long-chain

cr
branched PLA foamed samples with 9% CO2 content at various die temperatures. When a longer

us
die residence time was used, the isothermally induced crystallization improved, and the high

expanded foams were obtained in a wider processing window. This was due to the increased melt
an
strength of the PLA/CO2 mixture wherein the molecules were connected through crystals, which

reduced gas loss during foam expansion. According to the reported SEM images, the induced
M
crystallization also provided more nucleated cells and more closed-cell content (i.e., minimized

cell coalescence) due to the increased melt strength.


d

Figure 18.
e
pt

Nofar et al., whose work is reported in [40], also used the same tandem extrusion system to
ce

investigate the induction of isothermal melt crystallization along the second extruder in a tandem

line. They varied its temperature profile, and therefore, they controlled the extruded PLA foam
Ac

properties by varying the amount of induced crystallization along the tandem line. Nofar et al.

used a linear PLA with a D-lactide content of 4.6% and PLA-clay nanocomposites [40]. They

showed that reducing the temperature profile along the second extruder, with a constant residence

time, significantly enhanced the cell density and expansion ratios of the foamed PLA and the

PLA-clay nanocomposites. As noted earlier, the isothermally induced crystallization improved

the PLA’s melt strength and provided a larger number of cell nucleation sites during extrusion

17 
 

Page 17 of 63
foaming. Figure 19 shows the expansion ratio enhancement of the foamed samples, when the

temperature profile in the tandem line was reduced in the order of Profile 1 > Profile 2 > Profile 3.

Figure 19.

Recent studies of PLA bead foaming also focused on producing expanded PLA (EPLA)

t
bead foams with a double-crystal melting peak structure [81-82]. In this technique, the high-

ip
temperature melting peak crystals that form during the isothermal saturation step in a batch-based

cr
bead foaming process maintain the bead geometry [83-84]. The formation of this high melting

us
temperature crystal results from the crystal perfection during which saturation occurs. The gas

saturation occurs around the PLA’s melting temperature at which time the unmelted crystals
an
become more perfect with the higher melting temperature [85-86]. These induced high-melting

temperature crystals can significantly affect cell nucleation and the expansion behaviors of PLA
M
bead foams. As explained earlier, cell nucleation can be promoted around the formed crystals [70-

71]. On the other hand, the low-melt strength PLA molecules become high-melt strength material.
d

This increases the PLA’s ability to expand by minimizing both gas loss and cell coalescence. But
e

due to the increased stiffness of the matrix, too high a crystallinity will depress the foam’s
pt

expansion ability [66-67].


ce

3-2- Effects of Molecular Architecture and Configuration


Several studies have explored PLA foaming behavior’s dependency on its molecular
Ac

architecture (i.e., the degree of branching) [16, 37-38, 87-90] and on its molecular configuration

(i.e., the D-lactide content) [39, 91]. PLA foaming behavior depends greatly on these molecular

structures because they affect its molecular mobility, crystallization kinetics, and rheological

properties. By modifying PLA’s molecular structure, researchers have attempted to achieve low

density PLA foams with fine and uniform cell morphologies.

18 
 

Page 18 of 63
3-2-1- Effects of Chain Branching
Corre et al. [87] studied the effects of various branching degrees on PLA foaming behavior

in batch foaming. They showed that the addition of chain extender (CE), Joncryl® 4368-BASF,

broadened the foaming window. This was mainly due to the increased rheological properties and

t
the PLA’s melt strength. At the highest saturation temperature of 140oC, the PLA samples with

ip
branching degrees that increased from 0% to 3% had expansion ratios of 2.5 to 32, respectively.

cr
All of the samples had cell densities of between 105-106 cells/cm3. At the lowest saturation

temperature of 110oC, when the chain extender content was increased from 0% to 3% the

us
expansion ratio decreased from 21 to 1.3, although the cell densities increased from around 3x104

an
to around 4x1010, respectively. This was explained by a high concentration in the crystalline

phase at lower saturation temperatures, where cell nucleation dominated cell growth due to the
M
increased stiffness of the matrix.

Di et al. [16] and Marrazzo et al. [88] also reported on the reactive modification of the PLA
d

chain structure. Their use of 1,4-butanediol and 1,4-butane diisocyanate as low molecular weight
e

chain extenders enhanced the PLA’s melt viscosity and elasticity. These improvements resulted
pt

in batch-based PLA foams that had lower densities as well as larger cell densities. They showed
ce

that the expansion ratio, the average cell size and the average cell density of the modified PLA

foamed samples increased to 19-fold, 28 µm, and 6.7 x 108 cells/cm3, respectively. These values
Ac

in unmodified foamed samples were, respectively, 10-fold, 227 µm and 7.6 x 105 cells/cm3.

Using continuous extrusion foaming to produce low density foams, a few researchers [37-

38, 89] studied the impact of chain branching on PLA’s foaming features. In these studies, CO2

was used as the physical blowing agent. Figure 20 compares the maximum expansion ratio and

the average cell density of the PLA foamed samples achieved with various chain extender (CE)

content. Induction of the CE promoted the PLA’s molecular weight, complex viscosity, elasticity,

19 
 

Page 19 of 63
and elongational viscosity. The samples also showed strain hardening behavior. Therefore, the

resultant foam morphology and its uniformity were significantly improved with a wider

processing window, and the final expansion ratio was increased with a higher closed-cell content.

However, only one study reported an increase in the final cell density of the PLA foamed

t
samples with CE. Wang et al. [38] reported that the highest expansion ratio of 42 was achieved

ip
with a cell density of around 109 cells/cm3, which is the threshold for microcellular foam. This

cr
could have been caused by the use of a tandem extrusion system, which provides a more uniform

us
cooling profile along the second extruder in the PLA/gas mixture. Being uniformly cooled, the

branched PLA may have better crystallized because branched molecules can act as crystal
an
nucleating sites. It should be noted that Mihai et al. [37] and Pilla et al. [89] used twin- and

single-screw extruders, respectively.


M
Figure 20.
d

In addition to the extrusion process, Pilla et al. [90] also investigated the injection foam
e

molding behavior of linear and branched PLA using supercritical nitrogen as the blowing agent.
pt

They showed that the addition of 0.8 wt% CE to PLA improved the foam morphology of the
ce

injected PLA foamed samples and promoted cell density (Figure 21). They attributed such

increased final mechanical properties as specific strength, specific toughness and strain at break
Ac

to the foamed samples’ structural enhancements.

Figure 21.

3-2-2- Effects of D-Lactide Content


It is well known that reduced D-lactide content increases PLA’s crystallization kinetics

while increasing its melting temperature [19-20, 76, 91-92]. This indicates that, although

20 
 

Page 20 of 63
enhanced crystallization kinetics might be beneficial for PLA foaming purposes, a higher

temperature will be required for batch processing. Fujiwara et al. [93] investigated the effect of

D-lactide content on PLA’s foaming behavior using batch foaming with supercritical CO2 as the

blowing agent. They reported that the average cell sizes of semi-crystalline PLA foamed samples

t
with D-lactide content of 1% and 4.2% were 5.4 µm and 3.3 µm, respectively. By contrast, the

ip
amorphous PLA samples with a D-lactide content of 10% and 28.5% were not foamable. This

cr
suggested that the achieved microcellular morphology of PLA samples with low D-lactide

us
content was crystallinity dependent. At the same time, Garancher et al. [92] reported that the

batch foaming of amorphous PLA with a D-lactide content of 11.8% resulted in a finer cell
an
morphology, with a higher closed-cell content. Meanwhile, the low D-lactide content PLA grades

with higher crystallinity had larger cell sizes and a higher open-cell content, with a perforated
M
surface morphology, which contradicted Ref [93].

Mihai et al. [39] explored the effects of varying D-lactide content on PLA’s final foam
d

morphology through continuous extrusion foaming using a twin-screw extruder. In Figure 22, the
e

expansion ratio and the final crystallinity of the PLA foamed samples are compared with various
pt

D-lactide and CO2 content. Mihai et al. reported that, during foam processing, the crystallinity of
ce

the PLAs with low D-lactide content was further enhanced by dissolved CO2. This resulted in

foams with finer cells and a more uniform morphology. Higher expansion ratios were also
Ac

achieved. These improvements were caused by the increased melt strength of the PLA/CO2

mixture during the extrusion process, in which a larger number of cell nucleation sites were

created around the nucleated crystals. This is the first paper that claimed the formed crystals

during extrusion foaming enhance both cell nuclei density and the expansion ratio, which is

confirmed by Mihai et al. [37], Wang et al. [38] and Nofar et al. in [40]. The maximum obtained

21 
 

Page 21 of 63
expansion ratio was 39-fold, when PLA with D-lactide content of 2% and CO2 with 7 wt% were

used. As Figure 22 shows, the strain-induced crystallization that occurred during the foam

expansion further promoted the foamed samples’ crystallinity.

Figure 22.

t
ip
3-3- PLA Compounds

cr
3-3-1- PLA Blends
Several studies have investigated the foaming behavior of PLA samples blended with other

us
biodegradable polymers such as starch, poly(butylene adipate-co-terephthalate) (PBAT),

an
polyvinyl alcohol (PVA or PVOH), and polyhydroxybutyrate-valerate (PHBV) [94-103]. Table 1

shows the detailed blend information, the processing methods, the blowing agent types, the
M
maximum achieved expansion ratios, and the minimum obtained average cell sizes.

Table 1.
d

A few studies have investigated the foaming behavior of PLA/starch blends using various
e

approaches [94-98]. Preechawong et al. [94] prepared PLA/starch foams through a baking
pt

process in a hot mold using water. The PLA improved the starch-based foams’ resistance to

water absorption as well as the blends’ ultimate tensile strength and ductility. Zhang et al. [97-
ce

98] used extrusion foaming with water as the blowing agent. They showed that increased water
Ac

content reduced the melting temperature of the blend. However, to achieve high foam expansion

and uniform foam morphology an optimum water content was required. Hao et al. [95] and Mihai

et al. [96] also investigated the foaming behavior of PLA/starch blends, but they used

supercritical CO2 as the blowing agent. Hao et al. [95] aimed to confirm that PLA/starch blends

have the capacity for use in medical applications. But their investigation was limited to batch

processing on the foaming behaviors of PLA (60 wt%)/starch (40 wt%). By contrast, Mihai et al.

22 
 

Page 22 of 63
[96] explored PLA/starch blends with various starch content through continuous extrusion

foaming. They demonstrated that low-density PLA/starch foams with an open-cell structure were

achieved, but in order to obtain a finer cell structure with less open-cell content, interfacial

modification of the PLA was necessary.

t
Some studies also reported on the foaming of PLA/PBAT blends [99, 101-102]. PBAT has

ip
been used to improve PLA’s low melt strength and low elasticity. Different compatibilizers were

cr
also employed to increase the compatibility between PLA and PBAT. To increase the cell

us
nucleation rate, all of these studies used such cell nucleating agents as nanosilica [99], talc [101],

and nanoclay [102]. Yuan et al. [99] and Pilla et al. [101] conducted continuous extrusion
an
foaming while Yuan et al. used a chemical blowing agent (Azodicarbonamide). Pilla et al.

applied CO2 as a physical blowing agent. Yuang et al. showed that the addition of 10 wt% PBAT
M
significantly promoted PLA’s foam uniformity. When a small amount of maleic anhydride, a

compatibilizer, was added, it heightened the blends’ physical and mechanical properties, but it
d

reduced the foamed samples’ cell density. On the other hand, Pilla et al. reported that the addition
e

of a compatibilizer to PLA/PBAT blends with 55 wt% of PBAT increased the cell density of the
pt

foamed samples, although it suppressed the expansion ratio. According to their DSC analysis,
ce

the two distinct melting peaks of PLA and PBAT merged into one peak at the lower temperatures.

Li et al. [102] also studied foam injection molding of PLA/PBAT blends. They showed that
Ac

nanoclay platelets improved the interfacial adhesion between PLA and PBAT. Consequently, cell

density was significantly enhanced through the presence of nanoclay particles and by the

improved interfacial adhesion of the two polymers.

Kramschuster et al. [100] and Zhao et al. [103] also used foam injection molding to

produce PLA blends with PVOH and PHBV, respectively. Kramschuster et al. [100] added salt

23 
 

Page 23 of 63
particulates to the blends and leached the foamed samples in water after the injection process to

produce a porous interconnected foam structure as a scaffold for tissue engineering. Since the

PVOH was used in granule shape, further mixing of the PLA with the PVOH granules in a twin-

screw extruder further reduced the dispersed PVOH phase size in the PLA matrix. Consequently,

t
the finely dispersed PVOH phase improved the uniformity of the blend’s final foam morphology.

ip
Zhao et al. [103] also reported that blending 30 wt% PHBV with PLA increased the cell density

cr
of the injection-molded foam samples. However, as the PHBV content increased beyond 30 wt%,

us
the miscibility of the PLA and the PHBV became a serious concern, and the foam morphology

was suppressed.

3-3-2- PLA Composites


an
Many researchers have also studied foaming of PLA composites using micro-sized natural
M
and synthetic additives [42, 104-108]. Micro-sized additives can indeed increase the cell

nucleation rate during foaming, because they act as heterogeneous cell nucleation agents
d

(Equations 1 and 2). Table 2 shows different research group’s examinations of PLA composite
e

foaming. Table 2 also lists the foaming processes, the blowing agent types, the maximum
pt

achieved expansion ratios, the minimum obtained average cell densities, and the applied analysis
ce

for each study.

Table 2.
Ac

In order to manufacture fully degradable biocomposite foams with enhanced final

mechanical properties, a few researchers have investigated PLA foaming behavior using several

natural fibers/additives, which have included flax fiber [105], silk fibroin powder [106], wood

flour [107], and microfibrillated cellulose (MFC) [108]. Pilla et al. [105] reported on the foam

injection molding of PLA biocomposites using flax fiber contents of 1, 10, and 20 wt%. Silane

was also used as a coupling agent to create a strong interface bonding between the PLA and the

24 
 

Page 24 of 63
fibers. The foaming results showed that fiber content improved both the cell density and the

foams’ crystallinity. The average cell size of 3 µm was achieved when 20 wt% fibers were used.

As Pilla et al. reported, only the specific modulus of the foamed samples was improved by

increased fiber content. Kang et al. [106] also investigated the effects of 1, 3, 5 and 7 wt% of silk

t
fibroin powder on the PLA’s foaming behavior in batch foaming using CO2 as the blowing agent.

ip
The biocomposites were prepared using a solution casting technique. Kang et al showed that the

cr
addition of silk fibroin powder decreased the average cell size from 52 µm in neat PLA to around

us
15 µm in PLA with 7 wt% silk. Further, Matuana et al. [107] explored PLA’s batch foaming

behavior with 10, 20, 30, and 40 wt% of wood fiber. Silane was also used as the coupling agent.
an
They reported that an increase in the wood flour content reduced the expansion ratio from around

10-fold in neat PLA to 2-fold in PLA-40 wt% wood flour due to the matrix’s high stiffness.
M
However, the cell size was reduced to an average of around 35 µm. Boissard et al. [108]

described the production of microfibrillated cellulose (MFC)-reinforced PLA biocomposite


d

foams using supercritical CO2. Similar to the previous study, the expansion ratio of the neat PLA
e

was reduced from 6.8-fold to 3.8-fold in PLA with 5 wt% MFC. The compression modulus of
pt

these samples was also increased from 25 MPa in the neat PLA to 47 MPa in PLA with 5 wt%
ce

MFC.

In another study, Kramschuster et al. [104] investigated the PLA’s foam injection molding
Ac

behavior with recycled paper shopping bag fibers. The fiber contents of 10 wt% and 30 wt%

were used, and silane was the coupling agent. They showed that the addition of fibers up to 30

wt% enhanced the cell density of the foamed samples as well as their storage modulus, specific

modulus, and specific tensile strength. Ameli et al. [42] studied PLA’s foam injection molding

behavior with talc particles up to 5 wt%. They confirmed that the talc and the improved

25 
 

Page 25 of 63
crystallinity created a more uniform foam morphology with larger cell density and smaller cell

sizes. They demonstrated that the toughening effect of the talc and the induced crystallinity in

conjunction with the improved cell morphology enhanced the mechanical properties of the PLA-

talc foamed samples.

t
ip
3-3-3- PLA Nanocomposites
Compared to conventional polymer foams, microcellular polymer foams are described by

cr
cell densities that range between 109-1012 cells/cm3 [109]. For more than two decades, the

us
manufacture of low-density microcellular polymer has been of intense interest to both

researchers and industry. This is because microcellular polymers have several advantages over

an
their unfoamed or conventional polymer foam counterparts. These include their reduced weight,

improved toughness, higher impact strength [110-112]; longer fatigue life [113]; superior heat-
M
insulation properties [114-116]; excellent sound insulation properties [117]; and high light

reflectability [118-119].
d

The presence of a large number of dispersed nanoparticles in polymers can significantly


e

promote cell density in foamed samples. Nanoparticles act as heterogeneous cell nucleating sites
pt

[120-121] through the local pressure variations that are created around them [68-69, 72, 122].
ce

Thus, their effectiveness as cell nucleating agents strongly depends on how well they disperse

and exfoliate in the polymer matrix [31, 40, 73, 123-127]. While nanoparticles can promote cell
Ac

density, the dispersed particles can simultaneously enhance the melt strength of the polymer

matrix, thereby stabilizing the nucleated cells by minimizing cell coalescence [124]. At the same

time, the reduced gas diffusion rate can minimize the cell ripening phenomenon [128].

Consequently, a larger number of nucleated cells will last longer during cell growth, and the final

expansion ratio of the foam products will also be promoted. Recently, Wong et al. [73] proposed

that at a very low processing temperature, the intercalated particles have greater rigidity and

26 
 

Page 26 of 63
thereby can create a higher cell density because of the higher local stresses that are favorable for

cell nucleation. In other words, better dispersion would not play a role but the rigidity of the

intercalated particles would be of prime importance in generating more stress variations.

As shown in the previous section, the effectiveness of nanoparticles on cell nucleation was

t
much greater than were the micro-sized additives, although micro-sized additives and fibers can

ip
significantly promote both cell density and the final mechanical properties of foamed samples.

cr
Nanocellular polymer foams have also been considered as the new generation of low

us
density foams with cell densities beyond 1012 cells/cm3 and cell sizes below 1 µm [129], which

distinguishes them from microcellular polymer foams. The most important application of
an
nanocellular foams is in super-thermal insulation [129]. The first studies were done by Fujimoto

et al. [130] and Ema et al. [131], who introduced nanocellular foams into PLA-clay
M
nanocomposites through the batch foaming process. As previously noted, according to

heterogeneous cell nucleation theory [69, 72-75], cell nucleation occurred on the boundary of the
d

PLA and the nanoclay particles. The resultant effects of PLA’s melt strength and the presence of
e

nanoparticles could, therefore, have hindered cell growth and induced nano-cells under specific
pt

processing conditions [130-131].


ce

Several studies over the past decade have looked at the effects of nanoparticles on PLA’s

microcellular and nanocellular foaming behaviors [25, 31, 40, 88, 102, 130-140]. Table 3
Ac

outlines these studies, showing the foaming process used, the blowing agent type, the maximum

achieved expansion ratio, the minimum obtained average cell size, and the applied analysis for

each one.

Some of these studies used a batch foaming apparatus to explore the influence of

nanoparticles on PLA’s foaming behavior [25, 31, 88, 130-131, 134-135, 139]. All of these

27 
 

Page 27 of 63
studies used nanoclay due to its long aspect ratio and platelet-shaped structure, which provides

superior viscoelastic behavior in the polymer matrix [25, 31, 88, 130]. They found that the PLA-

clay nanocomposite foams produced foams with homogeneous cell morphology and average cell

sizes that ranged from 165 nm to 25 µm. The majority of the nucleated cells appeared to be of a

t
closed-cell structure. In fact, due to the biaxial stretching during cell growth, the long aspect ratio

ip
platelet-shaped nanoclay particles were oriented along the cell walls, and the increased cell wall

cr
strength inhibited cell rupture [120]. All of these studies found that the nanoparticles behaved as

us
heterogeneous cell nucleating agents, and that a varying range of cell sizes were obtained with

different nanoclay loadings. As the nanoclay particles increased, the cell density of the foamed
an
samples increased. Thereby, the cell size of the PLA foamed samples could be controlled by the

amount of nanoparticles loaded. Further, Tsimpliaraki et al. [139] reported that while PLA cell
M
nucleation behavior depended on the nanoclay loading, it could also be significantly affected by

chemically modifying the nanoparticles. They observed that adding organically modified
d

nanoclay particles resulted in more heterogeneous cell nucleation, with smaller cell sizes. This
e

was due to the nanoclay’s different degree of dispersion in the PLA nanocomposite [40,130].
pt

Some of the literature has explored microcellular foam injection molding of PLA
ce

nanocomposites. Pilla et al. [132] investigated the microcellular foaming behavior of PLA carbon

nanotube nanocomposites. They explained how a small amount of carbon nanotube noticeably
Ac

promoted the foamed samples’ cell density due to its effect on cell nucleation. They

demonstrated that the microcellular foaming process further dispersed the nanotubes within the

PLA, and that this was caused by the dissolved gas’s plasticization effect. Other researchers have

also explored the effects of nanoclay on microcellular foam injection molding of PLA

nanocomposites [102, 133, 136-137, 140]. In all of these studies, the average cell size of the PLA

28 
 

Page 28 of 63
nanocomposites were significantly reduced with increased nanoclay loading. Kramschuster et al.

[133] found that the strain at break and the specific toughness of the microcellular PLA

nanocomposites showed more improvement than did the neat PLA foamed samples. They

attributed this improvement to the homogenous foam morphology and to the smaller cell sizes

t
caused by the nanoclay particles. Recently, Ameli et al. [140] reported that, during microcellular

ip
foam injection molding of PLA clay nanocomposites, not only the nanoparticles but also the

cr
induced crystallization promoted the cell nucleation rate and the homogeneity of the cell

us
morphology. They showed that an average cell size of less than 50 μm was induced in the PLA

nanocomposites when an expansion ratio of 2.8-fold was achieved. They also found that
an
nanoclay enhanced the flexural and impact properties and the strain at break in the injected

microcellular PLA foams with a high void fraction.


M
A few studies also explored the continuous extrusion foaming behavior of PLA

nanocomposites using supercritical CO2 [40, 138]. It was demonstrated that nanoclay, dissolved
d

CO2, and the shearing effect during extrusion expedited the PLA’s slow crystallization kinetics
e

[40]. Nofar et al., whose work is reported in [40], showed that nanoparticles, in conjunction with
pt

the crystallization induced during foam processing, behaved as cell nucleating sites and thereby
ce

created foams with a larger cell density, confirming Mihai et al.’s earlier results [37, 39]. These

two factors further enhanced the PLA’s melt strength and, subsequently, low-density
Ac

microcellular foams were created. Matuana et al. [138] also reported that microcellular PLA

foams were successfully produced during the extrusion foaming of PLA clay nanocomposites.

Table 3.

29 
 

Page 29 of 63
4- PLA Foam Processing Technologies
Many efforts have been devoted to produce PLA foams, specifically microcellular foams,

through various manufacturing methods. A few studies have focused only on lab-scale batch

foaming systems. Batch foaming is mainly used to identify the effects of material composition,

t
ip
the blowing agent type and content, and the use of solvent on PLA’s foaming behaviors. But a

few studies have reported that batch-based foams can be geared towards biomedical scaffolding

cr
and tissue engineering applications [141-157]. Since the early 1990s, some attempts have been

us
made to produce PLA porous foam structures by using organic solvents [141-153]. However, the

organic residue that remained in the porous PLA foams could be harmful in biomedical
an
applications [154]. For the first time, Mooney et al. [154] introduced the use of high pressure

CO2 as a physical blowing agent to produce porous PLA foams through a static batch foaming
M
process. Since this supercritical physical blowing agent does not remain in the foam as a toxic or

harmful residue, the method received a great deal of attention. After the gas dissolved in the
d

PLA, the foams were produced by rapid depressurization. Subsequently, due to the
e

thermodynamic instability that was created, cell nucleation and cell growth occurred, and PLA
pt

foams with almost 93% porosity were produced. Later on, other researchers [155-157] attempted
ce

PLA batch foaming to produce PLA scaffolds by using supercritical CO2. Nevertheless, it should

be noted that, due to its small and non-continuous production scale, batch process foaming is not
Ac

cost effective. Therefore, to produce PLA foams for biomedical and commodity applications it is

preferable to develop continuous processing technologies. In the sections that follow, we explain

the three main foam processing technologies currently in use in PLA foam manufacturing. While

these may not be completely well established, research to develop the desired final PLA foam

products is ongoing.

30 
 

Page 30 of 63
4-1- PLA Extrusion Foaming
In extrusion foaming processes, the polymer is first fed into the extruder, and then the

blowing agent is injected into the extruder barrel to dissolve the gas into the polymer melt under

high pressure. The dissolved gas will subsequently plasticize the polymer melt, and a uniform

t
polymer/gas mixture will flow along the extruder. The polymer/gas mixture will then emerge

ip
from the die. The induced pressure drop across the die lip will cause foaming. The pressure drop

cr
will create thermodynamic instability and cause phase separation, i.e., cell nucleation and cell

us
growth [109]. Foam stabilization is the next important parameter, and it depends on the

polymer’s viscoelastic and strain-hardening behaviors [158]. The use of variable die geometries
an
will shape the extruded foams into two-dimensional geometric products.

Recently, efforts have been made to manufacture low-density microcellular PLA foams
M
through the extrusion process using CO2 as the physical blowing agent [37-40, 89, 96-97, 101,

138, 159]. These foams can be used in various applications such as packaging, food trays,
d

construction, and insulation [9-15]. The use of gaseous and supercritical CO2 is important
e

because it is environmentally-friendly, inexpensive, and non-flammable. When CO2 is in its


pt

supercritical condition, its solubility and diffusivity in polymers significantly increase.


ce

Consequently, the plasticization of the dissolved CO2 decreases the polymer’s Tc and Tg. Thus,

the temperatures required for foam processing will be reduced [160].


Ac

Using high pressure CO2, Lee et al. [66] and Reignier et al. [159] investigated the extrusion

foaming behavior of commercially available linear amorphous PLA. Although they both

achieved foam expansion of around 25-fold with the use of 9 wt% CO2, the foam morphologies

were not uniform and the majority of cells had coalesced. Moreover, the PLA foams had poor

mechanical properties and shrank an average of 70%, after remaining in an atmospheric

31 
 

Page 31 of 63
condition for 48 hours. On the other hand, Mihai et al. [37, 39, 96] investigated the extrusion

foaming behavior of linear and branched, amorphous and semi-crystalline PLAs as well as

PLA/thermoplastic starch blends using a twin-screw extrusion system with a distinct screw

configuration. They reported that PLA foams with an expansion ratio of up to 40-fold with a fine

t
cell morphology (i.e., an average cell size ranging from 50-100 µm) were achieved when 9 wt%

ip
CO2 was used. They also reported that biaxial stretching during foaming further induced

cr
crystallinity in the PLA foamed samples, and that foams with a high degree of crystallinity at

us
around 30% were obtained. Matuana et al. [138] claimed that low-density microcellular PLA

foams were obtained during extrusion foaming, when cell nucleating agents were used. Pilla et al.
an
[89, 101] investigated the extrusion foaming behavior of linear and branched PLA as well as

PLA/PBAT blends using a single-screw extruder. Although the samples did not reach foam
M
expansion ratios beyond 5-fold, the cell morphology was uniform, and the minimum average cell

size was 10 µm. Furthermore, Wang et al. [38] and Nofar et al. in [40] all used a tandem-line
d

extrusion system to investigate the foaming behavior of PLA with various branching degrees and
e

different nanoparticles. The tandem-line extrusion system consisted of two single-screw


pt

extruders. The first extruder was designed to create a uniform polymer/gas mixture. The second
ce

extruder provided additional mixing and a uniform cooling profile within the polymer/gas

mixture along the barrel. These studies confirmed that PLA foams with expansion ratios of over
Ac

40-fold were achievable with average cell sizes of below 50 µm.

In other studies, different blowing agents were applied during the extrusion foaming of

PLA [97-99, 161]. Zhang et al. [97-98] examined the extrusion behavior of PLA/starch blends

using water as a foaming agent. They showed that cell sizes as large as 500 µm were induced,

although an expansion ratio of over 50-fold could be achieved. Yuan et al. [99] and Matuana et al.

32 
 

Page 32 of 63
[161] used a chemical blowing agent as the driving force for foaming. Both of their studies

obtained PLA foams with fine cells that had average cell sizes of 10-20 µm [99] and cell

densities of between 106-108 cells/cm3 [161]. However, the foam expansion ratio did not exceed

3-fold.

t
ip
4-2- PLA Foam Injection Molding
Foam injection molding provides various advantages such as lower material costs, high

cr
dimensional stability, greater energy efficiency, and a shorter time cycle. It can also improve

us
mechanical properties such as fatigue life, toughness, and impact strength [104, 132-133, 162-

165]. Generally, in foam injection molding, supercritical nitrogen is used as the foaming agent.
an
Although nitrogen has a lower solubility in polymers than CO2 [45], it has a powerful cell

nucleating force [166]. Supercritical nitrogen reduces the polymer melt’s viscosity due to the
M
blowing agent’s plasticization effect [167-168]. Therefore, the processing temperature can be

decreased, and both energy use and processing costs will be reduced [169]. The decreased
d

processing temperature is also helpful in working with temperature-sensitive bio-based polymers


e

such as PLA [132-133].


pt

Extensive research and characterization has been done on foam injection molding of
ce

various PLA materials [42, 90, 100, 102-105, 132-133, 136-137, 140]. The majority of these

studies have focused on high-pressure foam injection molding (i.e., MuCell Technology -
Ac

microcellular injection molding) of PLA. MuCell Technology is being widely used in structural

foaming because it offers greater control of cell nucleation and cell coalescence [170]. In this

technique, the cavity is fully filled, and the void fraction is limited to only 5-15%. Studies

applying the high-pressure foam injection molding technique showed that the use of chain

extender and the addition of micro-/nano-size additives could enhance the PLA’s foaming ability

33 
 

Page 33 of 63
and its cell nucleation power [90, 104-105, 132-133, 136-137]. These studies reported that a

uniform foam morphology with average cell sizes ranging between 3 µm- 40 µm were achieved

by foam injection molding of branched PLA and PLA micro/nanocomposites.

The high-pressure foam injection molding technique has recently been developed by

t
introducing a mold opening as an additional step [165, 171-174]. In this technique, the mold

ip
cavity is first pressurized and then fully filled with the polymer/gas mixture. Subsequently, the

cr
mold is opened in the thickness direction while the gas is depressurized from the cavity. Thus,

us
more uniform foam samples with a high void fraction can be produced. This technique has been

applied with polyethylene and polypropylene (Egger et al. [171], Sporres et al. [172] and
an
Ishikawa et al. [173-174]) and with thermoplastic polyolefin (Wong et al. [165]). Recently,

Ameli et al. [140] also adopted it to create foam injection molded PLA and PLA
M
micro/nanocomposites. The injected foam samples had a high surface quality, a uniform cell

morphology, a minimum average cell size of 38 µm, and high void fractions in the range of 50%
d

- 65%.
e

Low-pressure foam injection molding is another technique used to produce foam injected
pt

samples. Here, the cavity is only partially filled, and void fractions of up to 40% can be obtained.
ce

However, there is less control of cell nucleation, growth and coalescence. Therefore, achieving a

uniform foam morphology with a high cell density and a void fraction of up to 20% has been a
Ac

serious challenge [164]. Recently, Ameli et al. [42] used low-pressure foam injection molding

and fine cell PLA foams were achieved with an average cell size of 20 µm and a void fraction of

30%. They claimed that this foam structure was obtained because of presence of talc as crystal

and cell nucleating agent. The talc and the nucleated crystals promoted the PLA’s low melt

strength and enhanced heterogeneous cell nucleation.

34 
 

Page 34 of 63
4-3- PLA Bead Foaming
As discussed earlier, extrusion foaming has been used to obtain low-density PLA foam

products with simple geometries. On the other hand, foam injection molding has been used to

manufacture high-density foam products with three-dimensional complex geometries. Bead

t
foaming is another means by which to produce low-density foam products with complex three-

ip
dimensional geometries [81-86, 92, 175]. In this method, low-density bead foams are molded

cr
into the desired shape of the final foam product.

us
Although producing expanded PLA (EPLA) bead foams is not yet commercially well

established, several attempts have been made to manufacture PLA bead foams using a method
an
similar to that used for expanded polystyrene (EPS) [175-181]. This method involves the

saturation of the PLA particles at temperatures below the PLA’s Tg with the blowing agent,
M
specifically CO2. Further expansion in a pre-expander machine follows this, and then the

molding of the foamed beads (i.e., the particles) at a high temperature into the desired shape.
d

Although a few companies currently produce EPLA bead foams using this method [176-
e

179], good sintering among bead foams to make three-dimensional final foam products with
pt

strong mechanical properties remains a serious challenge. Recent studies on PLA bead foaming
ce

focused on producing EPLA bead foams with a double-crystal melting peak structure, [81-82] the

technique that is being used to produce expanded polypropylene (EPP) bead foams [83-86]. In
Ac

this technique, the high-temperature melting peak crystals formed during isothermal saturation in

a batch-based bead foaming process maintain the bead geometry, even at the high temperature

required for good sintering. The formation of this crystal-melting peak at a temperature higher

than the original crystals is due to the crystal perfection induced during the gas saturation stage

35 
 

Page 35 of 63
that occurred around the polymer’s melting temperature. After gas saturation, the low-

temperature melting peak forms during cooling while foaming takes place.

In the steam-chest molding process (i.e., the final step taken to induce bead sintering to

produce three-dimensional foam products), the EPLA beads are supplied to the mold cavity.

t
Then, steam at a temperature between the low-temperature melting peak and the newly created

ip
high-temperature melting peak is supplied to the mold cavity to heat the beads. When exposed to

cr
this high-temperature steam, the low-temperature melting-peak crystals of the PLA beads will

us
melt. The beads will then sinter to each other while the unmolten high-temperature melting peak

crystals in them will maintain their overall geometry. The subsequent foam products will have
an
good geometry in the shape of the mold cavity, as outstanding intra-bead sintering will have

occurred due to the molecular diffusion. This high-quality PLA bead sintering will give the
M
products exceptional mechanical properties.
d

5- Conclusions
e

Investigations of PLA foaming behaviors showed that enhanced crystallization kinetics


pt

significantly increases the expansion ratio and the cell density of foamed samples. The crystal

nuclei that are induced during foam processing can increase PLA’s inherently low melt strength
ce

through the crystal-to-crystal network. The crystal nuclei can also act as heterogeneous cell
Ac

nucleating agents and can promote cell density. It was further shown that too high a crystallinity

can also hinder the expansion ratio due to the PLA’s increased stiffness.

Moreover, modification of PLA molecules can influence PLA foaming behavior by

affecting its crystallization and its melt strength. Adding nanoparticles also improved the PLA

foams’ expansion and cell nucleation behavior by increasing its melt strength and enhancing its

36 
 

Page 36 of 63
heterogeneous cell nucleation power. Nanoparticles can also induce more crystal nucleation,

which further improves the expansion ratio and the cell density.

The literature showed that extrusion foam processing produced microcellular PLA foams

with an expansion ratio of up to around 45-fold. However, manufacturing nanocellular PLA

t
foams with a high expansion ratio remains a serious challenge.

ip
Foam injection molding produced PLA foamed samples with a minimum cell size of 3 µm.

cr
However, the achieved void fraction was around 20%. Injected PLA foamed samples with a void

us
fraction of 65% had been successfully achieved with an average cell size of 38 µm.

Several attempts have also been made to produce PLA bead foams by applying the
an
manufacturing technique used for EPS. But bead sintering during the molding stage has been a

steady challenge. Recently, PLA bead foams with a double-crystal melting peak have been
M
developed, and these apply the sintering technique used in EPP manufacturing.
d

Acknowledgments
e

The authors gratefully acknowledge financial support from the NSERC-Alexander Graham
pt

Bell Canada Graduate Scholarship (CGS), the Consortium of Cellular and MicroCellular Plastics

(CCMCP), the NSERC Network for Innovative Plastic Materials and Manufacturing Processes
ce

(NIPMMP), and from AUTO21.


Ac

References
[1] Grijpma DW, Pennings AJ. (Co)polymers of L-lactide, 2. Mechanical properties.
Macromol Chem Phys 1994;195:1649-1663.
[2] Perego G, Gella GD, Bastioli C. Effect of molecular weight and crystallinity on
poly(lactic acid) mechanical properties. J Appl Polym Sci 1996;59:37-43.

37 
 

Page 37 of 63
[3] Sinclair RG. The case for polylactic acid as a commodity packaging plastic. J Macromol
Sci Pure Appl Chem 1996;33:585-597.
[4] Tsuji H, Ikada Y. Blends of aliphatic polyesters. II. Hydrolysis of solution-cast blends
from poly(L-lactide) and poly(E-caprolactone) in phosphate-buffered solution. J Appl
Polym Sci 1998;67:405-415.

t
[5] Martin O, Averous L. Poly(lactic acid): plasticization and properties of biodegradable

ip
multiphase systems. Polymer 2001;42:6209-6219.
[6] Gupta B, Revagade N, Hilborn J. Poly(lactic acid) fiber: An overview. Prog Polym Sci

cr
2007;32:455-482.
[7] Lunt J. Large scale production, properties and commercial applications of polylactic acid

us
polymers. Polym Degrad Stab 1998;59:145-152.
[8] Drumright RE, Gruber PR, Henton DE. Polylactic acid technology. Adv Mater

[9]
2000;12:1841-1846. an
Fang Q, Hanna MA. Characteristics of biodegradable Mater-Bi®-starch based foams as
affected by ingredient formulations. Ind Crops Prod 2001;13:219-227.
M
[10] Auras R, Harte B, Selke S. An overview of polylactides as packaging materials.
Macromol Biosci 2004;4:835-864.
d

[11] Garlotta DJ. A literature review of poly(lactic acid). J Polym Environ 2001;9:63-84.
e

[12] Bandyopadhyay A, Basak GC. Studies on photocatalytic degradation of polystyrene.


Mater Sci Technol 2007;23:307-314.
pt

[13] Kosior E, Braganca RM, Fowler P. Lightweight Compostable Packaging: Literature


Review. Banbury UK: The Waste & Resources Action Programme, 2006. 49 pp.
ce

[14] Ajioka M, Enomoto K, Yamaguchi A, Suzuki K, Watanabe T, Kitahara Y. Degradable


foam and use of same. US Pat 5,447,962 1995.
Ac

[15] Lim LT, Auras R, Rubino M. Processing technologies for poly (lactic acid). Prog Polym
Sci 2008;33:820-852.
[16] Di Y, Iannace S, Di Maio E. L. Nicolais. Reactively modified poly(lactic acid): properties
and foam processing. Macromol Mater Eng 2005;290:1083-1090.
[17] Mikos AG, Thorsen AJ, Czerwonka LA, Bao Y, Langer R, Winslow DG, Vacanti JP.
Preparation and characterization of Poly (L-lactide acid) foams. Polymer 1994;35:1068-
1077.

38 
 

Page 38 of 63
[18] Dorgan J, Lehermeier RH, Mang M. Thermal and rheological properties of commercial-
grade poly(lactic acid). J Polym Environ 2000;8:1-9.
[19] Dorgan J. Rheology of poly(lactic acid). In: Auras R, Lim LT, Selke SEM, Tsuji H,
editors. Poly(lactic acid): Synthesis, Structures, Properties, Processing and Applications.
New York: John Wiley & Sons, 2010. p. 125-139.

t
[20] Dorgan J, Janzen J, Clayton M, Hait S, Knauss D. Melt rheology of variable L-content

ip
poly(lactic acid). J Rheol 2005;45:607-619.
[21] Palade LI, Lehermeier H, Dorgan J. Melt rheology of high l-content poly(lactic acid).

cr
Macromolecules 2001;34:1384-1390.
[22] Dorgan J, Williams J. Melt rheology of poly(lactic acid): Entanglement and chain

us
architecture effects. J Rheol 1999;43:1141-1155.
[23] Sungsanit K, Kao N, Bhattacharya SN, Pivsaart S. Physical and rheological properties of

[24]
an
plasticized linear and branched PLA. Korea-Aust Rheol J 2010;22:187-195.
Carlson D, Dubois P. Free radical branching of polylactide by reactive extrusion. Polym
Eng Sci 1998;38:311-321.
M
[25] Di Y, Iannace S, Di Maio E, Nicolais L. Poly(lactic acid)/organoclay nanocomposites:
thermal, rheological properties and foam processing. J Polym Sci Part B Polym Phys
d

2005;43:689-698.
e

[26] Gu SY, Ren J, Dong B. Melt rheology of polylactide/montmorillonite nanocomposites. J


Polym Sci Part B Polym Phys 2007;45:3189-3196.
pt

[27] Gu SY, Zou CY, Zhou K, Ren J. Structure-rheology responses of polylactide/calcium


carbonate composites. J Appl Polym Sci 2009;114:1648-1655.
ce

[28] Krishnamoorti R, Yurekli K. Rheology of polymer layered silicate nanocomposites. Curr


Opin Colloid Interface Sci 2001;6:464-470.
Ac

[29] Pluta M. Melt compounding of polylactide/organoclay: structure and properties of


nanocomposites. J Polym Sci Part B Polym Phys 2006;44:3392-3405.
[30] Ray SS. Rheology of polymer/layered silicate nanocomposites. J Ind Eng Chem
2006;12:811-842.
[31] Ray SS, Okamoto M. New polylactide/layered silicate nanocomposites. 6 Melt rheology
and foam processing. Macromol Mater Eng 2003;288:936-944.

39 
 

Page 39 of 63
[32] Ray SS, Maiti P, Okamoto M, Yamada K, Ueda K. New polylactide/layered silicate
nanocomposites. 1. preparation, characterization and properties. Macromolecules
2002;35:3104-3110.
[33] Wang B, Wan T, Zeng W. Dynamic rheology and morphology of polylactide/ organic
montmorillonite nanocomposites. J Appl Polym Sci 2011;121:1032-1039.

t
[34] Wu D, Wu L, Wu L, Zhang M. Rheology and thermal stability of polylactide/clay

ip
nanocomposites. Polym Degrad Stab 2006;91:3149-3155.
[35] Saeidlou S, Huneault M, Li H, Park CB. Poly(lactic acid) crystallization. Prog Polym Sci

cr
2012;37:1657-1677.
[36] Rasal RM, Janorkar AV, Hirt DE. Poly(lactic acid) modifications. Prog Polym Sci

us
2010;35:338-356.
[37] Mihai M, Huneault MA, Favis BD. Rheology and extrusion foaming of chain-branched

[38]
an
poly(lactic acid). Polym Eng Sci 2010;50:629-642.
Wang J, Zhu W, Zhang H, Park CB. Continuous processing of low-density, microcellular
poly(lactic acid) foams with controlled cell morphology and crystallinity. Chem Eng Sci
M
2012;75:390-399.
[39] Mihai M, Huneault MA, Favis BD. Crystallinity development in cellular poly(lactic acid)
d

in the presence of supercritical carbon dioxide. J Appl Polym Sci 2009;113:2920-2932.


e

[40] Nofar M, Park CB. Heterogeneous cell nucleation mechanisms in polylactide foaming.
In: Iannace S. Park CB, editors. Biofoams: Science and Applications of Bio-Based
pt

Cellular and Porous Materials.


Boca Raton FL: CRC Press, 2015;in press.
ce

[41] Srithep Y, Nealey P, Turng LS. Effects of annealing time and temperature on the
crystallinity, heat resistance behavior and mechanical properties of injection molded
Ac

poly(lactic acid) (PLA). Polym Eng Sci 2013;53:580-588.


[42] Ameli A, Jahani D, Nofar M, Jung P, Park CB. processing and characterization of solid
and foamed injection-molded polylactide with talc. J Cell Plast 2013;49:351-374.
[43] Li YG, Park CB, Li HB, Wang J. Measurement of the PVT property of PP/CO2 solution.
Fluid Phase Equilib 2008;270:15-22.

40 
 

Page 40 of 63
[44] Sato Y, Yurugi M, Fujiwara K, Takishima S, Masuoka H. Solubilities of carbon dioxide
and nitrogen in polystyrene under high temperature and pressure. Fluid Phase Equilib
1996;125:129-138.
[45] Li G, Li H, Turng LS, Gong S, Zhang C. Measurement of gas solubility and diffusivity in
polylactide. Fluid Phase Equilib 2006;246:158-166.

t
[46] Li H, Lee LJ, Tomasko DL. Effect of carbon dioxide on the interfacial tension of polymer

ip
melts. Ind Eng Chem Res 2004;43:509-514.
[47] Hasan MM, Li YG, Li G, Park CB, Chen P, Simha R. Determination of solubilities of

cr
CO2 in linear and branched polypropylene using a magnetic suspension balance and a
PVT apparatus. J Chem Eng Data 2010;55:4885-4895.

us
[48] Mahmood SH, Keshtkar M, Park CB. Determination of carbon dioxide solubility in
polylactide acid with accurate PVT properties. J Chem Thermodyn 2013;70:13-23.
[49] an
Sato Y, Inohara K, Takishima S, Masuoka H, Imaizumi M, Yamamoto H, Takasugi M.
Pressure-volume-temperature behavior of polylactide, poly (butylenes succinate), and
poly (butylenes succinate-co-adipate). Polym Eng Sci 2000;40:2602-2609.
M
[50] Li G, Park CB. A crystallization kinetics study of polycarbonate under high-pressure
carbon dioxide and various crystallization temperatures by using magnetic suspension
d

balance. J Appl Polym Sci 2010;118:2898-2903.


e

[51] Oliveira NS, Dorgan J, Coutinho JAP, Ferreira A, Daridon JL, Marrucho IM. Gas
solubility of carbon dioxide in poly(lactic acid) at high pressures. J Polym Sci Part B
pt

Polym Phys 2006;44:1010-1019.


[52] Aionicesei E, Skerget M, Knez Z. Measurement of CO2 solubility and diffusivity in
ce

poly(l-lactide) and poly(d,l-lactide-co-glycolide) by magnetic suspension balance. J


Supercrit Fluids 2008;47:296-301.
Ac

[53] Mahmood SH, Ameli A, Hossieny N, Park CB. The interfacial tension of polylactide acid
in supercritical carbon dioxide. J Chem Thermodyn 2014; in press.
[54] Takada M, Hasegawa S, Ohshima M. Crystallization kinetics of poly(l-lactide) in contact
with pressurized CO2. Polym Eng Sci 2004;44:186-196.
[55] Yu L, Liu H, Dean K. Thermal behaviour of poly(lactic acid) in contact with compressed
carbon dioxide. Polym Int 2009;58:368-372.

41 
 

Page 41 of 63
[56] Li D, Liu T, Zhao L, Lian X, Yuan W. Foaming of poly(lactic acid) based on its
nonisothermal crystallization behavior under compressed carbon dioxide. Ind Eng Chem
Res 2011;50:1997-2007.
[57] Yu L, Liu H, Dean K, Chen L. Cold crystallization and post-melting crystallization of
PLA plasticized by compressed carbon dioxide. J Polym Sci Part B Polym Phys

t
2008;46:2630-2636.

ip
[58] Reignier J, Tatibouet J, Gendron R. Effect of dissolved carbon dioxide on the glass
transition and crystallization of poly(lactic acid) as probed by ultrasonic measurements. J

cr
Appl Polym Sci 2009;112:1345-1355.
[59] Nofar M, Zhu W, Park CB. Effect of dissolved CO2 on the crystallization behavior of

us
linear and branched PLA. Polymer 2012;53:3341-3353.
[60] Nofar M, Tabatabaei A, Park CB. Effects of nano-/micro-sized additives on the

[61]
an
crystallization behaviors of PLA and PLA/CO2 mixtures. Polymer 2013;54:2382-2391.
Nofar M, Tabatabaei A, Ameli A, Park CB. Comparison of melting and crystallization
behaviors of polylactide under high-pressure CO2, N2, and He. Polymer 2013;54:6471-
M
6478.
[62] Nofar M, Ameli A, Park CB. The thermal behavior of polylactide with different D-lactide
d

content in the presence of dissolved CO2. Macromol Mater Eng 2014;submitted.


e

[63] Avrami M. Kinetics of phase change. II transformation time relations for random
distribution of nuclei. J Chem Phys 1940;8:212-224.
pt

[64] Liao X, Nawaby AV, Whitfield PS. Carbon dioxide-induced crystallization in poly(L-
lactic acid) and its effect on foam morphologies. Polym Int 2010;59:1709-1718.
ce

[65] Kissinger HE. Variation of peak temperature with heating rate in different thermal
analysis. J Res Nat Bur Stand 1956;57:217-221.
Ac

[66] Lee ST, Leonard L, Jun J. Study of thermoplastic PLA foam extrusion. J Cell Plast
2008;44:293-305.
[67] Marrazzo C, Di Maio E, Iannace S. Conventional and nanometric nucleating agents in
poly(e-caprolactone) foaming: crystals vs. bubbles nucleation. Polym Eng Sci
2008;48:336-344.

42 
 

Page 42 of 63
[68] Wang C, Leung SN, Bussmann M, Zhai WT, Park CB. Numerical investigation of
nucleating agent-enhanced heterogeneous nucleation. Ind Eng Chem Res 2010;49:12783-
12792.
[69] Leung SN, Wong A, Wang C, Park CB. Mechanism of extensional stress-induced cell
formation in polymeric foaming processes with the presence of nucleating agents. J

t
Supercrit Fluids 2012;63:187-198.

ip
[70] Lips PAM, Velthoen W, Dijkstra PJ, Wessling M, Feijen J. Gas foaming of segmented
poly(ester amide) films. Polymer 2005;46:9396-9403.

cr
[71] Taki K, Kitano D, Ohshima M. Effect of growing crystalline phase on bubble nucleation
in poly(L-Lactide)/CO2 batch foaming. Ind Eng Chem Res 2011;50:3247-3252.

us
[72] Wong A, Guo Y, Park CB. Fundamental mechanisms of cell nucleation in polypropylene
foaming with supercritical carbon dioxide - effects of extensional stresses and crystals. J

[73]
Supercrit Fluids 2013;79:142-151. an
Wong A, Wijnands SFL, Kuboki T, Park CB. Mechanisms of nanoclay-enhanced plastic
foaming processes – effects of nanoclay intercalation and exfoliation. J Nanoparticle Res
M
2013;15:1815-1829.
[74] Colton JS, Suh NP. The nucleation of microcellular thermoplastic foam with additives:
d

Part I: Theoretical considerations. Polym Eng Sci 1987;27:485-492.


e

[75] Colton JS, Suh NP. The nucleation of microcellular thermoplastic foam with additives:
Part II: Experimental results and discussion. Polym Eng Sci 1987;27:493-499.
pt

[76] Kolstad JJ. Crystallization kinetics of poly(l-lactide-co-mesolactide). J Appl Polym Sci


1996;62:1079-1091.
ce

[77] Li H, Huneault MA. Effect of nucleation and plasticization on the crystallization of


poly(lactic acid). Polymer 2007;48:6855-6866.
Ac

[78] Tsuji H, Takai H, Saha SK. Isothermal and non-isothermal crystallization behavior of
poly(l-lactic acid): effects of stereocomplex as nucleating agent. Polymer 2006;47:3826-
3837.
[79] Pei A, Zhou Q, Berglund LA. Functionalized cellulose nanocrystals as biobased
nucleation agents in poly(l-lactide) (PLLA)- crystallization and mechanical property
effects. Comp Sci Technol 2010;70:815-821.

43 
 

Page 43 of 63
[80] De Santis F, Pantani R, Titomanlio G. Nucleation and crystallization kinetics of
poly(lactic acid). Thermochim Acta 2011;522:128-134.
[81] Park CB, Nofar M. A method for the preparation of PLA bead foams. US Pat PCT appl
PCT/ 050231, filed on 28 March 2013.
[82] Nofar M. Expanded PLA Bead Foaming: Analysis of Crystallization Kinetics and

t
Development of a Novel Technology. PhD thesis University of Toronto, 2013. 301 pp.

ip
[83] Sasaki H, Ogiyama K, Hira A, Hashimoto K, Tokoro H. Production method of foamed
polypropylene resin beads. US Pat 6,838,488 B2 2005.

cr
[84] Braun F. Method for producing expanded or expandable polyolefin particles. US Pat
6,723,760 B2 2004.

us
[85] Choi JB, Chung MJ, Yoon JS. Formation of double melting peak of poly(propylene-co-
ethylene-co-1-butene) during the preexpansion process for production of expanded

[86]
an
polypropylene. Ind Eng Chem Res 2013;44:2776-2780.
Nofar M, Guo Y, Park CB. Double crystal melting peak generation for expanded
polypropylene bead foam manufacturing. Ind Eng Chem Res 2013;52:2297-2303.
M
[87] Corre YM, Maazouz A, Duchet J, Reignier J. Batch foaming of chain extended PLA with
supercritical CO2: Influence of the rheological properties and the process parameters on
d

the cellular structure. J Supercrit Fluids 2011;58:177-188.


e

[88] Marrazzo C, Di Maio E, Iannace S. Foaming of synthetic and natural biodegradable


polymers. J Cell Plast 2007;43:123-133.
pt

[89] Pilla S, Kim SG, Auer GK, Gong S, Park CB. Microcellular extrusion-foaming of
polylactide with chain-extender. Polym Eng Sci 2009;49:1653-1660.
ce

[90] Pilla S, Kramschuster A, Yang L, Lee J, Gong S, Turng LS. Microcellular injection-
molding of polylactide with chain-extender. Mater Sci Eng C 2009;29:1258-1265.
Ac

[91] Bigg DM. Polylactide copolymers: effect of copolymer ratio and end capping on their
properties. Adv Polym Tech 2005;24:69-82.
[92] Garancher JP, Fernyhough A. Crystallinity effects in polylactic acid-based foams. J Cell
Plast 2012;48:387-397.
[93] Fujiwara T, Yamaoka T, Kimura Y, Wynne KJ. Poly(lactide) swelling and melting
behavior in supercritical carbon dioxide and postventing porous material.
Biomacromolecules 2005;6:2370-2373.

44 
 

Page 44 of 63
[94] Preechawong D, Peesan M, Supaphol P, Rujiravanit R. Preparation and characterization
of starch/poly(L-lactic acid) hybrid foams. Carbohydr Polym 2005;59:329-337.
[95] Hao A, Geng Y, Xu Q, Lu Z, Yu L. Study of different effects on foaming process of
biodegradable pla/starch composites in supercritical/compressed carbon dioxide. J Appl
Polym Sci 2008;109:2679-2686.

t
[96] Mihai M, Huneault MA, Favis BD, Li H. Extrusion foaming of semi-crystalline PLA and

ip
pla/thermoplastic starch blends. Macromol Biosci 2007;7:907-920.
[97] Zhang JF, Sun X. Biodegradable foams of poly(lactic acid)/starch. I. extrusion condition

cr
and cellular size distribution. J Appl Polym Sci 2007;106:857-862.
[98] Zhang JF, Sun X. Biodegradable foams of poly(lactic acid)/starch. II. cellular structure

us
and water resistance. J Appl Polym Sci 2007;106:3058-3062.
[99] Yuan H, Liu Z, Ren J. Preparation, characterizationand foaming behavior of poly(lactic
an
acid)/poly(butylene adipate-co-butylene terephthalate) blend. Polym Eng Sci
2009;49:1004-1012.
[100] Kramschuster A, Turng LS. An injection molding process for manufacturing highly
M
porous and interconnected biodegradable polymer matrices for use as tissue engineering
scaffolds. J Biomed Mater Res Part B 2010;92:366-376.
d

[101] Pilla S, Kim SG, Auer GK, Gong S, Park CB. Microcellular extrusion foaming of
e

poly(lactide)/poly(butylene adipate-co-terephthalate) blends. Mater Sci Eng C


2010;30:255-262.
pt

[102] Li K, Cui Z, Sun X, Turng LS, Huang HX. Effects of nanoclay on the morphology and
physical properties of solid and microcellular injection molded polyactide/poly(butylenes
ce

adipate-co-terephthalate) (PLA/PBAT) nanocomposites and blends. J Biobased Mater Bio


2011;5:442-451.
Ac

[103] Zhao HB, Cui Z, Sun X, Turng LS, Peng XF. The morphology and properties of injection
molded solid and microcellular polylactic acid/polyhydroxybutyrate-valerate
(PLA/PHBV) blends. Ind Eng Chem Res 2013;52:2569-2581.
[104] Kramschuster A, Pilla S, Gong S, Chandra A, Turng LS. Injection molded solid and
microcellular polylactide compounded with recycled paper shopping bag fibers. Int
Polym Process 2007;22:436-445.

45 
 

Page 45 of 63
[105] Pilla S, Kramschuster A, Lee J, Auer GK, Gong S, Turng LS. Microcellular and solid
polylactide-flax fiber composites. Compos Interfaces 2009;16:869-890.
[106] Kang DJ, Xu D, Zhang ZX, Pal K, Bang DS, Kim JK. Well-controlled microcellular
biodegradable PLA/silk composite foams using supercritical CO2. Macromol Mater Eng
2009;294:620-624.

t
[107] Matuana LM, Faruk O. Effect of gas saturation conditions on the expansion ratio of

ip
microcellular poly(lactic acid)/wood-flour composites. Express Polym Lett 2010;4:621-
631.

cr
[108] Boissard CI, Bourban PE, Plummer CJG, Neagu RC, Manson JAE. Cellular
biocomposites from polylactide and microfibrillated cellulose. J Cell Plast 2012;48:445-

us
458.
[109] Park CB, Baldwin DF, Suh NP. Effect of the pressure drop rate on cell nucleation in
an
continuous processing of microcellular polymers. Polym Eng Sci 1995;35:432-440.
[110] Matuana LM, Park CB, Balatinecz JJ. Cell morphology and property relationships of
microcellular foamed PVC/wood-fiber composites. Polym Eng Sci 1998;38:1862-1872.
M
[111] Rachtanapun P, Selke SEM, Matuana LM. Relationship between cell morphology and
impact strength of microcellular foamed high density polyethylene/polypropylene blends.
d

Polym Eng Sci 2004;44:1551-1560.


e

[112] Shimbo M, Higashitani I, Miyano Y. Mechanism of Strength Improvement of Foamed


Plastics Having Fine Cell. J Cell Plast 2007;43:157-167.
pt

[113] Seeler KA, Kumar V. Tension-tension fatigue of microcellular polycarbonate: initial


results. J Reinforced Plast Comp 1993;12:359-376.
ce

[114] Hilyard NC, Cunningham A, editors. Low Density Cellular Plastics: Physical Basis of
Behaviour. London: Chapman and Hall, 1994. 369 pp.
Ac

[115] Martínez-Díez JA, Rodríguez-Pérez MA, de Saja JA, Arcos y Rábago LO, Almanza OA.
The thermal conductivity of a polyethylene foam block produced by a compression
molding process. J Cell Plast 2001;37:21-42.
[116] Solórzano E, Rodriguez-Perez MA, Lázaro J. J.A. de Saja; Influence of Solid
Phase Conductivity and Cellular Structure on the Heat Transfer Mechanisms of Cellular
Materials: Diverse Case Studies. Adv Eng Mater 2009;11:818-824.

46 
 

Page 46 of 63
[117] Ahmed MS. Design and Manufacturing of Novel Microcellular Acoustical Foams. PhD
thesis, University of Toronto, 2007. 178 pp.
[118] Kabumoto A, Nakayama K, Ito M, Ono S, Yoshida N. Method for manufacturing a
foamed plastics of saturated polyester using a cyclic tetramer as foaming agent. US Pat
5,458,832 1995.

t
[119] Kabumoto A, Yoshida N, Ito M, Okada M. Method of manufacturing thermoplastic

ip
polyester foam sheet. US Pat 5,723,510 1998.
[120] Okamoto M, Nam PH, Maiti P, Kotaka T, Nakayama T, Takada M, Ohshima M, Usuki

cr
A, Hasegawa N, Okamoto H. Biaxial flow-induced alignment of silicate layers in
polypropylene/clay nanocomposite foam. Nano Lett 2001;1:503-505.

us
[121] Srithep Y, Turng LS, Sabo R, Clemons C. Nanofibrillated cellulose (NFC) reinforced
polyvinyl alcohol (pvoh) nanocomposites: properties, solubility of carbon dioxide, and
foaming. Cellulose 2012;19:1209-1223. an
[122] Wong A, Park CB. The effects of extensional stresses on the foamability of polystyrene-
talc composites blown with carbon dioxide. Chem Eng Sci 2012;75:49-62.
M
[123] Lee YH, Wang KH, Park CB, Sain M. Effects of clay dispersion on the foam morphology
of LDPE/clay nanocomposites. J Appl Polym Sci 2007;103:2129-2134.
d

[124] Zheng W, Lee YH, Park CB. Use of nanoparticles for improving the foaming behaviors
e

of linear PP. J Appl Polym Sci 2010;117:2972-2979.


[125] Shen J, Zeng CC, Lee LJ. Synthesis of polystyrene] carbon nanofibers nanocomposite
pt

foams. Polymer 2005;46:5218-5224.


[126] Kharbas H, Nelson P, Yuan M, Gong S, Turng LS. Effects of nano-fillers and process
ce

conditions on the microstructure and mechanical properties of microcellular injection


molded polyamide nanocomposites. Polym Compos 2003;24:655-671.
Ac

[127] Han X, Zeng C, Lee LJ, Koelling KW, Tomasko DL. Extrusion of polystyrene
nanocomposite foams with supercritical CO2. Polym Eng Sci 2003;43:1261-1275.
[128] Zhu Z, Park CB, Zong J. Challenges to the formation of nano cells in foaming processes.
Int Polym Process 2008;23:270-276.
[129] Thiagarajan C, Sriraman R, Chaudhari D, Kumar M, Pattanayak A. Nano-cellular
polymer foam and methods for making them. US Pat 7,838,108 B2 2010.

47 
 

Page 47 of 63
[130] Fujimoto Y, Ray SS, Okamoto M, Ogami A, Yamada K, Ueda K. Well-Controlled
biodegradable nanocomposite foams: from microcellular to nanocellular. Macromol
Rapid Commun 2003;24:457-461.
[131] Ema Y, Ikeya M, Okamoto M. Foam processing and cellular structure of polylactide-
based nanocomposites. Polymer 2006;47:5350-5359.

t
[132] Pilla S, Kramschuster A, Gong S, Chandra A, Turng LS. Solid and microcellular

ip
polylactide-carbon nanotube nanocomposites. Int Polym Process 2007;5:418-428.
[133] Kramschuster A, Gong S, Turng LS, Li T, Li T. Injection molded solid and microcellular

cr
polylactide and polylactide nanocomposites. J Biobased Mater Bio 2007;1:37-45.
[134] Bitou M, Okamoto M. Fabrication of porous 3-D structure from poly(L-lactide)-based

us
nanocomposite foam via enzymatic degradation. Int Polym Process 2007;5:1-9.
[135] Bitou M, Okamoto M. Fabrication of porous 3-D structure from poly(L-lactide)-based
an
nano-composite foams. Effect of foam structure on enzymatic degradation. Polym Degrad
Stab 2008;93:1081-1087.
[136] Hwang SS, Hsu PP, Yeh JM, Chang KC, Lai YZ. The mechanical/thermal properties of
M
microcellular injection-molded poly-lactic-acid nanocomposites. Polym Compos
2009;30:1625-1630.
d

[137] Pilla S, Kramschuster A, Lee J, Clemons C, Gong S, Turng LS. Microcellular processing
e

of polylactide-hyperbranched polyester-nanoclay composites. J Mater Sci 2010;45:2732-


2746.
pt

[138] Matuana LM, Diaz CA. Study of cell nucleation in microcellular poly(lactic acid) foamed
with supercritical CO2 through a continuous-extrusion process. Ind Eng Chem Res
ce

2010;49:2186-2193.
[139] Tsimpliaraki A, Tsivintzelis I, Marras SI, Zuburtikudis I, Panayiotou C. The effect of
Ac

surface chemistry and nanoclay loading on the microcellular structure of porous poly(d,l
lactic acid) nanocomposites. J Supercrit Fluids 2011;57:278-287.
[140] Ameli A, Jahani D, Nofar M, Park CB. Development of high void fraction polylactide
composite foams using injection molding: Mechanical and thermal insulation properties.
Comp Sci Technol 2013;90:88-95.
[141] Mikos AG, Lyman MD, Freed LE, Langer R. Wetting of poly(L-lactic acid) and
poly(DL-lactic-co-glycolic acid) foams for tissue culture. Biomaterials 1994;15:55-58.

48 
 

Page 48 of 63
[142] Lo H, Kadiyala S, Guggino SE, Leong KW. Poly (L-lactide acid) foams with cell seeding
and controlled-release capacity. J Biomed Mater Res 1996;30:475-484.
[143] Goldstein AS, Zhu G, Morris G, Meszlenyi RK, Mikos AG. Effect of osteoblastic culture
conditions on the structure of poly(DL-lactic-co-glycolic acid) foam scaffolds. Tissue
Eng 1999;5:421-433.

t
[144] Zhang R, Ma PX. Porous poly(L-lactic acid)/apatite composites created by biomimetic

ip
process. J Biomed Mater Res 1999;45:285-293.
[145] Nam YS, Park TG. Biodegradable polymeric microcellular foams by modified thermally

cr
induced phase separation method. Biomaterials 1999;20:1783-1790.
[146] Nam YS, Yoon JJ, Park TG. A novel fabrication method of macroporous biodegradable

us
polymer scaffolds using gas foaming salt as a porogen additive. J Biomed Mater Res
2000;53:1-7.
an
[147] Mikos AG, Temenoff JS. Formation of highly porous biodegradable scaffolds for tissue
engineering. J Biotechnol 2000;3:1-6.
[148] Chen G, Ushida T, Tateishi T. Preparation of poly(L-lactic acid) and poly(DL-lactic-co-
M
glycolic acid) foams by use of ice microparticulates. Biomaterials 2001;22:2563-2567.
[149] Yang J, Shi G, Bei J, Wang S, Cao Y, Shang Q, Yang G, Wang W. Fabrication and
d

surface modification of macroporous poly(L-lactic acid) and poly(L-lactic-co-glycolic


e

acid) (70/30) cell scaffolds for human skin fibroblast cell culture. J Biomed Mater Res
2002;62:438-446.
pt

[150] Tu C, Cai Q, Yang J, Wan Y, Bei J, Wang S. The fabrication and characterization of
poly(lactic acid) scaffolds for tissue engineering by improved solid–liquid phase
ce

separation. Polym Adv Technol 2003;14:565-573.


[151] Lee JH, Park TG, Park HS, Lee DS, Lee YK, Yoon SC, Nam JD. Thermal and
Ac

mechanical characteristics of poly(l-lactic acid) nanocomposite scaffold. Biomaterials


2003;24:2773-2778.
[152] Karageorgiou V, Kaplan D. Porosity of3D biomaterial scaffolds and osteogenesis.
Biomaterials 2005;26:5474-5491.
[153] Jung Y, Kim SS, Kim YH, Kim SH, Kim BS, Kim S, Choi CY, Kim SH. A poly(lactic
acid)/calcium metaphosphate composite for bone tissue engineering. Biomaterials
2005;26:6314-6322.

49 
 

Page 49 of 63
[154] Mooney DJ, Baldwin DF, Suh NP, Vacantis JP, Larger R. Novel approach to fabricate
porous sponges of poly(D,L-lactic-co-glycolic acid) without the use of organic solvents.
Biomaterials 1996;17:1417-1422.
[155] Montjovent M, Mathieu L, Hinz B, Applegate LL, Bourban PE, Zambelli PY, Månson
JA, Pioletti DP. Biocompatibility of bioresorbable poly(l-lactic acid) composite scaffolds

t
obtained by supercritical gas foaming with human fetal bone cells. Tissue Eng

ip
2005;11:1640-1649.
[156] Tsivintzelis I, Pavlidou E, Panayiotou C. Porous scaffolds prepared by phase inversion

cr
using supercritical CO2 as antisolvent I. Poly(l-lactic acid). J Supercrit Fluids
2007;40:317-322.

us
[157] Zhu XH, Lee LY, Jackson JSH, Tong YW, Wang CH. Characterization of porous
poly(d,l-lactic-co-glycolic acid) sponges fabricated by supercritical CO2 gas-foaming
an
method as a scaffold for three-dimensional growth of Hep3B cells. Biotechnol Bioeng
2008;100:998-1009.
[158] Gendron R, Champagne MF. Rheological Behavior Relevant to Extrusion Roaming. In:
M
Gendron R, editor. Thermoplastic Foam Processing: Principles and Development. Bocca
Raton FL: CRC Press, 2005. p. 42-104
d

[159] Reignier J, Gendron R, Champagne MF. Extrusion foaming of poly(lactic acid) blown
e

with CO2: Toward 100% green material. Cell Polym 2007;26:83-115.


[160] Nalawade S, Picchioni F, LPBM Janssen. Supercritical carbon dioxide as a green solvent
pt

for processing polymer melts: processing aspects and applications. Prog Polym Sci
2006;31:19-43.
ce

[161] Matuana LM, Faruk O, Diaz CA. Cell morphology of extrusion foamed poly(lactic acid)
using endothermic chemical foaming agent. Bioresour Technol 2009;100:5947-5954.
Ac

[162] Lee JWS, Park CB. Use of nitrogen as a blowing agent for the production of fine-celled
high-density HDPE foams. Macromol Mater Eng 2006;291:1233-1244.
[163] Lee JWS, Park CB, Kim SG. Reducing material costs with microcellular/fine-celled
foaming. J Cell Plast 2007;43:297-312.
[164] Lee JWS, Wang J, Yoon JD, Park CB. Strategies to achieve a uniform cell structure with
a high void fraction in advanced structural foam molding. Ind Eng Chem Res
2009;47:9457-9464.

50 
 

Page 50 of 63
[165] Wong S, Lee JWS, Naguib HE, Park CB. Effect of processing parameters on the
mechanical properties of injection molded thermoplastic polyolefin (TPO) cellular foams.
Macromol Mater Eng 2008;293:605-613.
[166] Guo Q. Visualization of Polymer Foaming Using a Batch Foaming Simulation System
with a High Pressure-Drop Rate. PhD thesis University of Toronto, 2007. 202 pp.

t
[167] Kwag C, Manke CW, Gulari E. Effects of dissolved gas on viscoelastic scaling and glass

ip
transition temperature of polystyrene melts. Ind Eng Chem Res 2001;40:3048-3052.
[168] Kwag C, Manke CW, Gulari E. Rheology of molten polystyrene with dissolved

cr
supercritical and near-critical gases. J Polym Sci Part B Polym Phys 1999;37:2771-2781.
[169] Suh NP. Microcellular plastics. In: Stevenson JF, editor. Innovation in Polymer

us
Processing: Molding. Munich: Hanser Publishers, 1996. p. 93-133.
[170] Anonymous, Trexel Inc. Sustainable solutions for engineered plastic parts
an
http://www.trexel.com 2012. Accessed Dec. 2013.
[171] Egger P, Fischer M, Kirschling H, Bledzki AK. Versatility for Mass Production in Mucell
Injection Moulding. Kunstst Plast Eur 2005;95:70-73.
M
[172] Sporrer ANJ, Altstadt V. Controlling morphology of injection molded structural foams by
mold design and processing parameters. J Cell Plast 2007;43:313-330.
d

[173] Ishikawa T, Ohshima M. Visual observation and numerical studies of polymer foaming
e

behavior of polypropylene/carbon dioxide system in a core-back injection molding


process. Polym Eng Sci 2011;51:1617-1625.
pt

[174] Ishikawa T, Taki K, Ohshima M. Visual observation and numerical studies of N2 vs.
CO2 foaming behavior in core-back foam injection molding. Polym Eng Sci 2012;52:875-
ce

883.
[175] Mills N. Bead Foam Microstructure and Processing. Chapt 4. Polymer Foams Handbook.
Ac

Oxford: Butterworth-Heinemann, 2007. p. 69-83.


[176] Parker K, Garancher JP, Shah S, Weal S, Fernyhough A. Polylactic acid (PLA) foams for
packaging applications. In: Pilla S, editor. Handbook of Bioplastics and Biocomposites
Engineering Applications. Salem MA: Scrivener Publishing LLC 2011. p. 161-175.
[177] Witt M, Shah S. Methods of manufacture of polylactide acid foams. US Pat US 8,283,389
B2 2012.

51 
 

Page 51 of 63
[178] Noordegraaf J, Kuijstermans FPA, De Jong JPM. Particulate, expandable polymer,
method for producing particulate expandable polymer, as well as a special use of the
obtained foam material. US Pat 20110218257 A1 2011.
[179] Britton RN, Van Doormalen FAHC, Noordegraaf J, Molenveld K, Schennink GGJ,
Kuijstermans FPA, Van Sas JC, De Jong JPM. Polymer blend containing polylactic acid

t
and a polymer having a Tg higher than 60°C foam. Eur Pat EP2137249 A2 2009.

ip
[180] Shinohara M, Tokiwa T, Sasaki H. Expanded polylactide acid resin beads and foamed
molding obtained therefrom. Eur Pat EP1378538 A1 2004.

cr
[181] Haraguchi K, Ohta H. Expanded polylactide acid resin particles. Eur Pat EP1683 828 A2
2005.

us
 

Figure 1.
an
Synthesis of polylactide: ring-opening polymerization [6-7].

Figure 2. SS-corrected volume swelling ratio of PLA 3001D–CO2 and PLA 3001D–N2 (a) and
M
PLA with varied D-content and CO2 (b) at 180◦C and Copyright 200◦C. Data
adapted with permission from [45, 48], Copyright 2006, Elsevier Ltd; Copyright
2013, Elsevier Ltd; respectively.
d

Figure 3. SS-corrected solubility of PLA 3001D–CO2 and PLA 3001D–N2 (a) and PLA with
e

varied D-content and CO2 (b) at 180◦C and Copyright 200◦C. Data adapted with
permission from [45, 48], Copyright 2006, Elsevier Ltd; Copyright 2013, Elsevier
pt

Ltd; respectively.

Figure 4. Diffusivity of three different N2 and CO2 pressures in PLA 3001D at 180oC. Data
ce

adapted with permission from [45], Copyright 2006, Elsevier Ltd.

Figure 5. CO2 solubility in the amorphous (80:Copyright 20) PLA and in the semi-crystalline
Ac

(98:2) PLA at 30.9◦C. Data adapted with permission from [51], Copyright 2006 ,
John Wiley & Sons.

Figure 6. Solubility of supercritical CO2 in PLLA at 35oC, 40oC, and 50oC. Data adapted with
permission from [52], Copyright 2008, Elsevier Ltd.

Figure 7. Effects of temperature and CO2 pressures on the interfacial tension of PLA 3001D.
Data adapted with permission from [53], Copyright 2014, Elsevier Ltd.

Figure 8. Crystallization rate at various isothermal melt crystallization temperatures. Data


adapted with permission from [59, 61-62], Copyright 2012, John Wiley & Sons;

52 
 

Page 52 of 63
Copyright 2013, John Wiley & Sons; Copyright 2013, John Wiley & Sons;
respectively.

Figure 9. Activation energy of PLA’s non-isothermal melt crystallization at various CO2 gas
pressures (a) [56],, Crystallization half-time at various CO2 pressures during a
cooling cycle (b). Data adapted with permission from [56, 59], Copyright 2011, the
American Chemical Society; Copyright 2012, John Wiley & Sons; respectively.

t
ip
Figure 10. Avrami n value exponent of the isothermally treated PLA samples. Data adapted
with permission from [54, 57, 59], Copyright 2004, John Wiley & Sons; Copyright
2008, John Wiley & Sons; Copyright 2012, John Wiley & Sons; respectively.

cr
Figure 11. Final degree of crystallinity in the PLA samples cooling at various rates and at

us
different CO2 gas pressures. Data adapted with permission from [56], Copyright
2011, the American Chemical Society.

Figure 12. Final degree of crystallinity of various types of PLA and PLA composite samples
an
cooled at a rate of 2oC/min. at different CO2 gas pressures. Data adapted with
permission from [59-62], Copyright 2012, John Wiley & Sons; Copyright 2013, John
Wiley & Sons; Copyright 2013, John Wiley & Sons; Copyright 2013, John Wiley &
M
Sons; respectively.

Figure 13. (a) Tc and (b) Tg variations of different types of PLA versus CO2 pressure. Data
adapted with permission from [54-57, 59, 62], Copyright 2004, John Wiley & Sons;
d

Copyright 2009, John Wiley & Sons; Copyright 2011, the American Chemical
Society; Copyright 2008, John Wiley & Sons; Copyright 2012, John Wiley & Sons;
e

Copyright 2013, John Wiley & Sons; respectively.


pt

Figure 14. Tm variations of different PLA materials versus CO2 pressure. Data adapted with
permission from [54-56, 61], Copyright 2004, John Wiley & Sons; Copyright 2009,
ce

John Wiley & Sons; Copyright 2011, the American Chemical Society; Copyright
2013, John Wiley & Sons; respectively.

Figure 15. Foam visualization of a PLLA/CO2 system under a polarized optical microscope.
Ac

The saturation pressure and temperature were 11 MPa and 180oC, respectively. The
crystallization and bubble nucleation occurred at 110oC after 6.1sec [71].

Figure 16. SEM micrographs and expansion ratios of the linear PLA (D-content=2) samples
with foamed at various CO2 content. Data adapted with permission from [39],
Copyright 2009, John Wiley & Sons.

Figure 17. SEM micrographs and expansion ratios of the branched PLA (D-content=2) samples
with 2% chain extender foamed at various CO2 content. Data adapted with
permission from [37], Copyright 2010, John Wiley & Sons.

53 
 

Page 53 of 63
Figure 18. Expansion ratios of branched PLA foams after 0 sec and 90 sec die residence time
for crystallization. Data adapted with permission from [38], Copyright 2012, Elsevier
Ltd.

Figure 19. Expansion ratios of PLA and PLA nanocomposite foams while varying the
temperature profile of the second extruder.

t
Figure 20. Maximum expansion ratios (a) and cell density (b) of the PLA foams with varying

ip
CE content. Data adapted with permission from [37-38, 89], Copyright 2010, John
Wiley & Sons; Copyright 2012, Elsevier Ltd; Copyright 2009, John Wiley & Sons;
respectively.

cr
Figure 21. Cell morphology of the injected PLA foamed samples with (a) 0 wt% and (b) 0.8

us
wt% CE. Data adapted with permission from [90], Copyright 2009, Elsevier Ltd.

Figure 22. Expansion ratios (a) and foam crystallinity (b) of PLA foams with varying D-lactide
content. Data adapted with permission from [39], Copyright 2009, John Wiley &
Sons. an
 
M
 

Table 1. Studies implemented on PLA blends. Data adapted from [94-103]


d

Polymer Minimum
Maximum
Blend Foam Average
Researcher Blowing Agent Expansion Analysis Techniques
e

Processing Cell size


Base Ratio ~
~
pt

Preechawong
PLLA/ Compression
et al. (2005) Water 6 - Tensile & Flexural Tests
Starch Molding
ce

[94]

Hao et al. PLA/


Batch CO2 13 5 µm Gas Sorption/ DSC
(2008) [95] Starch
Ac

Mihai et al. PLA/ Rheology/Solubility/


Extrusion CO2 50 25 µm
(2007) [96] Starch DSC/XRD

Zhang et al. PLA/


Extrusion Water 55 500 µm Compression Test/ DSC
(2007) [97, 98] Starch

PLA/
Yuan et al.
PBAT+ Extrusion (Azodicarbonamide) 3.2 10-20 µm FTIR/Rheology/ DSC
(2009) [99]
nanosilica

Kramschuster PLA/PVOH Injection + CO2 4 200 µm TGA


t l (2010) + lt l hi
54 
 

Page 54 of 63
et al. (2010) + salt leaching
[100]

PLA/
Pilla et al.
PBAT + Extrusion CO2 1.8 10 µm DSC
(2010) [101]
talc

PLA/
Li et al. XRD/TEM/DSC/Tensile

t
PBAT + Injection N2 1.05 25 µm
(2011) [102] Test

ip
nanoclay

Zhao et al. PLA/ Rheology/DSC/TGA/


Injection N2 - 25 µm

cr
(2013) [103] PHBV DMA/Tensile test

us
an
M
e d
pt
ce
Ac

55 
 

Page 55 of 63
 
Table 2. Studies made of PLA composites. Data adapted from [42, 104-108]
PLA composites

Maximum Minimum
Polymer Foam Blowing
Researcher Expansion Average Analysis Techniques
Composite Processing Agent
Ratio ~ Cell size ~

t
PLA/ Recycled
Kramschuster et DSC/ DMA/ Tensile

ip
Paper Shopping Injection N2 1.2 3 µm
al. (2007) [104] Test
Bag Fibers

cr
Pilla et al. DSC/ DMA/ Tensile
PLA/ Flax Fiber Injection N2 1.25 3 µm
(2009) [105] Test

us
Kang et al. PLA/ Silk Fibroin
Batch CO2 - 15 µm DSC/ XRD
(2009) [106] powder

Matuana et al.
PLA/ Wood Flour Batch CO2 10 35 µm Gas Sorption
(2010) [107]

Boissard et al.
PLA/
an
microfibrillated Batch CO2 6.8 200 µm Compression Test
(2012) [108]
cellulose (MFC)
M
Ameli et al.
PLA/ talc Injection N2 1.4 20 µm DSC/ Tensile Test
(2013) [42]
d

 
e
pt
ce
Ac

56 
 

Page 56 of 63
 

Table 3. Studies implemented on PLA nanocomposites. Data adapted from [25, 31, 40, 88, 130-140]
PLA nanocomposites

Minimum
Maximum
Polymer Foam Blowing Average
Researcher Expansion Analysis Techniques
Nanocomposite Processing Agent Cell size
Ratio ~

t
~

ip
Fujimoto et al.
(2003) [130], Rheology/DSC/

cr
PLA/Nanoclay Batch CO2 2.7 360 nm
Ray et al. XRD/TEM/GPC
(2003) [31]

us
Di et al.
(2005) [25]
Rheology/DSC/ XRD/
PLA/Nanoclay Batch CO2/N2 13.5 25 µm
Marrazzo et FTIR
al. (2007) [88]

Ema et al.
an
PLA/Nanoclay Batch CO2 5 200 nm Gas Sorption/ TEM
(2006) [131]
M
Pilla et al. PLA/Carbon Tensile Test/DMA/
Injection N2 1.28 40 µm
(2007) [132] Nanotube TEM/ DSC
d

Kramschuster PLA/Nanoclay
et al. (2007) (Cloisite 20A & Injection N2 1.4 13 µm XRD/TEM/ Tensile Test
e

[133]
30B)
pt

Bitou et al.
(2007, 2008) PLLA/Nanoclay Batch CO2 3.1 165 nm GPC/DSC
[134-135]
ce

Hwang et al. XRD/TEM/TGA/


PLA/Nanoclay Injection CO2 1.35 35 µm
(2009) [136] DSC/Tensile Test
Ac

Pilla et al. PLA/HBP XRD/TEM/DSC/ Tensile


Injection N2 1.2 10 µm
(2010) [137] Nanoclay Test/DMA

Matuana et al.
PLA/Nanoclay Extrusion CO2 - 6.9 µm Rheology
(2010) [138]

Tsimpliaraki

et al. (2011) PDLLA/Nanoclay Batch CO2 6.2 6 µm XRD


[139]

Li et al. PLA/ PBAT + Injection N2 1.05 25 µm XRD/TEM/DSC/Tensile


(2011) [102] l T t

57 
 

Page 57 of 63
(2011) [102] nanoclay Test

Nofar et al. XRD/TEM/Solubility/


PLA/Nanoclay Extrusion CO2 44 40 µm
(2013) [40] DSC/Rheometer

Ameli et al. XRD/TEM/DSC/Flexural


PLA/Nanoclay Injection N2 2.85 38 µm
(2013) [140] & Impact Tests

t
 

ip
 

cr
us
an
M
e d
pt
ce
Ac

58 
 

Page 58 of 63
ep
c
Ac
Page 59 of 63
c
Ac Page 60 of 63
Ac Page 61 of 63
p
ce
Ac
Page 62 of 63
an
M
ed
pt
ce
Ac

Page 63 of 63

You might also like