SlideShare ist ein Scribd-Unternehmen logo
1 von 16
Downloaden Sie, um offline zu lesen
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 1
Journal of Polymer & Composites
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
Volume 4, Issue 2
www.stmjournals.com
Depolymerization of Polypropylene: A Systematic Review
Parag Kulkarni*, Ruchika Pache
Department Polymer and Surface Engineering, Institute of Chemical Technology, Mumbai-400019,
Maharashtra, India
Abstract
Polypropylene have become an indispensable ingredient of human life and most widely used
thermoplastic in commodity applications. Polymerization of propylene is carried out by using
Ziegler-Natta or Metallocene catalysts which results in high molecular weights and
crystalline polypropylene which are not required for several applications. Reduction in
molecular weight of polypropylene in conventional reactors is uneconomical and complicated.
Therefore to overcome this problem researchers and technologists proposed a post reactor
operation to reduce molecular weight (depolymerization) of polypropylene. To induce
modification in polypropylene different techniques of depolymerization are proposed. In this
review article, we have discussed various types of polypropylene depolymerization methods
with their mechanisms like oxidative, thermal, high energy radiation and chemical (free
radical induced) depolymerization along with their effect on thermal, rheological, crystalline
and chemical properties of the polypropylene.
Keywords: Polypropylene, molecular weight, depolymerization, thermal depolymerization,
rheology
*Author for Correspondence E-mail: kulkarniparag05@gmail.com
INTRODUCTION
Polyolefins are the most widely used polymers
for manufacturing of commodity plastics
article. They are manufactured usually by
using transition metal catalysts and have very
high molecular weights. Polypropylene is an
enormously adaptable member of the
polyolefin family, which is due to the
prochiral nature of the propylene monomer.
Highly crystalline polypropylene is built up by
the chain growth polymerization of propylene,
a gaseous compound obtained by the thermal
cracking of ethane, propane and butane or
naptha fraction of petroleum. In 1951
scientists at Phillips Robert L Banks et al.
found that one of their processes of converting
propylene into gasoline led to the formation of
whitish substance called crystalline
polypropylene. Ziegler modified the
equipment and prepared polypropylene in
1954. In 1954 Natta managed to synthesize
polypropylene, followed shortly by Ziegler.
The first industrially important crystalline,
high molecular weight polypropylene was
synthesized by Natta in 1955 from organo-
metallic catalysts based on titanium and
aluminum [1]. Polypropylene was developed
by scientists at Phillips, Hoechst, Montecatini,
Hercules, Farbwerke-Hoechst, by Ziegler and
Natta and in separate development by standard
oil of Indiana, Du Pont in 1957 [2].
Polypropylene has been widely used as
injection molding plastics as well as in many
extruded forms including film, fiber and split
yarns [3].
Polypropylene is one of the world’s most
important thermoplastic materials and it is
used in numerous applications in the plastics
industry. It has desirable properties such as
low-cost, high-melting point, low density,
appropriate process ability, high-strength, high
stiffness, hinge properties, easy to handle and
excellent chemical resistance [4]. Its
applications have expanded continuously for
the last decade. Since commodity plastics do
not require extremely high mechanical
properties, molecular weights and crystallinity
always, highly crystalline isotactic
polypropylene has been reported to be having
very high melt viscosity with consequently
poor flow properties and deficient in impact
strength at low temperatures, mainly owing to
its relatively high Tg and large spherulite
dimensions. It’s moderately elevated glass
transition temperature results in it being too
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 2
brittle and hard at application temperatures
below 0ºC. These high molecular weights of
polypropylene give resistance to flow at
elevated temperatures and hence operations of
shaping by compression molding or injection
molding or extrusion are unnecessarily
complex. The toughness and rheological
properties of polypropylene can be enhanced
by reducing the molecular weight, molecular
weight distribution, introduction of branching
in polymer backbone and controlling average
size of the spherulites in depolymerization.
Control of molecular weight in the reactor,
during the process of polymerization is
difficult. The process parameters are set and
any modification in the process like addition
of a chain transfer agent would result in
resetting the process, which in turn is very
tedious, difficult and economically not
feasible. To achieve the diversity in
polypropylene grades suitable for the different
applications the molecular weight and
molecular weight distribution (MWD) must be
tailor-made to fit the performance
requirements of each application.
DEPOLYMERIZATION OF
POLYPROPYLENE
Depolymerization leads to changes in
molecular properties such as molecular
weight, molecular weight distribution, thermal
properties, and crystallisability, etc. which are
considered to be more beneficial in mechanical
properties. This foregoing defects and
deficiencies can be overcome by
depolymerization process for reducing the
molecular weight of the polypropylene.
Polypropylene is depolymerized to achieve
easy processing for high-quality products in
injection molding and fiber spinning with
enhancing mechanical properties [5–7].
Depolymerization is the reversion of a
polymer to a low molecular weight polymer
fragments or splitting of polymers into
molecules [8]. Reverse of the propagation step
in chain polymerization is depolymerization or
unzipping, which is characterized by reduction
in molecular weight of the polymer to low
molecular weight polymer fragments [9].
Depolymerization of polymer may be brought
about by physical factor such as heat, light,
mechanical stress, high energy radiation, and
ultrasonication or by chemical agent such as
initiator, oxygen, ozone etc. Depolymerization
of the polymer is subdivided according to its
various modes of initiation. There are
oxidative, thermal, high energy radiation and
chemical (free radical induced) modes for
depolymerization of polypropylene. Lanrence
has patented a process which gives a
procedure of controlled thermal
depolymerization of a crystalline polymer of
an aliphatic mono-a-olefin which comprises
treating the polymer as slurry with water,
containing oxygen and including free radicals,
at a temperature of 60 to 90o
C and recovering
from the slurry, a polymer of increased melt
index [10]. The depolymerization of 16.7
weight % slurry of isotactic polypropylene
dispersed in water was carried out in presence
of free radical initiator and it was found that in
the presence of 0.2% benzoyl peroxide at
95°C, the melt flow index increased to 13 from
9 after 1 h of reaction. Similarly in the
presence of 0.3% AIBN the melt flow index
increased from 0.1 to 1.2 after 2 h of reaction.
It is important to note that this reduction in
polymer is brought about by molecular chains
scission of polypropylene. Hans et al. gave a
process for reduction in border viscosities of
the polymer, by treating polypropylene in
fluidized bed, in presence of oxygen, within
temperature range of 100–130°C. Powdered
polypropylene, of border viscosity 6, was
treated in an eddy furnace for 40 min and at
temperature of 130°C. The resulting polymer
had a border viscosity of 3 [11]. The Dow
Chemical company has patented a process for
reduction in molecular weight of isotactic
polypropylene to improve its proccessability
[12]. The process comprises of thermally
depolymerizing polypropylene at temperatures
of 160 to 300°C, in the presence of small
amount of depolymerizing agent such as
bromine or bromine containing compound (to
result in a bromide radical). The treated
polypropylene had a melt index improved by
2.2. Davis et al. subjected polypropylene
samples to various degrees of thermal
depolymerization [13]. The molecular weight
distributions of these samples were
experimentally determined and compared with
those expected theoretically for random
scission of the polymer chains. The
comparison confirmed that the chain breakage
was predominately random and also indicated
that determination of molecular weight by
viscosity average molecular weight were
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 3
adequate for use in evaluating hypothesized
depolymerization mechanisms. The most
common types of depolymerization occur
through chemical reactions that alter the
molecular weight of the polypropylene,
leading to a change in its mechanical, thermal,
chemical and rheological properties. Such
reactions include (a) chain scission, (b) cross-
linking, (c) modification of branched chains,
or (d) a combination of all of these reactions.
The agent(s) initiating the depolymerization
process defines the type of depolymerization.
A review of the different types of
depolymerizations and their initiating agents
are given in Table 1.
Table 1: Initiating Agents and
Depolymerization Methods.
Initiating Agent Depolymerization Type
Heat Thermal
Oxygen, Ozone Oxidative
X-rays, γ-rays, electron
beam radiation
High energy radiation
Free radical induced Chemical
The chemical reactions in depolymerization
process are comparatively similar, with barely
trivial differences owing to variations in the
initiation mechanism [14].
Depolymerization Mechanism of
Polypropylene
Depolymerization of polypropylene is usually
done by means of free radical chain reaction
mechanism shown in Figure 1 consisting of
the steps of: (i) initiation (Eq. (1)), (ii)
propagation (Eq. (2)) and (iii) termination
(Eq. (3)). Throughout the steps, heat-
facilitated hydrogen abstraction from the
tertiary carbons may lead to the formation of
tertiary alkyl radicals shown in Figure 1. In
PP, tertiary radicals are formed mostly, due to
the poorer dissociation energy of a tertiary C-
H bond (ca. 373 kJ.mol-1 at 25ºC), compared
to that of a secondary C-H bond (ca.
394 kJ.mol-1 at 25ºC) [19]. During the later
propagation step, formed tertiary radicals will
follow separate reaction paths. The reaction
paths followed by tertiary radicals are depicted
in schemes.
Termination of polymer reactions involving
radicals or macroradicals may take place [20].
Recombination is influenced by cage effects,
steric control, mutual diffusion and the
molecular dynamics of the polymer matrix
[21–23].
Oxidative Depolymerization
Free radical depolymerization of
polypropylene cannot be controlled, to avoid
such problems in depolymerization; oxidation
on neat polymer is to work with solutions and
initiator begun by Dulog, Radlman and Kerns
[24]. In oxidative depolymerization, the
formation of free radical sites on the chain
backbone by attack of molecular oxygen,
ozone etc. is done [9]. The free-radical chain
reactions depicted in Figure 2 describe
oxidative depolymerization of polypropylene
[20, 25], consisting of individual steps of
initiation, propagation and termination.
During the initiation stage of oxidative
depolymerization, initiating agent stimuli such
as heat or UV radiation are significant for the
growth of radicals. Cleavage of a polymer
chain is done, due to the abstraction of a
hydrogen atom from the polymer backbone
(reaction 2.1).
Fig. 1: Depolymerization Mechanism of Polypropylene.
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 4
Polymer chain scission is frequently caused by
severe deformation of a polymer. In the
propagation steps, the radicals formed during
initiation, react with molecular oxygen and are
changed into macro-alkyl peroxyl radicals
(reaction 2.2). Then abstraction of a hydrogen
atom from another polymer molecule gives
macro-hydroperoxides and more alkyl radicals
(reaction 2.3). Hydroperoxides formed through
this step of reaction might be cleaved
homolytically to give up radical. These kinds
of radicals are capable to abstract hydrogen
form adjacent polymer chains ensuing to the
growth of more radicals accountable for
initiating depolymerization.
Free radicals produced through the first two
steps may also go through further reactions.
Such reactions are identified as β-scission
(reaction 2.5), reduce the molecular weight of
chains and also persuade the crystallinity of
polymers [26]. Considering as amorphous
regions of semi-crystalline polypropylene
depolymerized faster than the crystalline phase
[27–29], the scission of polymer backbone,
dependable for relating two adjoining
crystalline areas, will cause a rapid devastation
of the physical and mechanical properties of a
polypropylene. Elevated mobility of oxygen in
non-crystalline material leads to propagation
within amorphous regions. The termination
reactions are reliant on the molecular structure
of the polymer as well as the existing
depolymerization conditions and termination
(reactions 2.7, 2.8) prevails [20]. Though, in
the case of oxygen starvation alkyl radical
dominate and bimolecular termination
reactions are of superior importance. This
leads to cross-linking which is evidenced by a
rise in molecular weight. Oxidative cycle of
polypropylene is shown in Figure 3 [4].
Fig. 2: Oxidative Free Radical Depolymerization of Polypropylene.
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 5
Sunggyu et al. have patented a process for
selectively depolymerizing isotactic PP in the
oxidative depolymerization pilot plant system,
at 383°C and 232 atm pressure. The batch
charge varied from 0.5 to 20 gm of isotactic
PP, with oxygen flow rates varying from 50 to
2000 scc/min [30]. Upadhyaya et al. patented
a method for depolymerizing polyolefins like
amorphous polypropylene, Fischer-Tropsch
waxes, petroleum waxes, and mixtures.
The process comprises of reacting the molten
polymers, at atmospheric pressure, in the
absence of catalysts at a temperature between
about 130 and about 215o
C. The molten
material was constantly stirred as it was
sparged with air [31]. Jansson et al.
hypothesized that hydroperoxides formed
during the ageing step, decompose in the
subsequent processing step, thereby causing
faster degradation of the materials.
Hydroperoxide formation and decomposition
into radicals are known to play a key role in
polymer auto-oxidation [32].
Thermal Depolymerization
Thermal depolymerization of polypropylene is
molecular reduction as a result of overheating.
At high temperatures the components of the
long chain backbone of the polypropylene can
begin to separate the molecular scission it
involves change in molecular weight,
molecular weight distribution. The
conventional model for thermal
depolymerization involves the major steps of
depolymerization such as initiation (4.1, 4.2),
propagation (4.3), branching (4.4) and
termination (4.5) shown in Figure 4 [33].
Fig. 3: Oxidative Cycle of Polypropylene [4].
Fig. 4: Mechanism of Thermal Depolymerization.
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 6
Polypropylene is very susceptible to thermal
depolymerization even at normal temperature
which causes chain scission; the reduced chain
length reduces molecular weight. This can
considerably change the mechanical
properties, thermal properties and crystallinity.
Ying et al. explained the mechanism of
thermal oxidative and thermal mechanical
degradation of isotactic polypropylene (PP) in
the presence of incorporated organic peroxide.
The change in MWD was explained by a
random chain scission mechanism and
mechanical chain scission [34]. Polypropylene
subjected to multiple extrusion shows that the
chain scission processes during thermo-
mechanical depolymerization causes the
reduction in molecular weight. This indicates
that the probability of chain breaking is
dependent on the depolymerization and the
molecular weight of the chain [35].
Free Radical Induced Depolymerization
Free-radical initiators are chemical substances
that, under certain conditions, initiate chemical
reactions by producing free radicals as cause
of homolytically cleavage [36]. Radicals are
reactive chemical species possessing a free
electron and ions. Initiator-derived radicals are
very reactive chemical intermediates and
generally have short lifetimes [37], i.e., half-
life times less than 10-9. The decomposition
rate of these peroxides depends on the class of
peroxide as well as on the type of alkyl group.
The eight classes of organic peroxides that are
produced commercially for use as initiators are
listed as follows Diacyl peroxides, Dialkyl
peroxydicarbonates, tert-Alkyl peroxyesters,
OO-tert-Alkyl O-alkyl monoperoxy
carbonates, Di (tert-alkylperoxy) ketals, Di-
tert-alkyl peroxides, tert-Alkyl
hydroperoxides, Ketone peroxides.
Fig. 5: Depolymerization of PP in Presence of Free Radical Initiators.
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 7
The Saule et al. studied of the modification of
polypropylene based on the decomposition of
unsaturated peroxides which helped to
decrease the molecular weight of the polymer.
Mechanisms for the branching and the
depolymerization of the polypropylene were
also proposed [38]. To optimize desired
processing capabilities and the end use
properties the controlled degradation of
polypropylene (PP) using peroxides is
proposed by Iedema et al.
The author performed a series of controlled
degradation experiments on a twin-screw
extruder with polypropylene of varying MW
under addition of various amounts of initiator
and developed a model which was adapted to
the geometry of the extruder entrance and the
peroxide feed practice. Effect of thermal
heterogeneities, residence time distribution,
micromixing on molecular weight distribution
was studied [39]. Balke et al. presents a new
kinetic model for the free radical initiated
degradation of polypropylene in the reactive
extrusion. The peroxide used was 2,5-
Dimethyl-2,5-di(t-butylperoxy) hexane. The
resulting product is known to provide superior
processing properties; this is an example of the
production of a specialty polymer by chemical
modification of a commodity polymer [40].
Possible reactions mechanism of
polypropylene in presence of free radical
initiators or organic peroxide is shown in
Figure 5 [41].
There are several common processes for
supplying the energy required to variety
radicals from initiators thermal and radiation
[36]. Formerly produced, radicals experience
two basic types of reactions, propagation
reactions and termination reactions. The
initiator molecule is denoted by I and the free
radical produced is shown by I*. In a
propagation reaction, a radical reacts to form a
covalent bond and to create a new radical. The
three most ordinary propagating reactions are
atom abstraction, ß-scission, and addition to
carbon–carbon double bonds or aromatic
rings. The polypropylene radicals produced
will further react with other polymer
molecules. The main importance of shift to
polymer is chain branching which frequently
occurs in two ways such as:
i. Hydrogen might be abstracted from sites
on the backbone thus transferring radical
movement and initiating a long chain
branch.
ii. Short chain branches possibly will be
fashioned by intra-molecular atom transfer
(backbiting). In this case the propagating
radical abstracts backbone hydrogen as of
a close by carbon.
In a termination reaction, two radicals
interrelate in a mutually vicious reaction in
which both radicals fashioned covalent bonds
and reaction ceases. Termination of the
polymer radicals can also be done by
recombination, where two polymer radicals
formed would merge.
i. In which combination of two tertiary
radicals would effect in cross-linking,
which is slightest liable in case of
polypropylene.
ii. Amalgamation of two secondary, or a
secondary and tertiary radical would give
a polymer molecule or by
disproportionate.
In the provisions of the free radical
depolymerization theory, intermolecular
transfer of the radical was found to be the
leading process. An arch linking the degree of
depolymerization to the intrinsic viscosity was
obtained from calculations based on a random
depolymerization. With this arch, intrinsic
viscosities measured as a function of
depolymerization time could be used to gain
the time dependence of the degree of
depolymerization [13].
Depolymerization by High Energy
Radiation
Depolymerization of polypropylene can be
done by means of irradiation methods, such as
electron beams, ultrasonic, ultraviolet (UV)
and γ-rays, which have been favored to modify
the structure and properties of polypropylene
besides using chemical initiators and free
radicals.
Electron beams, ultrasonic, ultraviolet
radiation and γ-rays energy which possibly
will cause bond scission by free radical
formation for higher doses than 60 kGy for
long period depolymerization occurs. The
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 8
depolymerization of polypropylene resulting
from experience to high energy radiation is an
important subject for many reasons. The
mechanism of high energy induced
depolymerization is a free radical, one similar
to the general mechanism of depolymerization
as shown in Figure 6.
Fig. 6: Mechanism of High Energy Induced Depolymerization PP.
In case of PP irradiation it has been
established that the PP undergoes essentially
chain scission. Polymers are increasingly more
exposed to radiation in their service where
depolymerization may cause [42]. Radiation-
induced depolymerization has been well
reviewed [43–46]. High energy radiation is
defined as all forms of radiation with energies
much higher than those of chemical bonds. It
includes both electromagnetic radiation (x-
rays, γ-rays, etc.) and particulate beams (α-
and β-particles, electrons, neutrons, etc.). The
series of energies is exceptionally broad and
superior to chemical bond energies. Radiation
is absorbed by interaction with atomic nuclei
and electron clouds. He et al. developed UV
initiation based reactive procedure for
depolymerization of polypropylene for the
production of controlled rheology
polypropylene, benzophonene has been used
as the photo initiator. At low-irradiation time
and low-levels of photo initiator molecular
weight distribution was noticeably tailored.
After depolymerization, MFR improved with
drop off in viscosity and elasticity with
benzophonene concentration as predictable.
Crystallinity levels and crystallization
temperatures of the tailored polypropylene
were lower than those of virgin polypropylene
because of the reduced molecular weight and
narrower MWDs [47]. Radiation induced
depolymerization caused by chain scission was
successfully used. Controlling the degree of
degradation, uniform molecular weight is
friendly process are the beneficial effects of
using radiation technology. When polymeric
materials are irradiated by ionizing radiation,
they are divided into two types, degradation
(chain scission) and chain link (crosslinking).
Studying the effect of hydrogen peroxide
and/or γ irradiation on the degradation process
of Na-alginate was investigated. It was found
that the molecular weight of the polymer
decreases by using gamma radiation or H2O2.
However, combining both γ radiation and
H2O2 accelerates the degradation rate [48]. The
influence of three hindered amine light
stabilizers (HALS) and two EVA copolymers
on the radiation degradation processes in
isotactic polypropylene (iPP) was investigated
by Zimek et al. The process consequences
degradation, oxidation, release of harmful
gaseous products, etc. as the polymer is known
as one of the most degradable upon experience
to ionizing radiation. Polypropylene properties
are changed even before exposure to electron
beam due to structural modifications caused
due to chain scission initiated by additional
components [49]. The range of energies is
extremely broad and higher than chemical
bond energies. Radiation is absorbed by
interaction with atomic nuclei and electron
clouds. The radiation interaction with the
electron clouds of molecules consequences in
the transfer of energy to the molecules to form
ions, with the elimination of a secondary
electron [50, 51]. Further ionization and
excitation of nearby molecules is caused by
these secondary electrons having enough
kinetic energy. The immediate effect of
absorption of high energy radiation is
construction of energetic or excited species,
including trapped electrons, ions, and radical
ions, results in fragmentation to provide free
radicals. In polypropylene, scission of main-
chain C-C bonds gives radical pairs. The
scission of C-H bonds leads to comparatively
stable radicals and formation of molecular
hydrogen. The primary radical formed by
irradiation might not be stable. For example,
irradiation of polypropylene with γ-radiation at
doses of 25, 50, 100 and 150 kGy produces the
secondary radical [52]. In the absence of
oxygen, carbon-centered radicals might
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 9
undergo cross-linking, by recombination,
chain scission, β-elimination depending on the
radical structure. Polymers can consequently
be divided into two main groups, those which
react to radiation by cross-linking and those
which drop molecular weight by chain scission
[53]. Ultrasonic irradiation has been recently
looked upon as a new technique for
degradation of polymer compounds, mainly
due to the fact that the reduction in the
molecular weight is simply by splitting the
most susceptible chemical bond without
causing any changes in the chemical nature of
the polymer. Detailed analysis of cavitations
generated ultrasound radiation used for
molecular weight reduction of polypropylene
has been done by Desai et al. The effect of
various operating parameters including initial
concentration of the polymer, power density
into the system and type of solvent, on the
extent of depolymerization of polypropylene
are investigated. Cavitations results in liquid
turbulence associated with liquid shear and
generation of highly reactive free radicals. The
shear force leads to the rupture of chemical
bonds of the polypropylene, ultrasonic sound
waves produce a permanent reduction in
viscosity [54].
CHARACTERIZATIONS TO STUDY
DEPOLYMERIZATION
As it has been mentioned before,
polypropylene has extreme importance
commercially, but regrettably it is also less
susceptible to attack by heat, oxygen,
radiation, and free radical initiator. It is very
essential to study the depolymerization process
i.e., to set up the mechanisms throughout
oxidative depolymerization, the products
formed as well as the influence of molecular
weight and structure on the properties of
material. A full understanding of
depolymerization is of great significance,
since without this, successful stabilization
approaches and superior methods of lifetime
prediction would not have been possible. It is,
consequently, essential to consider the
characterization techniques employed to study
the depolymerization process. Analytical
chemistry field possesses a conventional
techniques that have been used for the purpose
of studying polypropylene depolymerization,
some of which include high temperature gel
permeation chromatography (GPC), Fourier
transform infrared spectroscopy (FTIR),
differential scanning calorimetry (DSC) and
viscometry. A brief overview on the use of
some of these techniques for the purpose of
studying polyolefin depolymerization will now
be presented.
High Temperature Gel Permeation
Chromatography
According to hydrodynamic volume high
temperature gel permeation chromatography
(HTGPC) separates molecules and is the
preferred method for the determination of
molecular weight and molecular weight
distribution. During the depolymerization of
polypropylene, changes in average molecular
weight and molecular weight distribution
(MWD) are often observed [55–57].
Depending on the polymer and the
depolymerization methods, many radical
reactions can take place. In the case of
polyolefin’s, branching and recombination
reactions predominate at lower temperatures,
yielding a long chain branching, which leads
to an increase in the average molecular weight
of the material [58]. In case of polypropylene,
depolymerization almost exclusively by means
of chain scission, leads to a reduction of its
average molecular weight [59]. Chain scission
during depolymerization consequences in
continuous breaking of polymer chains
yielding chains of shorter lengths. The average
result is that the number of short chains
increases with depolymerization time and is
accompanied by a broadening of the molecular
weight distribution.
The molecular weight curve obtained from
GPC measurements, as a result, shifts towards
the region of lower molar mass when chain
scission is the dominant depolymerization
mechanism [60–62]. As depolymerization
proceeds and chain scissions continue, the
molecular weight curve will exhibit bimodality
with the portion of highly depolymerized short
molecules appearing as a narrow distribution
on the lower molecular weight side of the
original material. David et al. proposed an
Eq. (A) for determining the number of chain
scissions in degraded samples [63], by shaping
the numerical average of the initial and final
number-average molecular weights.
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 10
𝑛 𝑅 =
𝑀 𝑛𝑜
𝑀 𝑛𝑓
− 1 (A)
Where,
𝑛 𝑅is the number of chain scission,
𝑀 𝑛𝑜 is initial number average molecular
weight,
𝑀 𝑛𝑓is final number average molecular weight.
High temperature gel permeation
chromatography can be used to record
molecular weight changes over a cross-section
of a thick polymer sample by means of
microtoming [64] or layer-by-layer milling
[65]. Computer-aided study method called
molecular weight distribution computer
analysis (MWDCA) has also proven to be
useful for the comparison of depolymerization
rates in polypropylene [66–68]. This technique
uses molecular weight data, obtained by
evaluation of experimental GPC molecular
weight distribution profiles with computer
simulations for determining the scission or
crosslink. Shyichuk investigated chain scission
and crosslink concentrations in photo-
degraded LDPE, polypropylene and ethylene-
propylene copolymer.
Differential Scanning Calorimetry
Differential scanning calorimetry is an
additional technique frequently used to obtain
information on the depolymerization behavior
of polymers [69, 70]. Profile of melt
endotherms and changes in the peak
temperatures and as well as the glass transition
temperature, can provide information on the
susceptibility of different crystalline phases or
arrangements to depolymerization [71]. DSC
is very useful in determining the oxidative
induction time (OIT) [72–74], oxidative
temperature (Tox) [75] and thermal stability
degree of depolymerization of polymers under
high temperature conditions of polymers. Rosa
et al. investigated the influence of several
parameters on DSC statistics and found that
sample preparation (shape and size), oxygen
flow and heating rate of the experiment had a
considerable influence on the data obtained
[70].
Fourier Transform Infrared Spectroscopy
For studying out the chemical changes brought
about by polymer depolymerization IR
spectroscopy is one of the most accepted
techniques. It has been used for qualitative as
well as quantitative characterization of
depolymerization products [76–80]. The
depolymerization route creates the formation
of different functional groups which are
strongly reliant on the chemical structure of
the polymer. The main chemical species
detectable by infrared spectroscopy are
hydroxyl and carbonyl groups [81]. The
configuration of these groups usually leads to
visible changes in infrared spectrum,
appearing in the regions of 1850–1550 and
3700–3200 cm-1
, respectively [82–84].
Moreover, change in the FTIR absorption of
depolymerization products in PE and PP can
also be experimental and used for studying the
difference in their oxidative depolymerization
mechanisms [85]. From the hydroxyl and
carbonyl groups, depolymerization products
such as peroxides, alcohols, can be identified
by means of derivatisation reactions [86].
Derivatisation methods for this point were first
applied to polyolefins by Carlsson et al. [87].
FTIR has also been sensible in determining
changes in unsaturation of PE during
depolymerization [88]. The degree of
depolymerization of polypropylene is
determined on the basis of their carbonyl index
[89–91]. The bands at 1892, 974, 2720 and
840 cm-1
have all been used as reference bands
for determining the carbonyl index in
polypropylene [89–92].
Viscometry
Viscometry is a useful technique for
determining the polymer molecular weight.
The molecular weight obtained by this
technique is the viscosity averaged molecular
weight. The increase in the viscosity imparted
by the macromolecules in solution is a direct
function of the hydrodynamic volume and,
hence, the molecular weight of the
macromolecule. The relationship between the
intrinsic viscosity and the viscosity average
molecular weight is given by the semi-
empirical Mark-Houwink Eq. (B) [9].
[𝜂] = 𝐾(𝑀 𝑉)⍺
(B)
Where η is intrinsic viscosity, K and ⍺ are
constants for a polymer-solvent system at a
given temperature and Mv is viscosity average
molecular weight. The η, values for the
polypropylene solution were calculated by the
one point intrinsic viscosity Eq. (C).
[𝜂] =
(2(ηsp−lnηr))
0.5
𝐶
(C)
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 11
Where ηr and ηsp are relative and specific
viscosities respectively, C is concentration of
polymer solution [93]. The viscometric studies
of polymer solutions as a means of molecular
characterization of polymers are well
recognized and widely practiced because of
simplicity in terms of experimental approach
and the apparatus needed. Dilute solution
viscosity can be conveniently measured in
capillary viscometers such as Ostwald type or
the Ubbelohde type [9]. Shenoy et al.
dissolved the depolymerized polypropylene
samples in xylene at 100°C and the viscosity
measurement at 85°C was carried out by using
the Ubbelohde viscometer. Relative (ηr) and
specific viscosities (ηsp) respectively were
calculated using standard formulae to
determine the viscosity average molecular
weight [94]. For PP-xylene system at 85°C the
Mark Houwink constants are K=9.6×10–3
and
α=0.63 [100].
EFFECTOF MOLECULAR WEIGHT
REDUCTION
Melt Flow Index
The melt flow index (MFI) of the samples is
considered as a critical parameter in polymer
processing and industrial designs. MFI of a
polyolefin resin refers to the rate at which it
extrudes from a capillary die under a standard
set of conditions. The MFI is reflected by the
average dimensions of the molecules in a resin
and their entanglements with one another so it
depends on molecular characteristics (Mw and
MWD) and branching characteristics, short
chain branching (SCB), of the sample. Fazeli
et al. have showed that the MFI of the samples
is related to their branching characteristics.
With increasing the degree of branching
(SCB/100°C), the MFI of the sample goes
through a minimum. The increase in SCB will
cause more entanglements between the chains
(inter-molecular entanglement) and therefore
will impede the flow [101]. However, after the
minimum point, increase in the SCB/100°C
will cause the chains to have a more compact
molecular profile (more intra-molecular
entanglements instead of inter-molecular
ones), so the chains will cause less hindrance
to the flow of other chains. In other words, the
degree of branching (number of branching per
100°C atoms of main chain) and the amount of
branching (the number of chains which have
the same degree of branching), have the same
effect on MFI. The melting point of a-form
iPP is strongly influenced by the
stereoregularity [102, 103]. Melting points in
the 160–168°C range are typical for
commercial homopolymer samples under
normal analysis conditions.
Melt Strength
The melt strength of a polymer is defined as
the maximum force at which a molten thread
can be drawn under standard conditions before
it breaks. High values of MS are desired in
processes where the material is stretched in its
molten state, such as in film blowing,
thermoforming, or foaming. Lagendijk et al.
have shown how the melt strength is enhanced
by the presence of strain hardening in
elongational viscosity [104]. Increasing
average molecular mass of a polymer results in
higher shear viscosity, as well as higher melt
strength. The melt strength also increases
when the molecular mass distribution (MWD)
becomes broader. Ghijsels et al. have
demonstrated how the melt strength increases
much more dramatically than shear viscosity
upon the addition of long chain branches on
the polymer backbone [105–107]. De Maio
and Dong have studied the effect of chain
structure on melt strength of polypropylene.
Comparison among several linear and
branched polypropylenes obtained by electron
beam irradiation has shown that the melt
strength of branched can be up to 10 times
higher than that of linear PP with the same
MFI [108]. Gotsis et al. found that the
enhancement of the melt strength is related to
the increase of the weight average number of
long chain branches per molecule [109].
Rheology
The molecular architecture of the polymer i.e.
short and long chain branching and MWD
affects the rheology of the melt. In a thorough
review on the effect of long chain branching
(LCB) on the linear viscoelasticity of
polyolefins, Vega et al. showed that the
introduction of LCB induces higher levels of
elasticity than broadening of the MWD of a
melt of linear chains of similar molecular
weight [110]. On the other hand, it is current
understanding that short chain branching
(SCB) cannot cause large increases in
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 12
elasticity. Yet, Vega et al. reported that SCB
resulted in higher zero shear rate viscosities,
higher relaxation times, higher values of the
elastic storage modulus and higher activation
energies of flow compared to linear polymers
[111]. Those authors surmised that the number
of entanglements per branch is a decisive
parameter that influences shear and
elongational behavior. Much less research has
been done on the extensional properties of PP
melts because appropriate samples have been
difficult to obtain. Hingmann and Marczinke
found that branched PP samples showed
distinct strain hardening and concluded that a
few branches on the chain had an enormous
effect on the extensional behavior of the melt
[112].
Crystallinity
Wide angle X-ray scattering (WAXS) patterns
of iPP, sPP, and aPP are shown in Figure 7
[113]. The regular molecular structure of iPP
and sPP readily enables crystallization of the
chains, leading to well defined crystalline
reflections differing in unit cell symmetry. aPP
lacks a regular molecular structure, and does
not crystallize [114]. Brucker et al. have
shown that the isotactic polypropylene can
crystallize into three different crystal forms
depending on the temperature, pressure and
mechanical stress state: monoclinic a-,
orthorhombic γ- and hexagonal ß-forms [115,
116]. The γ-crystals are formed only under
high pressure in high molar mass
homopolymer polypropylene. Polypropylene
samples with low molar mass or low tacticity
and polypropylene copolymers crystallize
partially in γ-form. The degree of crystallinity
varies between 0 for a completely amorphous
material (such as aPP) and 1 for a completely
crystalline material. The degree of crystallinity
plays a critical role in determining properties
of polypropylene. Commonly measured
properties such as; modulus, yield stress,
oxygen-moisture barrier resistance, and
hardness increase with increasing crystallinity.
In addition to tacticity, crystallinity generally
increases with decreasing molecular weight
(increased chain mobility), and is promoted by
slower cooling rates from the melt. Busico et
al. have explained how the density of iPP in
the a-form varies between the limit of 100%
amorphous (ϱa=0.850 to 0.855 g/cm3) and
100% crystalline (ϱc=0.936 to 0.946 g/cm3)
[117].
Fig. 7: Wide-Angle X-Ray Scattering (WAXS) Patterns of iPP, sPP, and aPP [113].
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 13
Balta-Calleja and Vonk indicated a method for
evaluating crystallinity as a number by
numerical means from the relationship of the
peak area to the total area [118]. Jinghua et al.
investigated the nonisothermal crystallization
kinetics of linear and long chain branched
polypropylene (LCB PP) by differential
scanning calorimetry (DSC) at various cooling
rates. It was shown that LCB has the role of
heterogeneous nucleating agent and
accelerates the crystallization process of PP.
Also the activation energy of LCB PPs are
higher than that of linear PP, indicating that
the presence of LCB baffles the transfer of
macromolecular segments from PP melt to the
crystal growth surface [119]. Furthermore, the
crystal morphology of linear PP and LCB PPs
was observed through polarized optical
microscopy (POM), and fine spherulites were
observed for LCB PPs. De Nicola et al. and
Wang et al. have independently investigated
the effect of branching on the crystallization
behaviors of PP resins [120, 121]. It was
observed that the introduction of branching in
the PP resins increases crystal nuclei density.
This promotes faster crystallization, and,
hence, higher crystallization temperatures
were observed for branched materials as
compared to linear materials. The cooling rate
was fixed to be 10°C/min. It was observed that
the branching of PP chains significantly
promoted the crystallization kinetics of the PP
resins by increasing the crystallization
temperature about 20°C.
Chemical Properties
iPP is soluble in high boiling aliphatic and
aromatic hydrocarbons at high temperature.
The high chemical resistance of iPP results in
exceptional stain resistance, and has led to the
use of iPP in automobile batteries. iPP has
outstanding resistance to water and other
inorganic environments. iPP resists most
strong mineral acids and bases, but like other
polyolefins is subject to attack by oxidizing
agents including 98% sulfuric acid and 30%
hydrochloric acid at high temperature
(13<100°C), and fuming nitric acid (ambient
temperature) [122]. PP reacts with oxygen in
several ways, causing chain scission and
brittleness that is associated with the loss in
molecular weight. This action is promoted by
high temperatures, light, or mechanical stress.
Several scientists have independently
demonstrated treatment of polypropylene with
peroxides which has led to controlled rheology
resins with reduced molecular weight and
narrow polydispersity relative to polymerized
product from Ziegler-Natta catalysts [123–
125]. These resins are used in some fiber
spinning and injection molding applications.
The creation of radical sites along the polymer
backbone, most often through peroxide-based
initiation, is also an essential condition for
many fictionalization/grafting chemistries.
CONCLUSION
It is concluded from the foregone discussion
that understanding about the mechanism can
go a long way in helping the researchers and
the technologists to induce the different types
of depolymerization methods in the
polypropylene. This depolymerization can
further be enhanced by the addition of the
initiator, additives in the polypropylene and by
understanding the various factors such as
mechanical, physical, and chemical which are
responsible for this depolymerization. It is also
concluded from this discussion that
polypropylene depolymerization could be
enhanced its properties with modification in
structure.
REFERENCES
1. Natta G, Pino P, Corradini P. J Am
Chem Soc. 1955; 77.
2. Seymour RB. History of Polyolefins. In:
Vasile C, Seymour RB, editors.
Handbook of Polyolefins. New York:
Marcel Dekker; 1993; 1p.
3. Qicong Ying, Yong Zhao, Yong Liu.
Makromol Chem. 1991; 192.
4. Denis Bertin, Marie Leblanc, Marque
Sylvain RA, et al. Polym Degrad Stabil.
2010; 95.
5. Imperial Chemical Industries Ltd. Germ
Pat. 1210562. Mar 17, 1961.
6. Chemische Werke Hls AG. French Pat.
1377951. Nov 6, 1964.
7. Caver Hill R, Taylor GW. Polymer.
1965; 6.
8. Gooch Jan W. Encyclopedic Dictionary
of Polymers. 2nd Edn. Springer; 2011.
9. Gowarikar VR, Vishwanathan NV,
Sreedhar Jayadev. Polymer Science.
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 14
Reprint Edn. India: New Age
Publishers; 1994.
10. Laurence Stephen Rayner. Imperial
Chemical Industries Ltd. US Pat.
3232918. Feb 1, 1966.
11. Boehm Hans, Henglein Friedrich,
Schmidt-Thomee, et al. BASF AG, DE
1201552. Sep 1965.
12. Saunders Frank Linwood. Dow
Chemical Company, DE 1221798. Jul
1966.
13. Thomas Davis E, Robert Tobias L,
Peterli Elizabeth B. J Poly Sci. 1962; 56.
14. Wood DL, Luongo JP. Mod Plast. 1961;
38.
15. Williams SD, Yoo HJ, Drickman MR.
US Patent 5820981. 1998.
16. Lagendijk RP, Hogt AH,
Buijtenhuijsand A, et al. Polymer. 2001;
42.
17. Sugimoto M, Tanaka T, Masubuchi Y,
et al. J Appl Polym Sci. 1999; 73.
18. De Nicola Jr., Anthony J. US Patent
5047446. Sep 01, 1991.
19. Gonzalez-Gonzalez VA, Neira-
Velazquez G, Angulo-Sanchez JL.
Polym Degrad Stabil. 1998; 60.
20. Al-Malaika S. Adv Polym Sci. 2004;
169.
21. Garton A, Carlsson DJ, Wiles DM. J
Polym Sci Polym Chem Ed. 1978; 16.
22. Garton A, Carlsson DJ, Wiles DM.
Macro. 1979; 12.
23. Al-Malaika S. Polym Plast Technol Eng.
1988; 27.
24. Dale Van Sickle E. J Polym Sci. 1972;
10(a-1).
25. George GA, Celina M, Hamid SH.
Handbook of Polymer Degradation. 2nd
Edn. New York: Marcel Dekker Inc.;
2000.
26. Rabello MS, White JR. Polymer. 1997;
38.
27. Blais P, Carlsson DJ, Wiles DM. J
Polym Sci. 1972; 10(A1).
28. Schlick S, Kruczala K. JCT Res. 2005;
2.
29. Knight JB, Calvert PD, Billingham NC.
Polymer. 1985; 26.
30. Lee Sunggyu, Gencer Mehmet A,
Fullerton Kathy L, et al. US Pat.
5386055. Jan 31, 1995.
31. Upadhyaya Janardan D. US Pat.
4624993. Nov 25, 1986.
32. Anna Jansson, Kenneth Moller, Thomas
Hjertberg. Polym Degrad Stabil. 2004;
84.
33. Thermal Degradation of Polymers a
Primer. Pop Plast Packag. 2013; 12.
34. Qicong Ying, Yong Zhao, Yong Liu.
Makromol Chem. 1991; 192.
35. Sebastiaao V, Canevarolo. Polym
Degrad Stabil. 2000; 709.
36. Myers Terry N, Mark Herman F.
Encyclopedia of Polymer Science and
Technology. 3rd Edn. John Wiley and
Sons; 2004.
37. Griller D, Ingold KU. Acc Chem Res.
1976; 9.
38. Myriam Saule, Laurence Moine, Marie
Degueil-Castaing, et al.
Macromolecules. 2005; 38.
39. Iedema PD, Remerie NK, Vanderhamb
BM, et al. Chem Eng Sci. 2011; 66.
40. Balke ST, Suwanda D, Lew R. J Polym
Sci: Part C, Polym Lett. 1987; 25.
41. Imoto Minoro, Takemoto Kiichi, Kono
Masatsugu. Angew Makromol Chem.
1967; 1.
42. Clough RL, Gillen KT. Nuc Techno.
1982; 59.
43. David C. Degrad Polym. 1975; 14.
44. Schnabel WS, Jellinek HHG.
Degradation and Stabilization of
Polymers. Amsterdam: Elsevier; 1978.
45. Schnabel WS. Polymer Degradation:
Principles and Practical Applications.
Munich: Hanser; 1981.
46. Ivanov VS. Radiation Chemistry of
Polymers. Amsterdam: VSP; 1992.
47. Guangjian He, Costas Tzoganakis.
Polym Eng Sci. 2011.
48. Hegazy EA, Abdel-Rehim H, Diaa DA,
et al. Controlling of Degradation Effects
in Radiation Processing of Polymers.
International Atomic Energy Agency;
2009.
49. Zimek Z, Przybytniak G, Rafalski A, et
al. International Atomic Energy Agency;
2009.
50. Swallow A. Radiation Chemistry.
London: Longman; 1973.
51. Spinks JWT, Woods RJ. An Introduction
to Radiation Chemistry. New York:
John Wiley and Sons, Inc.; 1976.
52. Krupa I, Luyt AS. Polym Degrad Stabil.
2001; 72.
Journal of Polymer & Composites
Volume 4, Issue 2
ISSN: 2321-2810(online), ISSN: 2321-8525(print)
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 15
53. Bilingham NC. Degradation. In: Mark
HF, editor. Encyclopedia of Polymer
Science and Technology. Vol. 6. John
Wiley and Sons; 2005.
54. Vaibhav Desai, Shenoy Mohan A,
Gogate Parag R. Chem Eng J. 2008;
140.
55. Fayolle B, Audouin L, Verdu J. Polym
Degrad Stab. 2002; 75.
56. Gonzalez-Gonzalez VA, Neira-
Velazquez G, Angulo-Sanchez JL.
Polym Degrad Stab. 1998; 60.
57. Lew R, Suwanda D, Balke ST. J Appl
Polym Sci. 1988; 35.
58. Yadav A, Vimal KK, Singh R. J Polym
Composites. 2016; 4(1).
59. Ying Q, Zhao Y, Liu Y. Macromol
Chem. 1991; 192.
60. Alam MS, Nakatani H, Goss BGS, et al.
J Appl Polym Sci. 2002; 86.
61. Manabe N, Yokota H, Suzuki S, et al. J
Appl Polym Sci. 2006; 100.
62. Nakatani H, Manabe N, Yokota Y, et al.
Polym Int. 2007; 56.
63. David C, Trojan M, Daro A, et al.
Polym Degrad Stab. 1992; 37.
64. Girois S, Audouin L, Verdu J, et al.
Polym Degrad Stab. 1996; 51.
65. Turton TJ, White JR. Polym Degrad
Stab. 2001; 74.
66. Shyichuk AV, White JR, Craig IH, et al.
Polym Degrad Stab. 2005; 88.
67. Shyichuk AV, White JR. J Appl Polym
Sci. 2000; 77.
68. Shyichuk AV, Stavychna DY, White JR.
Polym Degrad Stab. 2001; 72.
69. Olivares N, Tiemblo P, Gomez-Elvira
JM. Polym Degrad Stab. 1999; 65.
70. Rosa DS, Sarti J, Mei LHI, et al. Polym
Test. 2000; 19.
71. Elvira M, Tiemblo P, Gomez-Elvira JM.
Polym Degrad Stab. 2004; 83.
72. Camacho W, Karlsson S. Polym Degrad
Stab. 2002; 78.
73. Chang TC, Yu PY, Hong YS, et al.
Polym Degrad Stab. 2002; 77.
74. Groning M, Eriksson H, Hakkarainen
M, et al. Polym Degrad Stab. 2006; 9.
75. Ahlblad G, Gijsman P, Terselius B, et
al. Polym Degrad Stab. 2001; 73.
76. Gulmine JV, Janissek PR, Heise HM, et
al. Polym Degrad Stab. 2003; 79.
77. Edge M. Infrared Spectroscopy in
Analysis of Polymer Degradation. In:
Meyers RA, editor. Encycl Anal Chem.
Chichester: John Wiley and Sons Ltd.;
2000.
78. Gugumus F. Polym Degrad Stab. 1997;
55.
79. Nekhoroshev VP, Turov YP,
Nekhorosheva AV, et al. J Appl Chem.
2006; 79.
80. Costa L, Luda MP, Trossarelli L. Polym
Degrad Stab. 1997; 58.
81. Andreassen E, Karger-Kocsis J. Infrared
and Raman Spectroscopy of
Polypropylene. Dordrecht: Kluwer
Publishers; 1999.
82. Mani R, Singh RP, Sivaram S. Polym
Int. 1997; 44.
83. Rivaton A, Gardette JL, Mailhot B, et al.
Macromol Symp. 2005; 225.
84. Gugumus F. Polym Degrad Stab. 1999;
6.
85. Adams JH, Goodrich JE. J Polym Sci.
1970; Part A1(8).
86. Piton M, Rivaton A. Polym Degrd Stab.
1996; 53.
87. Carlsson DJ, Brousseau R, Zhang C, et
al. ACS Symp Ser. 1988; 364.
88. Kolbert AC, Didier JG, Xu L.
Macromolecules. 1996; 29.
89. Jansson A, Moller K, Gevert T. Polym
Degrad Stab. 2003; 82.
90. Miraftab M, Horrocks AR, Mwila J.
Polym Degrad Stab. 2002; 78.
91. Santos ASF, Agnelli JAM, Trevisan
DW, et al. Polym Degrad Stab. 2002;
77.
92. Rabello MS, White JR. Polym Degrad
Stab. 1997; 56.
93. Solomo O, Ciuta ZZ. J Appl Polym Sci.
1962; 6.
94. Shenoy MA, Mrinalini Patil. Polym Sci.
2010; 52.
95. Feng-Hua Su, Han-Xiong Huang. Polym
Eng Sci. 2009; 50(2).
96. Hamed Azizi, Ismaiel Ghasemi. Iran
Polym J. 2005; 14(5).
97. Manfred Raetlsch, Achim Hesse,
Harmut Bucka, et al. US Pat. 6136926.
Oct 24, 2000.
98. Horst David E, Michael Roth, Peter. US
Pat. 2007/0200272. Aug 30, 2007.
Depolymerization of Polypropylene Kulkarni and Pache
JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 16
99. Gaylord Norman G. US Pat. 4506056.
Mar 19, 1985.
100. Brandrup J, Immergut EH, Grulke EA.
Polymer Handbook. New York: Wiley;
1999.
101. Fazeli N, Arabi H, Bolandi. Polym Test.
2006; 25.
102. Wunderlich B. Macromolecular Physics.
1980; 3.
103. Paukkeri R, Lehtinen A. Polymer. 1993;
34.
104. Lagendijk RP, Hogt AH, Buijtenhuijs A,
et al. Polymer. 2001; 42.
105. Ghijsels A, Ente JSM, Raadsen J. Int
Polym Process. 1990; 5.
106. Ghijsels A, Clippeleir De J. Int Polym
Process. 1994; 9.
107. Ghijsels A, Massardier CHC, Bradley
RM. Int Polym Process. 1997; 12.
108. De Maio VV, Dong D. Proceedings SPE
ANTEC. 1997; 43.
109. Gotsis AD, Zeevenhoven BLF, Hogt
AH. Polym Eng Sci. 2004; 44.
110. Vega J, Aguilar M, Peon J, et al. e-
Polymers. 2002; 46.
111. Vega JF, Santamaria A, Munoz-Escalera
A, et al. Macromolecules. 1998.
112. Hingmann R, Marczinke BL. J Rheol.
1994; 38.
113. Phillips RA, Wolkowicz MD, Moore
EP. Polypropylene Handbook. Munich:
Hanser; 1996.
114. Natta G, Corradini P. Del Nuovo
Cimento. 1960; 15.
115. Takodoro H. Structure of Crystalline
Polymers. USA: John Wiley and Sons;
1979.
116. Brucker S, Meille V, Petraccone, et al.
Prog Polym Sci. 1991; 16.
117. Busico V, Corradini P, Debiasio R, et
al. Macromolecules. 1994; 27.
118. Balta-Calleja FJ, Vonk CG. X-ray
Scattering of Synthetic Polymers.
Amsterdam: Elsevier; 1989.
119. Tian Jinghu, Yu Wei, Zhou Chixing. J
Appl Polym Sci. 2007; 104.
120. De Nicola AJ, Galombos AF,
Wolcowicz MD. Polym Mater Sci Eng.
1992; 67.
121. Wang X, Tzoganakis C, Rempel GL. J
Appl Polym Sci. 1996; 61.
122. Pro-fax Polypropylene Chemical
Resistance. Montell Polyolefins
Technical Bulletin. 1996; TL-101.
123. Zeichner G, Patel P. 2nd World
Congress of Chemical Engineering,
Montreal, P.Q., Canada. 1981.
124. Dziemianowicz TS, Cox WW. SPE
ANTEC. 1985; 85.
125. Gahleitner M, Wolfschwenger J,
Bachner C, et al. J Appl Polym Sci.
1996; 61.
Copyright Notice
Declaration & Copyright Transfer Form
I, the undersigned author(s) of the submitted
manuscript, hereby declare, that the above
manuscript which is submitted for publication
in the STM Journals(s), is not published
already in part or whole (except in the form of
abstract) in any journal or magazine for private
or public circulation, and, is not under
consideration of publication elsewhere.
 I will not withdraw the manuscript after
1 week of submission as I have read the
author guidelines and will adhere to the
guidelines.
 I have neither given nor will give this
manuscript elsewhere for publishing after
submitting in STM Journal(s).
 I have read the original version of the
manuscript and am responsible for the
thought contents embodied in it. The work
dealt in the manuscript is my/our own, and
my/our individual contribution to this
work is significant enough to qualify for
authorship.
 I also agree to the authorship of the article
in the following order:
Author’s name
1) Parag Rajabhau Kulkarni
2) RuchikaDhanrajPache
Cite this Article
Parag Kulkarni, Ruchika Pache.
Depolymerization of Polypropylene: A
Systematic Review. Journal of Polymer
& Composites. 2016; 4(2): 1–16p.

Weitere ähnliche Inhalte

Was ist angesagt?

Pulp and paper mill waste
Pulp and paper mill wastePulp and paper mill waste
Pulp and paper mill waste
Rimpi Rimpy
 
Butyl and halobutyl rubber lecture 11
Butyl and halobutyl rubber lecture 11Butyl and halobutyl rubber lecture 11
Butyl and halobutyl rubber lecture 11
bistihai
 
Design.of.an.Ethanolamines.Production.Facility.Poster
Design.of.an.Ethanolamines.Production.Facility.PosterDesign.of.an.Ethanolamines.Production.Facility.Poster
Design.of.an.Ethanolamines.Production.Facility.Poster
Aaron Kirschen
 

Was ist angesagt? (20)

Renewable chemicals presentation final
Renewable chemicals presentation finalRenewable chemicals presentation final
Renewable chemicals presentation final
 
Pulp industries
Pulp industriesPulp industries
Pulp industries
 
EXTRACTION OF ESSENTIAL OILS FROM CASHEW NUTS (VEDANT LAL)ppt.pptx
EXTRACTION OF ESSENTIAL OILS FROM CASHEW  NUTS (VEDANT LAL)ppt.pptxEXTRACTION OF ESSENTIAL OILS FROM CASHEW  NUTS (VEDANT LAL)ppt.pptx
EXTRACTION OF ESSENTIAL OILS FROM CASHEW NUTS (VEDANT LAL)ppt.pptx
 
CATALYTIC PYROLYSIS OF PLASTIC
CATALYTIC PYROLYSIS OF PLASTICCATALYTIC PYROLYSIS OF PLASTIC
CATALYTIC PYROLYSIS OF PLASTIC
 
Lecture: Polymerization Reactions and Techniques
Lecture: Polymerization Reactions and TechniquesLecture: Polymerization Reactions and Techniques
Lecture: Polymerization Reactions and Techniques
 
Polypropylene (PP)
Polypropylene (PP)Polypropylene (PP)
Polypropylene (PP)
 
Poly olefins - Manufacturing & applications
Poly olefins - Manufacturing & applicationsPoly olefins - Manufacturing & applications
Poly olefins - Manufacturing & applications
 
Polypropene
PolypropenePolypropene
Polypropene
 
Polyethylene (PE)
Polyethylene (PE)Polyethylene (PE)
Polyethylene (PE)
 
Pulp and paper mill waste
Pulp and paper mill wastePulp and paper mill waste
Pulp and paper mill waste
 
Butyl and halobutyl rubber lecture 11
Butyl and halobutyl rubber lecture 11Butyl and halobutyl rubber lecture 11
Butyl and halobutyl rubber lecture 11
 
Polybutadiene
PolybutadienePolybutadiene
Polybutadiene
 
Design.of.an.Ethanolamines.Production.Facility.Poster
Design.of.an.Ethanolamines.Production.Facility.PosterDesign.of.an.Ethanolamines.Production.Facility.Poster
Design.of.an.Ethanolamines.Production.Facility.Poster
 
Styrene
StyreneStyrene
Styrene
 
CHAPTER 3 POLYMERIZATION.ppt
CHAPTER 3 POLYMERIZATION.pptCHAPTER 3 POLYMERIZATION.ppt
CHAPTER 3 POLYMERIZATION.ppt
 
Phenolic and amino resins
Phenolic and amino resinsPhenolic and amino resins
Phenolic and amino resins
 
Recovery of pulping liquors
Recovery of pulping liquorsRecovery of pulping liquors
Recovery of pulping liquors
 
Manufacturing pu foam
Manufacturing pu foamManufacturing pu foam
Manufacturing pu foam
 
Polyethylene
PolyethylenePolyethylene
Polyethylene
 
Solution polymerization technique
Solution polymerization techniqueSolution polymerization technique
Solution polymerization technique
 

Ähnlich wie Depolymerization OF PP Review Article

Role of ionic liquid [BMIMPF6] in modifying the
Role of ionic liquid [BMIMPF6] in modifying theRole of ionic liquid [BMIMPF6] in modifying the
Role of ionic liquid [BMIMPF6] in modifying the
SHALU KATARIA
 
Degradation of poly-L-lactide. Part 1, IMechE, 2004
Degradation of poly-L-lactide. Part 1, IMechE, 2004Degradation of poly-L-lactide. Part 1, IMechE, 2004
Degradation of poly-L-lactide. Part 1, IMechE, 2004
Dr Neill Weir
 
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A ReviewStudy of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
paperpublications3
 
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
Ali I. Al-Mosawi
 
lect dental-polymers.ppt including heat and cold
lect dental-polymers.ppt including heat and coldlect dental-polymers.ppt including heat and cold
lect dental-polymers.ppt including heat and cold
manjulikatyagi
 
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
ijtsrd
 
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
Shashi Kant
 
Recycling Of Polyurethane Wastes
Recycling Of Polyurethane WastesRecycling Of Polyurethane Wastes
Recycling Of Polyurethane Wastes
Ronak Vaghani
 

Ähnlich wie Depolymerization OF PP Review Article (20)

Role of ionic liquid [BMIMPF6] in modifying the
Role of ionic liquid [BMIMPF6] in modifying theRole of ionic liquid [BMIMPF6] in modifying the
Role of ionic liquid [BMIMPF6] in modifying the
 
Hgtsb6 2015 v16n1_62
Hgtsb6 2015 v16n1_62Hgtsb6 2015 v16n1_62
Hgtsb6 2015 v16n1_62
 
Degradation of poly-L-lactide. Part 1, IMechE, 2004
Degradation of poly-L-lactide. Part 1, IMechE, 2004Degradation of poly-L-lactide. Part 1, IMechE, 2004
Degradation of poly-L-lactide. Part 1, IMechE, 2004
 
Catalytic cracking of polypropylene waste over zeolite beta
Catalytic cracking of polypropylene waste over zeolite betaCatalytic cracking of polypropylene waste over zeolite beta
Catalytic cracking of polypropylene waste over zeolite beta
 
Plastic conversion copy
Plastic conversion   copyPlastic conversion   copy
Plastic conversion copy
 
Investigation of spatial configuration of Polypropylene and the influence of ...
Investigation of spatial configuration of Polypropylene and the influence of ...Investigation of spatial configuration of Polypropylene and the influence of ...
Investigation of spatial configuration of Polypropylene and the influence of ...
 
Thermal degradation kinetic study of polypropylene co-polymer (PPCP) nanocomp...
Thermal degradation kinetic study of polypropylene co-polymer (PPCP) nanocomp...Thermal degradation kinetic study of polypropylene co-polymer (PPCP) nanocomp...
Thermal degradation kinetic study of polypropylene co-polymer (PPCP) nanocomp...
 
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A ReviewStudy of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
Study of Bio-Nano Composite of Poly-Lactic Acid for Food Packaging- A Review
 
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
A new generation of cable grade poly(vinyl chloride) containing heavy metal f...
 
Light Stabilization of Polypropylene: An Independent Perspective
Light Stabilization of Polypropylene:  An Independent PerspectiveLight Stabilization of Polypropylene:  An Independent Perspective
Light Stabilization of Polypropylene: An Independent Perspective
 
lect dental-polymers.ppt including heat and cold
lect dental-polymers.ppt including heat and coldlect dental-polymers.ppt including heat and cold
lect dental-polymers.ppt including heat and cold
 
Aadrsh kumar tiwari bbau
Aadrsh kumar tiwari bbauAadrsh kumar tiwari bbau
Aadrsh kumar tiwari bbau
 
Poly(lactic acid) based biomaterials-tg-tm
Poly(lactic acid) based biomaterials-tg-tmPoly(lactic acid) based biomaterials-tg-tm
Poly(lactic acid) based biomaterials-tg-tm
 
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
The Effect of Formic Acid, Hydrogen Peroxyde and Other Conditions on Epoxidiz...
 
Click chemistry appliations in polymer science
Click chemistry appliations in polymer scienceClick chemistry appliations in polymer science
Click chemistry appliations in polymer science
 
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
Ind. Eng. Chem. Res. 2009, 48, 4866–4871_Synthesis of Ultrahigh Molecular Wei...
 
Recycling Of Polyurethane Wastes
Recycling Of Polyurethane WastesRecycling Of Polyurethane Wastes
Recycling Of Polyurethane Wastes
 
ISEF-Final
ISEF-FinalISEF-Final
ISEF-Final
 
Plastic Waste into Fuel using Pyrolysis Process
Plastic Waste into Fuel using Pyrolysis ProcessPlastic Waste into Fuel using Pyrolysis Process
Plastic Waste into Fuel using Pyrolysis Process
 
Addition polymerization, its examples and uses
Addition polymerization, its examples and usesAddition polymerization, its examples and uses
Addition polymerization, its examples and uses
 

Depolymerization OF PP Review Article

  • 1. JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 1 Journal of Polymer & Composites ISSN: 2321-2810(online), ISSN: 2321-8525(print) Volume 4, Issue 2 www.stmjournals.com Depolymerization of Polypropylene: A Systematic Review Parag Kulkarni*, Ruchika Pache Department Polymer and Surface Engineering, Institute of Chemical Technology, Mumbai-400019, Maharashtra, India Abstract Polypropylene have become an indispensable ingredient of human life and most widely used thermoplastic in commodity applications. Polymerization of propylene is carried out by using Ziegler-Natta or Metallocene catalysts which results in high molecular weights and crystalline polypropylene which are not required for several applications. Reduction in molecular weight of polypropylene in conventional reactors is uneconomical and complicated. Therefore to overcome this problem researchers and technologists proposed a post reactor operation to reduce molecular weight (depolymerization) of polypropylene. To induce modification in polypropylene different techniques of depolymerization are proposed. In this review article, we have discussed various types of polypropylene depolymerization methods with their mechanisms like oxidative, thermal, high energy radiation and chemical (free radical induced) depolymerization along with their effect on thermal, rheological, crystalline and chemical properties of the polypropylene. Keywords: Polypropylene, molecular weight, depolymerization, thermal depolymerization, rheology *Author for Correspondence E-mail: kulkarniparag05@gmail.com INTRODUCTION Polyolefins are the most widely used polymers for manufacturing of commodity plastics article. They are manufactured usually by using transition metal catalysts and have very high molecular weights. Polypropylene is an enormously adaptable member of the polyolefin family, which is due to the prochiral nature of the propylene monomer. Highly crystalline polypropylene is built up by the chain growth polymerization of propylene, a gaseous compound obtained by the thermal cracking of ethane, propane and butane or naptha fraction of petroleum. In 1951 scientists at Phillips Robert L Banks et al. found that one of their processes of converting propylene into gasoline led to the formation of whitish substance called crystalline polypropylene. Ziegler modified the equipment and prepared polypropylene in 1954. In 1954 Natta managed to synthesize polypropylene, followed shortly by Ziegler. The first industrially important crystalline, high molecular weight polypropylene was synthesized by Natta in 1955 from organo- metallic catalysts based on titanium and aluminum [1]. Polypropylene was developed by scientists at Phillips, Hoechst, Montecatini, Hercules, Farbwerke-Hoechst, by Ziegler and Natta and in separate development by standard oil of Indiana, Du Pont in 1957 [2]. Polypropylene has been widely used as injection molding plastics as well as in many extruded forms including film, fiber and split yarns [3]. Polypropylene is one of the world’s most important thermoplastic materials and it is used in numerous applications in the plastics industry. It has desirable properties such as low-cost, high-melting point, low density, appropriate process ability, high-strength, high stiffness, hinge properties, easy to handle and excellent chemical resistance [4]. Its applications have expanded continuously for the last decade. Since commodity plastics do not require extremely high mechanical properties, molecular weights and crystallinity always, highly crystalline isotactic polypropylene has been reported to be having very high melt viscosity with consequently poor flow properties and deficient in impact strength at low temperatures, mainly owing to its relatively high Tg and large spherulite dimensions. It’s moderately elevated glass transition temperature results in it being too
  • 2. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 2 brittle and hard at application temperatures below 0ºC. These high molecular weights of polypropylene give resistance to flow at elevated temperatures and hence operations of shaping by compression molding or injection molding or extrusion are unnecessarily complex. The toughness and rheological properties of polypropylene can be enhanced by reducing the molecular weight, molecular weight distribution, introduction of branching in polymer backbone and controlling average size of the spherulites in depolymerization. Control of molecular weight in the reactor, during the process of polymerization is difficult. The process parameters are set and any modification in the process like addition of a chain transfer agent would result in resetting the process, which in turn is very tedious, difficult and economically not feasible. To achieve the diversity in polypropylene grades suitable for the different applications the molecular weight and molecular weight distribution (MWD) must be tailor-made to fit the performance requirements of each application. DEPOLYMERIZATION OF POLYPROPYLENE Depolymerization leads to changes in molecular properties such as molecular weight, molecular weight distribution, thermal properties, and crystallisability, etc. which are considered to be more beneficial in mechanical properties. This foregoing defects and deficiencies can be overcome by depolymerization process for reducing the molecular weight of the polypropylene. Polypropylene is depolymerized to achieve easy processing for high-quality products in injection molding and fiber spinning with enhancing mechanical properties [5–7]. Depolymerization is the reversion of a polymer to a low molecular weight polymer fragments or splitting of polymers into molecules [8]. Reverse of the propagation step in chain polymerization is depolymerization or unzipping, which is characterized by reduction in molecular weight of the polymer to low molecular weight polymer fragments [9]. Depolymerization of polymer may be brought about by physical factor such as heat, light, mechanical stress, high energy radiation, and ultrasonication or by chemical agent such as initiator, oxygen, ozone etc. Depolymerization of the polymer is subdivided according to its various modes of initiation. There are oxidative, thermal, high energy radiation and chemical (free radical induced) modes for depolymerization of polypropylene. Lanrence has patented a process which gives a procedure of controlled thermal depolymerization of a crystalline polymer of an aliphatic mono-a-olefin which comprises treating the polymer as slurry with water, containing oxygen and including free radicals, at a temperature of 60 to 90o C and recovering from the slurry, a polymer of increased melt index [10]. The depolymerization of 16.7 weight % slurry of isotactic polypropylene dispersed in water was carried out in presence of free radical initiator and it was found that in the presence of 0.2% benzoyl peroxide at 95°C, the melt flow index increased to 13 from 9 after 1 h of reaction. Similarly in the presence of 0.3% AIBN the melt flow index increased from 0.1 to 1.2 after 2 h of reaction. It is important to note that this reduction in polymer is brought about by molecular chains scission of polypropylene. Hans et al. gave a process for reduction in border viscosities of the polymer, by treating polypropylene in fluidized bed, in presence of oxygen, within temperature range of 100–130°C. Powdered polypropylene, of border viscosity 6, was treated in an eddy furnace for 40 min and at temperature of 130°C. The resulting polymer had a border viscosity of 3 [11]. The Dow Chemical company has patented a process for reduction in molecular weight of isotactic polypropylene to improve its proccessability [12]. The process comprises of thermally depolymerizing polypropylene at temperatures of 160 to 300°C, in the presence of small amount of depolymerizing agent such as bromine or bromine containing compound (to result in a bromide radical). The treated polypropylene had a melt index improved by 2.2. Davis et al. subjected polypropylene samples to various degrees of thermal depolymerization [13]. The molecular weight distributions of these samples were experimentally determined and compared with those expected theoretically for random scission of the polymer chains. The comparison confirmed that the chain breakage was predominately random and also indicated that determination of molecular weight by viscosity average molecular weight were
  • 3. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 3 adequate for use in evaluating hypothesized depolymerization mechanisms. The most common types of depolymerization occur through chemical reactions that alter the molecular weight of the polypropylene, leading to a change in its mechanical, thermal, chemical and rheological properties. Such reactions include (a) chain scission, (b) cross- linking, (c) modification of branched chains, or (d) a combination of all of these reactions. The agent(s) initiating the depolymerization process defines the type of depolymerization. A review of the different types of depolymerizations and their initiating agents are given in Table 1. Table 1: Initiating Agents and Depolymerization Methods. Initiating Agent Depolymerization Type Heat Thermal Oxygen, Ozone Oxidative X-rays, γ-rays, electron beam radiation High energy radiation Free radical induced Chemical The chemical reactions in depolymerization process are comparatively similar, with barely trivial differences owing to variations in the initiation mechanism [14]. Depolymerization Mechanism of Polypropylene Depolymerization of polypropylene is usually done by means of free radical chain reaction mechanism shown in Figure 1 consisting of the steps of: (i) initiation (Eq. (1)), (ii) propagation (Eq. (2)) and (iii) termination (Eq. (3)). Throughout the steps, heat- facilitated hydrogen abstraction from the tertiary carbons may lead to the formation of tertiary alkyl radicals shown in Figure 1. In PP, tertiary radicals are formed mostly, due to the poorer dissociation energy of a tertiary C- H bond (ca. 373 kJ.mol-1 at 25ºC), compared to that of a secondary C-H bond (ca. 394 kJ.mol-1 at 25ºC) [19]. During the later propagation step, formed tertiary radicals will follow separate reaction paths. The reaction paths followed by tertiary radicals are depicted in schemes. Termination of polymer reactions involving radicals or macroradicals may take place [20]. Recombination is influenced by cage effects, steric control, mutual diffusion and the molecular dynamics of the polymer matrix [21–23]. Oxidative Depolymerization Free radical depolymerization of polypropylene cannot be controlled, to avoid such problems in depolymerization; oxidation on neat polymer is to work with solutions and initiator begun by Dulog, Radlman and Kerns [24]. In oxidative depolymerization, the formation of free radical sites on the chain backbone by attack of molecular oxygen, ozone etc. is done [9]. The free-radical chain reactions depicted in Figure 2 describe oxidative depolymerization of polypropylene [20, 25], consisting of individual steps of initiation, propagation and termination. During the initiation stage of oxidative depolymerization, initiating agent stimuli such as heat or UV radiation are significant for the growth of radicals. Cleavage of a polymer chain is done, due to the abstraction of a hydrogen atom from the polymer backbone (reaction 2.1). Fig. 1: Depolymerization Mechanism of Polypropylene.
  • 4. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 4 Polymer chain scission is frequently caused by severe deformation of a polymer. In the propagation steps, the radicals formed during initiation, react with molecular oxygen and are changed into macro-alkyl peroxyl radicals (reaction 2.2). Then abstraction of a hydrogen atom from another polymer molecule gives macro-hydroperoxides and more alkyl radicals (reaction 2.3). Hydroperoxides formed through this step of reaction might be cleaved homolytically to give up radical. These kinds of radicals are capable to abstract hydrogen form adjacent polymer chains ensuing to the growth of more radicals accountable for initiating depolymerization. Free radicals produced through the first two steps may also go through further reactions. Such reactions are identified as β-scission (reaction 2.5), reduce the molecular weight of chains and also persuade the crystallinity of polymers [26]. Considering as amorphous regions of semi-crystalline polypropylene depolymerized faster than the crystalline phase [27–29], the scission of polymer backbone, dependable for relating two adjoining crystalline areas, will cause a rapid devastation of the physical and mechanical properties of a polypropylene. Elevated mobility of oxygen in non-crystalline material leads to propagation within amorphous regions. The termination reactions are reliant on the molecular structure of the polymer as well as the existing depolymerization conditions and termination (reactions 2.7, 2.8) prevails [20]. Though, in the case of oxygen starvation alkyl radical dominate and bimolecular termination reactions are of superior importance. This leads to cross-linking which is evidenced by a rise in molecular weight. Oxidative cycle of polypropylene is shown in Figure 3 [4]. Fig. 2: Oxidative Free Radical Depolymerization of Polypropylene.
  • 5. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 5 Sunggyu et al. have patented a process for selectively depolymerizing isotactic PP in the oxidative depolymerization pilot plant system, at 383°C and 232 atm pressure. The batch charge varied from 0.5 to 20 gm of isotactic PP, with oxygen flow rates varying from 50 to 2000 scc/min [30]. Upadhyaya et al. patented a method for depolymerizing polyolefins like amorphous polypropylene, Fischer-Tropsch waxes, petroleum waxes, and mixtures. The process comprises of reacting the molten polymers, at atmospheric pressure, in the absence of catalysts at a temperature between about 130 and about 215o C. The molten material was constantly stirred as it was sparged with air [31]. Jansson et al. hypothesized that hydroperoxides formed during the ageing step, decompose in the subsequent processing step, thereby causing faster degradation of the materials. Hydroperoxide formation and decomposition into radicals are known to play a key role in polymer auto-oxidation [32]. Thermal Depolymerization Thermal depolymerization of polypropylene is molecular reduction as a result of overheating. At high temperatures the components of the long chain backbone of the polypropylene can begin to separate the molecular scission it involves change in molecular weight, molecular weight distribution. The conventional model for thermal depolymerization involves the major steps of depolymerization such as initiation (4.1, 4.2), propagation (4.3), branching (4.4) and termination (4.5) shown in Figure 4 [33]. Fig. 3: Oxidative Cycle of Polypropylene [4]. Fig. 4: Mechanism of Thermal Depolymerization.
  • 6. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 6 Polypropylene is very susceptible to thermal depolymerization even at normal temperature which causes chain scission; the reduced chain length reduces molecular weight. This can considerably change the mechanical properties, thermal properties and crystallinity. Ying et al. explained the mechanism of thermal oxidative and thermal mechanical degradation of isotactic polypropylene (PP) in the presence of incorporated organic peroxide. The change in MWD was explained by a random chain scission mechanism and mechanical chain scission [34]. Polypropylene subjected to multiple extrusion shows that the chain scission processes during thermo- mechanical depolymerization causes the reduction in molecular weight. This indicates that the probability of chain breaking is dependent on the depolymerization and the molecular weight of the chain [35]. Free Radical Induced Depolymerization Free-radical initiators are chemical substances that, under certain conditions, initiate chemical reactions by producing free radicals as cause of homolytically cleavage [36]. Radicals are reactive chemical species possessing a free electron and ions. Initiator-derived radicals are very reactive chemical intermediates and generally have short lifetimes [37], i.e., half- life times less than 10-9. The decomposition rate of these peroxides depends on the class of peroxide as well as on the type of alkyl group. The eight classes of organic peroxides that are produced commercially for use as initiators are listed as follows Diacyl peroxides, Dialkyl peroxydicarbonates, tert-Alkyl peroxyesters, OO-tert-Alkyl O-alkyl monoperoxy carbonates, Di (tert-alkylperoxy) ketals, Di- tert-alkyl peroxides, tert-Alkyl hydroperoxides, Ketone peroxides. Fig. 5: Depolymerization of PP in Presence of Free Radical Initiators.
  • 7. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 7 The Saule et al. studied of the modification of polypropylene based on the decomposition of unsaturated peroxides which helped to decrease the molecular weight of the polymer. Mechanisms for the branching and the depolymerization of the polypropylene were also proposed [38]. To optimize desired processing capabilities and the end use properties the controlled degradation of polypropylene (PP) using peroxides is proposed by Iedema et al. The author performed a series of controlled degradation experiments on a twin-screw extruder with polypropylene of varying MW under addition of various amounts of initiator and developed a model which was adapted to the geometry of the extruder entrance and the peroxide feed practice. Effect of thermal heterogeneities, residence time distribution, micromixing on molecular weight distribution was studied [39]. Balke et al. presents a new kinetic model for the free radical initiated degradation of polypropylene in the reactive extrusion. The peroxide used was 2,5- Dimethyl-2,5-di(t-butylperoxy) hexane. The resulting product is known to provide superior processing properties; this is an example of the production of a specialty polymer by chemical modification of a commodity polymer [40]. Possible reactions mechanism of polypropylene in presence of free radical initiators or organic peroxide is shown in Figure 5 [41]. There are several common processes for supplying the energy required to variety radicals from initiators thermal and radiation [36]. Formerly produced, radicals experience two basic types of reactions, propagation reactions and termination reactions. The initiator molecule is denoted by I and the free radical produced is shown by I*. In a propagation reaction, a radical reacts to form a covalent bond and to create a new radical. The three most ordinary propagating reactions are atom abstraction, ß-scission, and addition to carbon–carbon double bonds or aromatic rings. The polypropylene radicals produced will further react with other polymer molecules. The main importance of shift to polymer is chain branching which frequently occurs in two ways such as: i. Hydrogen might be abstracted from sites on the backbone thus transferring radical movement and initiating a long chain branch. ii. Short chain branches possibly will be fashioned by intra-molecular atom transfer (backbiting). In this case the propagating radical abstracts backbone hydrogen as of a close by carbon. In a termination reaction, two radicals interrelate in a mutually vicious reaction in which both radicals fashioned covalent bonds and reaction ceases. Termination of the polymer radicals can also be done by recombination, where two polymer radicals formed would merge. i. In which combination of two tertiary radicals would effect in cross-linking, which is slightest liable in case of polypropylene. ii. Amalgamation of two secondary, or a secondary and tertiary radical would give a polymer molecule or by disproportionate. In the provisions of the free radical depolymerization theory, intermolecular transfer of the radical was found to be the leading process. An arch linking the degree of depolymerization to the intrinsic viscosity was obtained from calculations based on a random depolymerization. With this arch, intrinsic viscosities measured as a function of depolymerization time could be used to gain the time dependence of the degree of depolymerization [13]. Depolymerization by High Energy Radiation Depolymerization of polypropylene can be done by means of irradiation methods, such as electron beams, ultrasonic, ultraviolet (UV) and γ-rays, which have been favored to modify the structure and properties of polypropylene besides using chemical initiators and free radicals. Electron beams, ultrasonic, ultraviolet radiation and γ-rays energy which possibly will cause bond scission by free radical formation for higher doses than 60 kGy for long period depolymerization occurs. The
  • 8. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 8 depolymerization of polypropylene resulting from experience to high energy radiation is an important subject for many reasons. The mechanism of high energy induced depolymerization is a free radical, one similar to the general mechanism of depolymerization as shown in Figure 6. Fig. 6: Mechanism of High Energy Induced Depolymerization PP. In case of PP irradiation it has been established that the PP undergoes essentially chain scission. Polymers are increasingly more exposed to radiation in their service where depolymerization may cause [42]. Radiation- induced depolymerization has been well reviewed [43–46]. High energy radiation is defined as all forms of radiation with energies much higher than those of chemical bonds. It includes both electromagnetic radiation (x- rays, γ-rays, etc.) and particulate beams (α- and β-particles, electrons, neutrons, etc.). The series of energies is exceptionally broad and superior to chemical bond energies. Radiation is absorbed by interaction with atomic nuclei and electron clouds. He et al. developed UV initiation based reactive procedure for depolymerization of polypropylene for the production of controlled rheology polypropylene, benzophonene has been used as the photo initiator. At low-irradiation time and low-levels of photo initiator molecular weight distribution was noticeably tailored. After depolymerization, MFR improved with drop off in viscosity and elasticity with benzophonene concentration as predictable. Crystallinity levels and crystallization temperatures of the tailored polypropylene were lower than those of virgin polypropylene because of the reduced molecular weight and narrower MWDs [47]. Radiation induced depolymerization caused by chain scission was successfully used. Controlling the degree of degradation, uniform molecular weight is friendly process are the beneficial effects of using radiation technology. When polymeric materials are irradiated by ionizing radiation, they are divided into two types, degradation (chain scission) and chain link (crosslinking). Studying the effect of hydrogen peroxide and/or γ irradiation on the degradation process of Na-alginate was investigated. It was found that the molecular weight of the polymer decreases by using gamma radiation or H2O2. However, combining both γ radiation and H2O2 accelerates the degradation rate [48]. The influence of three hindered amine light stabilizers (HALS) and two EVA copolymers on the radiation degradation processes in isotactic polypropylene (iPP) was investigated by Zimek et al. The process consequences degradation, oxidation, release of harmful gaseous products, etc. as the polymer is known as one of the most degradable upon experience to ionizing radiation. Polypropylene properties are changed even before exposure to electron beam due to structural modifications caused due to chain scission initiated by additional components [49]. The range of energies is extremely broad and higher than chemical bond energies. Radiation is absorbed by interaction with atomic nuclei and electron clouds. The radiation interaction with the electron clouds of molecules consequences in the transfer of energy to the molecules to form ions, with the elimination of a secondary electron [50, 51]. Further ionization and excitation of nearby molecules is caused by these secondary electrons having enough kinetic energy. The immediate effect of absorption of high energy radiation is construction of energetic or excited species, including trapped electrons, ions, and radical ions, results in fragmentation to provide free radicals. In polypropylene, scission of main- chain C-C bonds gives radical pairs. The scission of C-H bonds leads to comparatively stable radicals and formation of molecular hydrogen. The primary radical formed by irradiation might not be stable. For example, irradiation of polypropylene with γ-radiation at doses of 25, 50, 100 and 150 kGy produces the secondary radical [52]. In the absence of oxygen, carbon-centered radicals might
  • 9. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 9 undergo cross-linking, by recombination, chain scission, β-elimination depending on the radical structure. Polymers can consequently be divided into two main groups, those which react to radiation by cross-linking and those which drop molecular weight by chain scission [53]. Ultrasonic irradiation has been recently looked upon as a new technique for degradation of polymer compounds, mainly due to the fact that the reduction in the molecular weight is simply by splitting the most susceptible chemical bond without causing any changes in the chemical nature of the polymer. Detailed analysis of cavitations generated ultrasound radiation used for molecular weight reduction of polypropylene has been done by Desai et al. The effect of various operating parameters including initial concentration of the polymer, power density into the system and type of solvent, on the extent of depolymerization of polypropylene are investigated. Cavitations results in liquid turbulence associated with liquid shear and generation of highly reactive free radicals. The shear force leads to the rupture of chemical bonds of the polypropylene, ultrasonic sound waves produce a permanent reduction in viscosity [54]. CHARACTERIZATIONS TO STUDY DEPOLYMERIZATION As it has been mentioned before, polypropylene has extreme importance commercially, but regrettably it is also less susceptible to attack by heat, oxygen, radiation, and free radical initiator. It is very essential to study the depolymerization process i.e., to set up the mechanisms throughout oxidative depolymerization, the products formed as well as the influence of molecular weight and structure on the properties of material. A full understanding of depolymerization is of great significance, since without this, successful stabilization approaches and superior methods of lifetime prediction would not have been possible. It is, consequently, essential to consider the characterization techniques employed to study the depolymerization process. Analytical chemistry field possesses a conventional techniques that have been used for the purpose of studying polypropylene depolymerization, some of which include high temperature gel permeation chromatography (GPC), Fourier transform infrared spectroscopy (FTIR), differential scanning calorimetry (DSC) and viscometry. A brief overview on the use of some of these techniques for the purpose of studying polyolefin depolymerization will now be presented. High Temperature Gel Permeation Chromatography According to hydrodynamic volume high temperature gel permeation chromatography (HTGPC) separates molecules and is the preferred method for the determination of molecular weight and molecular weight distribution. During the depolymerization of polypropylene, changes in average molecular weight and molecular weight distribution (MWD) are often observed [55–57]. Depending on the polymer and the depolymerization methods, many radical reactions can take place. In the case of polyolefin’s, branching and recombination reactions predominate at lower temperatures, yielding a long chain branching, which leads to an increase in the average molecular weight of the material [58]. In case of polypropylene, depolymerization almost exclusively by means of chain scission, leads to a reduction of its average molecular weight [59]. Chain scission during depolymerization consequences in continuous breaking of polymer chains yielding chains of shorter lengths. The average result is that the number of short chains increases with depolymerization time and is accompanied by a broadening of the molecular weight distribution. The molecular weight curve obtained from GPC measurements, as a result, shifts towards the region of lower molar mass when chain scission is the dominant depolymerization mechanism [60–62]. As depolymerization proceeds and chain scissions continue, the molecular weight curve will exhibit bimodality with the portion of highly depolymerized short molecules appearing as a narrow distribution on the lower molecular weight side of the original material. David et al. proposed an Eq. (A) for determining the number of chain scissions in degraded samples [63], by shaping the numerical average of the initial and final number-average molecular weights.
  • 10. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 10 𝑛 𝑅 = 𝑀 𝑛𝑜 𝑀 𝑛𝑓 − 1 (A) Where, 𝑛 𝑅is the number of chain scission, 𝑀 𝑛𝑜 is initial number average molecular weight, 𝑀 𝑛𝑓is final number average molecular weight. High temperature gel permeation chromatography can be used to record molecular weight changes over a cross-section of a thick polymer sample by means of microtoming [64] or layer-by-layer milling [65]. Computer-aided study method called molecular weight distribution computer analysis (MWDCA) has also proven to be useful for the comparison of depolymerization rates in polypropylene [66–68]. This technique uses molecular weight data, obtained by evaluation of experimental GPC molecular weight distribution profiles with computer simulations for determining the scission or crosslink. Shyichuk investigated chain scission and crosslink concentrations in photo- degraded LDPE, polypropylene and ethylene- propylene copolymer. Differential Scanning Calorimetry Differential scanning calorimetry is an additional technique frequently used to obtain information on the depolymerization behavior of polymers [69, 70]. Profile of melt endotherms and changes in the peak temperatures and as well as the glass transition temperature, can provide information on the susceptibility of different crystalline phases or arrangements to depolymerization [71]. DSC is very useful in determining the oxidative induction time (OIT) [72–74], oxidative temperature (Tox) [75] and thermal stability degree of depolymerization of polymers under high temperature conditions of polymers. Rosa et al. investigated the influence of several parameters on DSC statistics and found that sample preparation (shape and size), oxygen flow and heating rate of the experiment had a considerable influence on the data obtained [70]. Fourier Transform Infrared Spectroscopy For studying out the chemical changes brought about by polymer depolymerization IR spectroscopy is one of the most accepted techniques. It has been used for qualitative as well as quantitative characterization of depolymerization products [76–80]. The depolymerization route creates the formation of different functional groups which are strongly reliant on the chemical structure of the polymer. The main chemical species detectable by infrared spectroscopy are hydroxyl and carbonyl groups [81]. The configuration of these groups usually leads to visible changes in infrared spectrum, appearing in the regions of 1850–1550 and 3700–3200 cm-1 , respectively [82–84]. Moreover, change in the FTIR absorption of depolymerization products in PE and PP can also be experimental and used for studying the difference in their oxidative depolymerization mechanisms [85]. From the hydroxyl and carbonyl groups, depolymerization products such as peroxides, alcohols, can be identified by means of derivatisation reactions [86]. Derivatisation methods for this point were first applied to polyolefins by Carlsson et al. [87]. FTIR has also been sensible in determining changes in unsaturation of PE during depolymerization [88]. The degree of depolymerization of polypropylene is determined on the basis of their carbonyl index [89–91]. The bands at 1892, 974, 2720 and 840 cm-1 have all been used as reference bands for determining the carbonyl index in polypropylene [89–92]. Viscometry Viscometry is a useful technique for determining the polymer molecular weight. The molecular weight obtained by this technique is the viscosity averaged molecular weight. The increase in the viscosity imparted by the macromolecules in solution is a direct function of the hydrodynamic volume and, hence, the molecular weight of the macromolecule. The relationship between the intrinsic viscosity and the viscosity average molecular weight is given by the semi- empirical Mark-Houwink Eq. (B) [9]. [𝜂] = 𝐾(𝑀 𝑉)⍺ (B) Where η is intrinsic viscosity, K and ⍺ are constants for a polymer-solvent system at a given temperature and Mv is viscosity average molecular weight. The η, values for the polypropylene solution were calculated by the one point intrinsic viscosity Eq. (C). [𝜂] = (2(ηsp−lnηr)) 0.5 𝐶 (C)
  • 11. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 11 Where ηr and ηsp are relative and specific viscosities respectively, C is concentration of polymer solution [93]. The viscometric studies of polymer solutions as a means of molecular characterization of polymers are well recognized and widely practiced because of simplicity in terms of experimental approach and the apparatus needed. Dilute solution viscosity can be conveniently measured in capillary viscometers such as Ostwald type or the Ubbelohde type [9]. Shenoy et al. dissolved the depolymerized polypropylene samples in xylene at 100°C and the viscosity measurement at 85°C was carried out by using the Ubbelohde viscometer. Relative (ηr) and specific viscosities (ηsp) respectively were calculated using standard formulae to determine the viscosity average molecular weight [94]. For PP-xylene system at 85°C the Mark Houwink constants are K=9.6×10–3 and α=0.63 [100]. EFFECTOF MOLECULAR WEIGHT REDUCTION Melt Flow Index The melt flow index (MFI) of the samples is considered as a critical parameter in polymer processing and industrial designs. MFI of a polyolefin resin refers to the rate at which it extrudes from a capillary die under a standard set of conditions. The MFI is reflected by the average dimensions of the molecules in a resin and their entanglements with one another so it depends on molecular characteristics (Mw and MWD) and branching characteristics, short chain branching (SCB), of the sample. Fazeli et al. have showed that the MFI of the samples is related to their branching characteristics. With increasing the degree of branching (SCB/100°C), the MFI of the sample goes through a minimum. The increase in SCB will cause more entanglements between the chains (inter-molecular entanglement) and therefore will impede the flow [101]. However, after the minimum point, increase in the SCB/100°C will cause the chains to have a more compact molecular profile (more intra-molecular entanglements instead of inter-molecular ones), so the chains will cause less hindrance to the flow of other chains. In other words, the degree of branching (number of branching per 100°C atoms of main chain) and the amount of branching (the number of chains which have the same degree of branching), have the same effect on MFI. The melting point of a-form iPP is strongly influenced by the stereoregularity [102, 103]. Melting points in the 160–168°C range are typical for commercial homopolymer samples under normal analysis conditions. Melt Strength The melt strength of a polymer is defined as the maximum force at which a molten thread can be drawn under standard conditions before it breaks. High values of MS are desired in processes where the material is stretched in its molten state, such as in film blowing, thermoforming, or foaming. Lagendijk et al. have shown how the melt strength is enhanced by the presence of strain hardening in elongational viscosity [104]. Increasing average molecular mass of a polymer results in higher shear viscosity, as well as higher melt strength. The melt strength also increases when the molecular mass distribution (MWD) becomes broader. Ghijsels et al. have demonstrated how the melt strength increases much more dramatically than shear viscosity upon the addition of long chain branches on the polymer backbone [105–107]. De Maio and Dong have studied the effect of chain structure on melt strength of polypropylene. Comparison among several linear and branched polypropylenes obtained by electron beam irradiation has shown that the melt strength of branched can be up to 10 times higher than that of linear PP with the same MFI [108]. Gotsis et al. found that the enhancement of the melt strength is related to the increase of the weight average number of long chain branches per molecule [109]. Rheology The molecular architecture of the polymer i.e. short and long chain branching and MWD affects the rheology of the melt. In a thorough review on the effect of long chain branching (LCB) on the linear viscoelasticity of polyolefins, Vega et al. showed that the introduction of LCB induces higher levels of elasticity than broadening of the MWD of a melt of linear chains of similar molecular weight [110]. On the other hand, it is current understanding that short chain branching (SCB) cannot cause large increases in
  • 12. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 12 elasticity. Yet, Vega et al. reported that SCB resulted in higher zero shear rate viscosities, higher relaxation times, higher values of the elastic storage modulus and higher activation energies of flow compared to linear polymers [111]. Those authors surmised that the number of entanglements per branch is a decisive parameter that influences shear and elongational behavior. Much less research has been done on the extensional properties of PP melts because appropriate samples have been difficult to obtain. Hingmann and Marczinke found that branched PP samples showed distinct strain hardening and concluded that a few branches on the chain had an enormous effect on the extensional behavior of the melt [112]. Crystallinity Wide angle X-ray scattering (WAXS) patterns of iPP, sPP, and aPP are shown in Figure 7 [113]. The regular molecular structure of iPP and sPP readily enables crystallization of the chains, leading to well defined crystalline reflections differing in unit cell symmetry. aPP lacks a regular molecular structure, and does not crystallize [114]. Brucker et al. have shown that the isotactic polypropylene can crystallize into three different crystal forms depending on the temperature, pressure and mechanical stress state: monoclinic a-, orthorhombic γ- and hexagonal ß-forms [115, 116]. The γ-crystals are formed only under high pressure in high molar mass homopolymer polypropylene. Polypropylene samples with low molar mass or low tacticity and polypropylene copolymers crystallize partially in γ-form. The degree of crystallinity varies between 0 for a completely amorphous material (such as aPP) and 1 for a completely crystalline material. The degree of crystallinity plays a critical role in determining properties of polypropylene. Commonly measured properties such as; modulus, yield stress, oxygen-moisture barrier resistance, and hardness increase with increasing crystallinity. In addition to tacticity, crystallinity generally increases with decreasing molecular weight (increased chain mobility), and is promoted by slower cooling rates from the melt. Busico et al. have explained how the density of iPP in the a-form varies between the limit of 100% amorphous (ϱa=0.850 to 0.855 g/cm3) and 100% crystalline (ϱc=0.936 to 0.946 g/cm3) [117]. Fig. 7: Wide-Angle X-Ray Scattering (WAXS) Patterns of iPP, sPP, and aPP [113].
  • 13. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 13 Balta-Calleja and Vonk indicated a method for evaluating crystallinity as a number by numerical means from the relationship of the peak area to the total area [118]. Jinghua et al. investigated the nonisothermal crystallization kinetics of linear and long chain branched polypropylene (LCB PP) by differential scanning calorimetry (DSC) at various cooling rates. It was shown that LCB has the role of heterogeneous nucleating agent and accelerates the crystallization process of PP. Also the activation energy of LCB PPs are higher than that of linear PP, indicating that the presence of LCB baffles the transfer of macromolecular segments from PP melt to the crystal growth surface [119]. Furthermore, the crystal morphology of linear PP and LCB PPs was observed through polarized optical microscopy (POM), and fine spherulites were observed for LCB PPs. De Nicola et al. and Wang et al. have independently investigated the effect of branching on the crystallization behaviors of PP resins [120, 121]. It was observed that the introduction of branching in the PP resins increases crystal nuclei density. This promotes faster crystallization, and, hence, higher crystallization temperatures were observed for branched materials as compared to linear materials. The cooling rate was fixed to be 10°C/min. It was observed that the branching of PP chains significantly promoted the crystallization kinetics of the PP resins by increasing the crystallization temperature about 20°C. Chemical Properties iPP is soluble in high boiling aliphatic and aromatic hydrocarbons at high temperature. The high chemical resistance of iPP results in exceptional stain resistance, and has led to the use of iPP in automobile batteries. iPP has outstanding resistance to water and other inorganic environments. iPP resists most strong mineral acids and bases, but like other polyolefins is subject to attack by oxidizing agents including 98% sulfuric acid and 30% hydrochloric acid at high temperature (13<100°C), and fuming nitric acid (ambient temperature) [122]. PP reacts with oxygen in several ways, causing chain scission and brittleness that is associated with the loss in molecular weight. This action is promoted by high temperatures, light, or mechanical stress. Several scientists have independently demonstrated treatment of polypropylene with peroxides which has led to controlled rheology resins with reduced molecular weight and narrow polydispersity relative to polymerized product from Ziegler-Natta catalysts [123– 125]. These resins are used in some fiber spinning and injection molding applications. The creation of radical sites along the polymer backbone, most often through peroxide-based initiation, is also an essential condition for many fictionalization/grafting chemistries. CONCLUSION It is concluded from the foregone discussion that understanding about the mechanism can go a long way in helping the researchers and the technologists to induce the different types of depolymerization methods in the polypropylene. This depolymerization can further be enhanced by the addition of the initiator, additives in the polypropylene and by understanding the various factors such as mechanical, physical, and chemical which are responsible for this depolymerization. It is also concluded from this discussion that polypropylene depolymerization could be enhanced its properties with modification in structure. REFERENCES 1. Natta G, Pino P, Corradini P. J Am Chem Soc. 1955; 77. 2. Seymour RB. History of Polyolefins. In: Vasile C, Seymour RB, editors. Handbook of Polyolefins. New York: Marcel Dekker; 1993; 1p. 3. Qicong Ying, Yong Zhao, Yong Liu. Makromol Chem. 1991; 192. 4. Denis Bertin, Marie Leblanc, Marque Sylvain RA, et al. Polym Degrad Stabil. 2010; 95. 5. Imperial Chemical Industries Ltd. Germ Pat. 1210562. Mar 17, 1961. 6. Chemische Werke Hls AG. French Pat. 1377951. Nov 6, 1964. 7. Caver Hill R, Taylor GW. Polymer. 1965; 6. 8. Gooch Jan W. Encyclopedic Dictionary of Polymers. 2nd Edn. Springer; 2011. 9. Gowarikar VR, Vishwanathan NV, Sreedhar Jayadev. Polymer Science.
  • 14. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 14 Reprint Edn. India: New Age Publishers; 1994. 10. Laurence Stephen Rayner. Imperial Chemical Industries Ltd. US Pat. 3232918. Feb 1, 1966. 11. Boehm Hans, Henglein Friedrich, Schmidt-Thomee, et al. BASF AG, DE 1201552. Sep 1965. 12. Saunders Frank Linwood. Dow Chemical Company, DE 1221798. Jul 1966. 13. Thomas Davis E, Robert Tobias L, Peterli Elizabeth B. J Poly Sci. 1962; 56. 14. Wood DL, Luongo JP. Mod Plast. 1961; 38. 15. Williams SD, Yoo HJ, Drickman MR. US Patent 5820981. 1998. 16. Lagendijk RP, Hogt AH, Buijtenhuijsand A, et al. Polymer. 2001; 42. 17. Sugimoto M, Tanaka T, Masubuchi Y, et al. J Appl Polym Sci. 1999; 73. 18. De Nicola Jr., Anthony J. US Patent 5047446. Sep 01, 1991. 19. Gonzalez-Gonzalez VA, Neira- Velazquez G, Angulo-Sanchez JL. Polym Degrad Stabil. 1998; 60. 20. Al-Malaika S. Adv Polym Sci. 2004; 169. 21. Garton A, Carlsson DJ, Wiles DM. J Polym Sci Polym Chem Ed. 1978; 16. 22. Garton A, Carlsson DJ, Wiles DM. Macro. 1979; 12. 23. Al-Malaika S. Polym Plast Technol Eng. 1988; 27. 24. Dale Van Sickle E. J Polym Sci. 1972; 10(a-1). 25. George GA, Celina M, Hamid SH. Handbook of Polymer Degradation. 2nd Edn. New York: Marcel Dekker Inc.; 2000. 26. Rabello MS, White JR. Polymer. 1997; 38. 27. Blais P, Carlsson DJ, Wiles DM. J Polym Sci. 1972; 10(A1). 28. Schlick S, Kruczala K. JCT Res. 2005; 2. 29. Knight JB, Calvert PD, Billingham NC. Polymer. 1985; 26. 30. Lee Sunggyu, Gencer Mehmet A, Fullerton Kathy L, et al. US Pat. 5386055. Jan 31, 1995. 31. Upadhyaya Janardan D. US Pat. 4624993. Nov 25, 1986. 32. Anna Jansson, Kenneth Moller, Thomas Hjertberg. Polym Degrad Stabil. 2004; 84. 33. Thermal Degradation of Polymers a Primer. Pop Plast Packag. 2013; 12. 34. Qicong Ying, Yong Zhao, Yong Liu. Makromol Chem. 1991; 192. 35. Sebastiaao V, Canevarolo. Polym Degrad Stabil. 2000; 709. 36. Myers Terry N, Mark Herman F. Encyclopedia of Polymer Science and Technology. 3rd Edn. John Wiley and Sons; 2004. 37. Griller D, Ingold KU. Acc Chem Res. 1976; 9. 38. Myriam Saule, Laurence Moine, Marie Degueil-Castaing, et al. Macromolecules. 2005; 38. 39. Iedema PD, Remerie NK, Vanderhamb BM, et al. Chem Eng Sci. 2011; 66. 40. Balke ST, Suwanda D, Lew R. J Polym Sci: Part C, Polym Lett. 1987; 25. 41. Imoto Minoro, Takemoto Kiichi, Kono Masatsugu. Angew Makromol Chem. 1967; 1. 42. Clough RL, Gillen KT. Nuc Techno. 1982; 59. 43. David C. Degrad Polym. 1975; 14. 44. Schnabel WS, Jellinek HHG. Degradation and Stabilization of Polymers. Amsterdam: Elsevier; 1978. 45. Schnabel WS. Polymer Degradation: Principles and Practical Applications. Munich: Hanser; 1981. 46. Ivanov VS. Radiation Chemistry of Polymers. Amsterdam: VSP; 1992. 47. Guangjian He, Costas Tzoganakis. Polym Eng Sci. 2011. 48. Hegazy EA, Abdel-Rehim H, Diaa DA, et al. Controlling of Degradation Effects in Radiation Processing of Polymers. International Atomic Energy Agency; 2009. 49. Zimek Z, Przybytniak G, Rafalski A, et al. International Atomic Energy Agency; 2009. 50. Swallow A. Radiation Chemistry. London: Longman; 1973. 51. Spinks JWT, Woods RJ. An Introduction to Radiation Chemistry. New York: John Wiley and Sons, Inc.; 1976. 52. Krupa I, Luyt AS. Polym Degrad Stabil. 2001; 72.
  • 15. Journal of Polymer & Composites Volume 4, Issue 2 ISSN: 2321-2810(online), ISSN: 2321-8525(print) JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 15 53. Bilingham NC. Degradation. In: Mark HF, editor. Encyclopedia of Polymer Science and Technology. Vol. 6. John Wiley and Sons; 2005. 54. Vaibhav Desai, Shenoy Mohan A, Gogate Parag R. Chem Eng J. 2008; 140. 55. Fayolle B, Audouin L, Verdu J. Polym Degrad Stab. 2002; 75. 56. Gonzalez-Gonzalez VA, Neira- Velazquez G, Angulo-Sanchez JL. Polym Degrad Stab. 1998; 60. 57. Lew R, Suwanda D, Balke ST. J Appl Polym Sci. 1988; 35. 58. Yadav A, Vimal KK, Singh R. J Polym Composites. 2016; 4(1). 59. Ying Q, Zhao Y, Liu Y. Macromol Chem. 1991; 192. 60. Alam MS, Nakatani H, Goss BGS, et al. J Appl Polym Sci. 2002; 86. 61. Manabe N, Yokota H, Suzuki S, et al. J Appl Polym Sci. 2006; 100. 62. Nakatani H, Manabe N, Yokota Y, et al. Polym Int. 2007; 56. 63. David C, Trojan M, Daro A, et al. Polym Degrad Stab. 1992; 37. 64. Girois S, Audouin L, Verdu J, et al. Polym Degrad Stab. 1996; 51. 65. Turton TJ, White JR. Polym Degrad Stab. 2001; 74. 66. Shyichuk AV, White JR, Craig IH, et al. Polym Degrad Stab. 2005; 88. 67. Shyichuk AV, White JR. J Appl Polym Sci. 2000; 77. 68. Shyichuk AV, Stavychna DY, White JR. Polym Degrad Stab. 2001; 72. 69. Olivares N, Tiemblo P, Gomez-Elvira JM. Polym Degrad Stab. 1999; 65. 70. Rosa DS, Sarti J, Mei LHI, et al. Polym Test. 2000; 19. 71. Elvira M, Tiemblo P, Gomez-Elvira JM. Polym Degrad Stab. 2004; 83. 72. Camacho W, Karlsson S. Polym Degrad Stab. 2002; 78. 73. Chang TC, Yu PY, Hong YS, et al. Polym Degrad Stab. 2002; 77. 74. Groning M, Eriksson H, Hakkarainen M, et al. Polym Degrad Stab. 2006; 9. 75. Ahlblad G, Gijsman P, Terselius B, et al. Polym Degrad Stab. 2001; 73. 76. Gulmine JV, Janissek PR, Heise HM, et al. Polym Degrad Stab. 2003; 79. 77. Edge M. Infrared Spectroscopy in Analysis of Polymer Degradation. In: Meyers RA, editor. Encycl Anal Chem. Chichester: John Wiley and Sons Ltd.; 2000. 78. Gugumus F. Polym Degrad Stab. 1997; 55. 79. Nekhoroshev VP, Turov YP, Nekhorosheva AV, et al. J Appl Chem. 2006; 79. 80. Costa L, Luda MP, Trossarelli L. Polym Degrad Stab. 1997; 58. 81. Andreassen E, Karger-Kocsis J. Infrared and Raman Spectroscopy of Polypropylene. Dordrecht: Kluwer Publishers; 1999. 82. Mani R, Singh RP, Sivaram S. Polym Int. 1997; 44. 83. Rivaton A, Gardette JL, Mailhot B, et al. Macromol Symp. 2005; 225. 84. Gugumus F. Polym Degrad Stab. 1999; 6. 85. Adams JH, Goodrich JE. J Polym Sci. 1970; Part A1(8). 86. Piton M, Rivaton A. Polym Degrd Stab. 1996; 53. 87. Carlsson DJ, Brousseau R, Zhang C, et al. ACS Symp Ser. 1988; 364. 88. Kolbert AC, Didier JG, Xu L. Macromolecules. 1996; 29. 89. Jansson A, Moller K, Gevert T. Polym Degrad Stab. 2003; 82. 90. Miraftab M, Horrocks AR, Mwila J. Polym Degrad Stab. 2002; 78. 91. Santos ASF, Agnelli JAM, Trevisan DW, et al. Polym Degrad Stab. 2002; 77. 92. Rabello MS, White JR. Polym Degrad Stab. 1997; 56. 93. Solomo O, Ciuta ZZ. J Appl Polym Sci. 1962; 6. 94. Shenoy MA, Mrinalini Patil. Polym Sci. 2010; 52. 95. Feng-Hua Su, Han-Xiong Huang. Polym Eng Sci. 2009; 50(2). 96. Hamed Azizi, Ismaiel Ghasemi. Iran Polym J. 2005; 14(5). 97. Manfred Raetlsch, Achim Hesse, Harmut Bucka, et al. US Pat. 6136926. Oct 24, 2000. 98. Horst David E, Michael Roth, Peter. US Pat. 2007/0200272. Aug 30, 2007.
  • 16. Depolymerization of Polypropylene Kulkarni and Pache JoPC (2016) 1-16 © STM Journals 2016. All Rights Reserved Page 16 99. Gaylord Norman G. US Pat. 4506056. Mar 19, 1985. 100. Brandrup J, Immergut EH, Grulke EA. Polymer Handbook. New York: Wiley; 1999. 101. Fazeli N, Arabi H, Bolandi. Polym Test. 2006; 25. 102. Wunderlich B. Macromolecular Physics. 1980; 3. 103. Paukkeri R, Lehtinen A. Polymer. 1993; 34. 104. Lagendijk RP, Hogt AH, Buijtenhuijs A, et al. Polymer. 2001; 42. 105. Ghijsels A, Ente JSM, Raadsen J. Int Polym Process. 1990; 5. 106. Ghijsels A, Clippeleir De J. Int Polym Process. 1994; 9. 107. Ghijsels A, Massardier CHC, Bradley RM. Int Polym Process. 1997; 12. 108. De Maio VV, Dong D. Proceedings SPE ANTEC. 1997; 43. 109. Gotsis AD, Zeevenhoven BLF, Hogt AH. Polym Eng Sci. 2004; 44. 110. Vega J, Aguilar M, Peon J, et al. e- Polymers. 2002; 46. 111. Vega JF, Santamaria A, Munoz-Escalera A, et al. Macromolecules. 1998. 112. Hingmann R, Marczinke BL. J Rheol. 1994; 38. 113. Phillips RA, Wolkowicz MD, Moore EP. Polypropylene Handbook. Munich: Hanser; 1996. 114. Natta G, Corradini P. Del Nuovo Cimento. 1960; 15. 115. Takodoro H. Structure of Crystalline Polymers. USA: John Wiley and Sons; 1979. 116. Brucker S, Meille V, Petraccone, et al. Prog Polym Sci. 1991; 16. 117. Busico V, Corradini P, Debiasio R, et al. Macromolecules. 1994; 27. 118. Balta-Calleja FJ, Vonk CG. X-ray Scattering of Synthetic Polymers. Amsterdam: Elsevier; 1989. 119. Tian Jinghu, Yu Wei, Zhou Chixing. J Appl Polym Sci. 2007; 104. 120. De Nicola AJ, Galombos AF, Wolcowicz MD. Polym Mater Sci Eng. 1992; 67. 121. Wang X, Tzoganakis C, Rempel GL. J Appl Polym Sci. 1996; 61. 122. Pro-fax Polypropylene Chemical Resistance. Montell Polyolefins Technical Bulletin. 1996; TL-101. 123. Zeichner G, Patel P. 2nd World Congress of Chemical Engineering, Montreal, P.Q., Canada. 1981. 124. Dziemianowicz TS, Cox WW. SPE ANTEC. 1985; 85. 125. Gahleitner M, Wolfschwenger J, Bachner C, et al. J Appl Polym Sci. 1996; 61. Copyright Notice Declaration & Copyright Transfer Form I, the undersigned author(s) of the submitted manuscript, hereby declare, that the above manuscript which is submitted for publication in the STM Journals(s), is not published already in part or whole (except in the form of abstract) in any journal or magazine for private or public circulation, and, is not under consideration of publication elsewhere.  I will not withdraw the manuscript after 1 week of submission as I have read the author guidelines and will adhere to the guidelines.  I have neither given nor will give this manuscript elsewhere for publishing after submitting in STM Journal(s).  I have read the original version of the manuscript and am responsible for the thought contents embodied in it. The work dealt in the manuscript is my/our own, and my/our individual contribution to this work is significant enough to qualify for authorship.  I also agree to the authorship of the article in the following order: Author’s name 1) Parag Rajabhau Kulkarni 2) RuchikaDhanrajPache Cite this Article Parag Kulkarni, Ruchika Pache. Depolymerization of Polypropylene: A Systematic Review. Journal of Polymer & Composites. 2016; 4(2): 1–16p.