Sei sulla pagina 1di 16

Trends in Analytical Chemistry, Vol. 28, No.

9, 2009

Trends

Environmental biodegradation
of synthetic polymers I. Test
methodologies and procedures
Jan P. Eubeler, Marco Bernhard, Sabine Zok, Thomas P. Knepper
Biodegradation of synthetic polymers is an important property that is used in
many applications. Evaluation of the extent of biodegradation has used different methods in recent years. For each environmental compartment, different approaches have to be made in order to obtain valuable data on
biodegradability.
This review describes validated and accepted methods based on standardized biodegradation tests, analytical tests, enzymatic tests or tests of
physical properties to evaluate the biodegradability of synthetic polymers for
different types of environmental compartments (e.g., soil, compost or aqueous media).
Part II of this review will subsequently report on the environmental biodegradation of different groups of synthetic polymers.
2009 Elsevier Ltd. All rights reserved.
Keywords: Analysis; Aqueous medium; Biodegradation; Compost; Environmental
compartment; Enzymatic test; Methodology; Soil; Standardized biodegradation test;
Synthetic polymer

1. Introduction
Jan P. Eubeler,
Sabine Zok
BASF SE, Experimental
Toxicology and Ecology,
GV/TC Z-570, D-67056
Ludwigshafen,
Germany
Marco Bernhard,
Thomas P. Knepper*
University of Applied Sciences
Fresenius, Limburger Strae 2,
D-65510 Idstein,
Germany

Corresponding author.
Tel.: +49 (0)6126-935264;
Fax: +49 (0)6126-935210;
E-mail:
knepper@hs-fresenius.de

Biodegradation can be a key feature of


synthetic polymers within the frame of
sustainable development. Polymers are
widely used and our daily life could not be
imagined without them. Polymers are
macromolecules generally with molecular
weight (MW) >1000 g/mol with different
consecutive repeating monomer units
combined in different ways to deliver
thousands of totally different products.
In the beginning, polymers were mostly
not biodegradable, but, today, certain
applications require specialized products
with real biodegradability because they
may otherwise accumulate in the environment.
In biodegradation research, assimilation
into biomass has been especially neglected
and the tools have been inadequate to
certify real biodegradability and integration of the material in biogeochemical life
cycles [1]. The concepts have been applied

0165-9936/$ - see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.trac.2009.06.007

and improved since the early 1970s, and


conventional tests and assessments have
been done since then for different groups
of chemical substances.
For polymers, people have been focusing
mainly on soil or compost biodegradation
[2,3]. By contrast, there is almost no systematic information on aqueous environments. This area can be important
because water-soluble polymers [e.g.,
polyvinyl pyrrolidones (PVPs) or polyethylene glycol (PEG) for PCPs, lacquers and
dyes, or kinetic hydrate inhibitors (KHIs)
for oil-field and gas-field applications] are
used in especially large amounts today.
In polymer research, biodegradation is a
useful property to obtain plastics for certain applications where biodegradation
enhances the value of an application (e.g.,
mulching films [4], food-packaging materials [5] or polymers used in oil-field and
gas-field chemicals or PCPs). The value of
the application increases because the
products may have lesser or no environmental risk when biodegradable. The literature in recent years has described
biodegradable polymers [6]. Synthetic
manufacture of bioplastics was investigated using natural and petrochemical or
fossil-fuel resources [7].
Generally, two different types of polymer can be degraded. The first are basically converted to carbon dioxide and
water in an appropriate environment
[e.g., cellulose, starch (ST) and aliphatic
polyesters] and the second comprises the
class of oxo-biodegradable polymers (natural rubber and lignocelluloses and some
oxo-polyolefins) that are broken down
(disintegrated) but not mineralized in the
main.

1057

Trends

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Generally, non-stabilized polyolefins oxidize rapidly in


the environment [e.g., polyethylene (PE) 20% weight
loss in 5 months; polypropylene (PP) 80% weight loss in
5 months]. Only their formulation and additional compounds (e.g., antioxidants) give them stability [8].
It is often assumed that bioplastics are, by definition,
environmentally friendly or sustainable. As life-cycle or
eco-efficiency analyses show, that may not always be the
case. If an assessment of environmental (and ecological)
facts for each stage of the life-cycle of a product
regarding natural resources, consumed energy (from
generally fossil fuels) and waste produced is done, relevant comparative information will be obtained [8].
In many different ways, renewable resources seemed
to be a solution for problems arising from common
polymers (e.g., unknown environmental fate and recalcitrance) [9]. But the overall performance in the complete life-cycle of a product determines whether natural

resources or petrochemical resources offer the best, most


effective source of energy and material [7,10] and directs
the development. The different stages of biodegradation
(biodeterioration, biofragmentation and assimilation)
have to be specifically investigated and validated because
changes can be observed when environmental conditions change. In Part II of this article, we will specifically
report on different groups of polymers and their potential
to biodegrade in natural environments by linking
material properties to biodegradation [11].

2. Assessing biodegradation with accepted


methods
Standardized methods for testing biodegradation of
plastics were introduced by the American Normative
Reference (ASTM), Japanese Industrial Standards (JIS)

Table 1. Standardized ASTM test methods for polymer-biodegradation tests


ASTM code
D5209-92
D5210-92
D5247-92

D5271-02

D5338-92

Purpose

Involved microorganisms and key features

Aerobic degradation of plastic materials


in municipal sewage sludge
Anaerobic degradation of plastic
materials in municipal sewage sludge
Aerobic biodegradability of degradable
plastics by specific microorganisms

Indigenous microorganisms in sewage sludge

CO2 evolved

OECD 301

Indigenous microorganisms in sewage sludge

CO2 and CH4


evolved
Weight loss, tensile
strength, elongation
and molecular
weight distribution
Oxygen
consumption

OECD 311

Aerobic biodegradation of plastic


materials in activated sludge and
wastewater
Aerobic biodegradation in composting
conditions

G21-90

Resistance to fungi

G22-76
D6691-01

Resistance to bacteria
Determination of aerobic biodegradation
of plastics in marine environment

D6692-01

Determination of biodegradation of
radiolabeled polymeric plastics in
seawater
Determination of aerobic biodegradation
of radiolabeled plastic materials in
aqueous or compost environments
Determination of weight loss from plastic
materials exposed to simulated municipal
solid-waste aerobic compost
environment
Standard specification for Non-floating
biodegradable plastics in the marine
environment

D6340-98

D6003-96

D7081-05

1058

http://www.elsevier.com/locate/trac

Streptomyces badius ATCC39117


Streptomyces setonii ATCC39115
Streptomyces viridosporus ATCC39115 or
other organisms agreed upon
Municipal sewage treatment plant sludge

24 months old compost

Aspergillus niger ATCC9642 Aureobasidium


pullulans ATCC15233 Chaetomium
globosum ATCC6205 Gliocladium virens
ATCC9645 Penicillum pinophilum
ATCC11797
Pseudomonas aeruginosa ATCC13388
Inoculum consists of a minimum of nine test
organisms. Marine solution is prepared in the
lab following a standard procedure. Inoculum
is verified by standard identification test
Most ASTM methods use natural sea water as
inoculum. No standardized method is
available. See also D7081-05
Natural mixed culture or compost matrix

Simulated compost, commercial compost


seed

Parameters
monitored

Similar to

OECD 303

Cumulative CO2
production, DMR,
CMR, GMR
Visual evaluation

Visual evaluation
CO2 evolved

OECD 306,
ISO 16221

CO2 evolved, DPM


counts (LSC),
specific radioactivity
CO2 evolved, DPM
counts (LSC),
specific radioactivity
Weight loss

OECD
Code
Tests on ready
biodegradability

301A

Static aerobic aquatic test using standard


conditions. (DOC-Die-Away-Test)

301B

CO2-Evolution Test. Static aerobic aquatic


test using standard conditions. Measurement
of biogenically evolved CO2 and comparison
to ThCO2

301C

MITI-I. designed for use in Japan. With special


inoculum preparation and obligatory specific
analysis
Closed bottle test. Static aerobic aquatic test
system using standard conditions

301D

http://www.elsevier.com/locate/trac

Tests on inherent
biodegradability

Purpose

301E

Modified test. Static aerobic aquatic test using


standard conditions with low bacteria
concentration (e.g. river water)

301F

Respirometric test. Static aerobic aquatic test


using standard conditions. Comparison of
BOD to ThBOD or COD

302A

Semi continuous activated sludge test (SCAS).


Semi static aerobic aquatic test system using
organic test compounds and easily
biodegradable organic medium
Static test (Zahn-Wellens test). Static aerobic
aquatic test using standard conditions

302B

302C

Standardized defined inorganic test medium


and mixed microorganisms. Aerated and
stirred test through 28 days and 3-4 samplings
a week. Non volatile and not significantly
adsorbable test compounds. Water-solubility
at 1040 mg/L DOC
Standardized defined inorganic test medium
and mixed microorganisms. Aerated and
stirred test through 28 days and 3-4 samplings
a week. For Polymers a modification with
higher buffer capacity, higher test
temperatures and longer test duration can be
used

Parameters monitored

Similar to

DOC removal. Comparison of


DOC(start) to DOC(end)

ISO 7827
EU 92/69/EWG C.4A

CO2 evolution

ISO 9439
EU 92/69/EWG C.4C

Only for Japanese region


EU 92/69/EWG C.4F
Oxygen supply from test water. Oxygen
measurement with electrode. Low inoculum
concentration. Comparison of BOD with
ThBOD or COD. DOC removal
determination possible
Standardized defined inorganic test medium
and mixed microorganisms. Aerated and
stirred test through 28 days and 3 or 4
samplings a week
Test compounds which are water-soluble or
insoluble at the test conc. of 100 mg/L
substance or ThOD

Daily fill and draw of test vessel. Watersoluble non volatile not significantly
adsorbable organic compounds at
concentration of 2050 mg/L DOC
Higher concentration of test substance and
activated sludge (up to 1 lg/L dry substance).
May be difficult do distinguish between
abiotic elimination by adsorption and
biodegradation. For water-soluble non
volatile organic compounds 50400 mg/L
DOC

BOD measurement. in completely


filled bottles

ISO 10707
EU 92/69/EWG C.4E

DOC removal. Comparison of


DOC(start) to DOC(end)

ISO 7827
EU 92/69/EWG C.4B

BOD measurement. In closed


respirometer. Add. information for
water-soluble compounds by DOC
removal

ISO 9408
EU 92/69/EWG C.4D

DOC measurement before and after


replacement to determine ultimate
biodegradation within a test time of
26 weeks
DOC removal. Comparison of
DOC(start) to DOC(end)

ISO 9887
EU 88/302/EWG C

ISO 9888
EU 88/302/EWG C

Only for Japanese


Government

(continued on next page)

Trends

1059

MITI-II same as MITI-I (OECD 301C) but with


different test and inoculum concentration to
improve biodegradability

Microorganisms involved and key features

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Table 2. Standardized general OECD test strategy for biodegradability testing

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

1060

Substance-specific analysis,
radiolabeled analytes
Substance-specific analysis,
radiolabeled analytes
Substance-specific analysis,
radiolabeled analytes
High environmental relevance. Soil system
from the environment
High environmental relevance. Sediment/
water system from the environment
High environmental relevance. Surface water
system from the environment

Test compounds at concentrations of 2


40 mg/L DOC

Evolved 14CO2 determination by


alkali adsorption and liquidscintillation counting
DOC removal. Comparison of
DOC(start) to DOC(end)
Incubation time up to 64 days

http://www.elsevier.com/locate/trac

309

308

307

306

304
Other tests

Biodegradability in soil. Static test using soil


as medium and inoculum in a closed system
Radiolabeled test substances
Marine biodegradation. Static aerobic test
using sea water. Shake flask (60 days) or
closed bottle test with 28-day test duration
Aerobic and anaerobic transformation in soil
systems
Aerobic and anaerobic transformation in
aquatic sediment systems
Aerobic mineralization in surface water
simulation biodegradation test

303
Simulation test

OECD
Code

Table 2 (continued)

Purpose

Activated sludge simulation test.


Continuously operated aerobic aquatic test
system using organic test compounds and
easily biodegradable organic medium

Microorganisms involved and key features

Water-soluble or satisfactorily dispersible


non-volatile organic compounds at
concentrations of normally 1020 mg/L DOC

Parameters monitored

DOC or COD measurement in


influent and effluent of test and blank.
Test time of 12 weeks

Similar to

ISO 11733
EU 88/302/EWG C

Trends

and European Normative Reference (ECN) and Organization for Economic Co-operation and Development
(OECD) for solid as well as liquid media [1216]. Tables
13 show the testing strategy used today in biodegradation research and evaluation. The advantages and the
limitations of accepted methods differ because of the
different structures of polymers. Also, most guidelines
were not exclusively designed for polymer biodegradation and have to be adapted for these special requirements so consideration needs to be given first to the
properties of the tested polymers (e.g., solubility, molecular structure and physico-chemical parameters). The
environmental compartment and microorganisms (MOs)
greatly influence the results. The major downside is that
experiments with polymers often take many days,
months or years, but the tests are mainly developed for a
period of 28100 days, depending on the test guideline.
Analytical procedures need to be assigned individually
taking into account the special requirements of the
polymer and the question addressed by the research. So
far, biodegradation tests for polymers are very particular
and standardized only to a little extent.
The current situation offers limited flexibility in
methodologies for testing biodegradation of polymers
because methods are developed based on material types
or the environment of application. For example, testing
microbial degradation generally offers only a small
variety of fungal and bacterial species or uses only a
small number of environments for testing. Finding the
same small variety of environmental conditions is hardly
possible [17]. Because there are many different metabolic
pathways in natural environments, we can also find
many different biodegradation processes that reflect
again a variety of different types of test methods (e.g.,
static, semi-continuous and continuous systems, anaerobic and aerobic, limnic and marine) [18].
Different tests for determination of the biodegradability
of water-insoluble polymers are described in the literature, ranging from simple die-away tests or biological
oxygen demand (BOD) tests to combined CO2-dissolved
organic carbon (DOC) tests [1921]. To make test results
comparable internationally, the tests are best carried out
following internationally recognized (OECD) or standardized (ISO or CEN) methods [18]. Methods are designed for certain environmental conditions [17] and
require specifically designed and characterized environments or certain strains of established cultures of MOs
(e.g., ISO 14851 and ISO 14852 methods are especially
designed for aerobic biodegradability determination of
polymeric and plastic materials and their additives in
aquatic batch tests). The sample is exposed to an inoculum under laboratory conditions in aqueous solution.
In these cases, the inoculum is either activated sludge or
a compost suspension [22]. They may also be divided
into three categories of OECD methods (e.g., tests for
ready biodegradation, tests for inherent biodegradation

ISO
code
Tests on ready
biodegradability

Tests on inherent
biodegradability

http://www.elsevier.com/locate/trac

Simulation test

Purpose

Involved microorganisms and key features

Parameters monitored

Standardized defined inorganic test medium


and mixed microorganisms. Aerated and
stirred test through 28 days and 3 or 4
samplings a week. Non-volatile and not
significantly adsorbable test compounds.
Water-solubility at 1040 mg/L DOC
Test compounds which are water-soluble or
insoluble at the test concentration of 100 mg/
L substance or ThOD

DOC removal. Comparison


of DOC(start) to DOC(end)

OECD 301 A

BOD measurement. In
closed respirometer. Add.
information for water-soluble
compounds by DOC
removal
CO2 evolution

OECD 301 F

BOD measurement. In
completely filled bottles

OECD 301 D

DOC measurement before


and after replacement to
determine ultimate
biodegradation within a test
time of 26 weeks
DOC removal. Comparison
of DOC(start) to DOC(end)

OECD 302 A

DOC or COD measurement


in influent and effluent of test
and blank. Test time of 12
weeks

OECD 303 A

7827

Static aerobic aquatic test using standard


conditions. (DOC-Die-Away-Test)

9408

Respirometric test. Static aerobic aquatic test


using standard conditions. Comparison of
BOD to ThBOD or COD

9439

CO2-Evolution Test. Static aerobic aquatic


test using standard conditions. Measurement
of biogenically evolved CO2 and comparison
to ThCO2

10707

Closed bottle test. Static aerobic aquatic test


system using standard conditions

9887

Semi continuous activated sludge test (SCAS).


Semi static aerobic aquatic test system using
organic test compounds and easily
biodegradable organic medium

Daily fill and draw of test vessel. Watersoluble non volatile not significantly
adsorbable organic compounds at
concentrations of 2050 mg/L DOC

9888

Static test (Zahn-Wellens test). Static aerobic


aquatic test using standard conditions

Higher concentration of test substance and


activated sludge (up to 1 lg/L dry substance).
May be difficult to distinguish between
abiotic elimination by adsorption and
biodegradation. For water-soluble, nonvolatile organic compounds of 50400 mg/L
DOC

11733

Activated sludge simulation test.


Continuously operated aerobic aquatic test
system using organic test compounds and
easily biodegradable organic medium

Water-soluble or satisfactorily dispersible


non-volatile organic compounds at
concentrations of normally 1020 mg/L DOC

Standardized defined inorganic test medium


and mixed microorganisms. Aerated and
stirred test through 28 days and 3 or 4
samplings a week. For polymers a
modification with higher buffer capacity,
higher test temperatures and longer test
duration can be used
Oxygen supply from test water. Oxygen
measurement with electrode. Low inoculum
concentration. Comparison of BOD with
ThBOD or COD. DOC removal
determination possible

Similar to

OECD 301 B

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Table 3. Standardized ISO test strategy for biodegradability testing

OECD 302 B

1061

Trends

(continued on next page)

http://www.elsevier.com/locate/trac

ISO
code
Other tests

Purpose

10634

Guidance for poorly water-soluble test


compounds

10708

Two-phase-closed-bottle-test. Static aerobic


aquatic test using standard conditions

11734

Anaerobic biodegradability. Static aquatic test


using anaerobic digested sludge medium and
inoculum
Low concentration water test. Static (shake
flask method) or dynamic (river simulation)
aerobic surface water system

14592

CO2-Headspace test. Static aerobic aquatic


test using standard conditions. Gas-tight
closed bottles with sufficient oxygen

14851

Test modification of MITI-I & II. Used for


polymer testing

14852

CO2-Evolution Test. Static aerobic aquatic


test using standard conditions

14855

Aerobic composting test. Static aerobic test

15462

Guidance for selection of biodegradation tests

16221

Water quality Guidance for determination


of biodegradability in the marine environment

Description of several techniques for


preparation of poorly water-soluble organic
compounds and introduction to the test vessel
with aqueous media
Closed test system with sufficient oxygen in
headspace. Oxygen measurement with
electrode. Comparison of BOD with ThBOD
or COD. DOC removal determination
possible
Test duration 60 days and test concentration
at 20100 mg/L organic carbon
Primary biodegradability and DEG kinetics of
test substance at realistic environmental
conditions. Substances known principally
degradable and water-soluble at test
concentrations. Non-volatile and suitable
analytical method is necessary
Measurement of biogenically-evolved CO2
after acidification in the gas phase or
alkalization in form of DIC and comparison to
ThCO2
Higher buffer capacity, test temperature and
longer test duration. Carbon balance is
possible
Especially for polymers: a modification with
higher buffer capacity, higher test
temperatures and longer test duration can be
used. Measurement of biogenically-evolved
CO2 and comparison to ThCO2
Determination of ultimate biodegradability of
test material in an optimized simulation of an
intensive aerobic composting process at
about 60C. Duration 45 days. Optional
disintegration and weight-loss determination
at the end of the test
Overview of current biodegradation methods
with a short description of their principles,
scopes and recommendations
Natural or artificial seawater

Parameters monitored

Similar to

n/a

BOD measurement

Measurement of biogas (CO2


& CH4) pressure or volume
and inorganic carbon
Specific analyses or
radiolabeled compounds

OECD 311

CO2-evolution. Volatile
substances can be analyzed

OECD 310

Oxygen consumption (BOD)

ISO 9408

CO2 evolution

ISO 9439

Part One is
similar to OECD
309, Part Two
describes a flow
rover model

Continuous CO2
measurement,, DMR, CMR,
GMR

n/a

TIC, DOC, CO2, BOD

OECD 306

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

14593

Involved microorganisms and key features

Trends

1062

Table 3 (continued)

Method

Polymer form (physiology,


morphology)

Inoculum and degradation criteria monitored

Gravimetry

Film or physical intact forms

Wide range of inocula from soil, water,


sewage or pure species from culture
collections

Respirometry

Film, powder, liquid and virtually all


forms and shapes

O2 consumed or CO2 produced under


aerobic conditions or CH4 produced under
methanogenic conditions

Surface hydrolysis

Films or sheets, pieces and others

EIS

Films or coatings resistant to water

Generally aerobic conditions, pure enzymes


are used. Hydrogen ions released are
monitored as incubation progresses
Test polymers should adhere on conductive
materials. Electrochemical conductance is
recorded

Radio labeling
GPC/SEC

All kinds of materials


Virtually most polymers soluble in
different solvents such as PEG, PVP,
Ecoflex, Ecovio
Ecoflex and others, PHB, Xanthan,
polysaccharide, Avicel. Requirement:
small molecules. MWD low
PEG, PVP, Requirement: small
molecules. MWD low
PEG, PVP, Ecoflex, Ecovio molecules
with higher molecular weight,
synthetic polymers
Particles adhered or dispersed to a
substrate
Thin and vacuum resistant, electron
transparent samples
Solid powder or liquid samples
Gold sputtered solid samples
PE-Wax, PHB, Xanthan,
polysaccharide, Avicel, solid or
liquid samples

GC, GC/MS

HPLC, LC/MS
MALDI-TOF
http://www.elsevier.com/locate/trac

AFM
TEM
NMR
SEM
FT-IR

Marine, soil, sewage, compost sediment etc.


Freshwater, Salt water, CO2 balance, DOC

Comments
Robust method, good for isolation of
degrading MOs. High reproducibility.
Disintegration cannot be differentiated from
biodegradation
Most adaptable to many materials.
Specialized instrumentation may be required.
When fermentation is the major mechanism
of DEG, the method gives underestimated
results
Prior information about DEG of sample by
MOs and particular enzymes is needed for a
target-specific test
Sample must be initial water impermeable for
signal transduction. DEG can proceed quickly
and, as it is registered, no further DEG process
can be distinguished
Samples need to be 14C labeled
Problems with environmental samples
because extraction may be required

Selected refs.
[17,7375]

[17,23,74,76,77]

[17,78,79]

[17,80]

[8183]
[84,29,85]

Soil leachate, CO2 balance; compost

Molecular weight can be limiting factor for


this type of analysis

[26,8691]

Freshwater, Salt water, CO2 balance, DOC

Molecular weight can be limiting factor for


this type of analysis
Parameters optimized, important for polymer
analysis

[29]

Surface analytical procedure

[9698]

Surface water, sea water, activated sludge

Surface analytical procedure

[96,99,100]

Bacterial degradation, surface area

Surface analytical procedure


Fingerprinting technique

[41,85,91]
[41,75,78,96]
[77,89,101]

Freshwater, Salt water, CO2 balance, DOC

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Table 4. Overview of analytical techniques for biodegradation of polymers

[29,85,88,9195]

Trends

1063

Trends

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Figure 1. GC-MS-chromatogram of water-soluble degradation intermediates of Ecoflex after 7 days incubation with Thermomonospora fusca
(10-fold magnification) [89]. The small diagram shows the initial scaling of the diagram. AB = Monoester of adipic acid (A) and butanediol
(B), BAB = Diester of A and B, ABAB = Oligomer of A and B.

and simulation tests) [23]. Since the working group on


biodegradability of plastics was created in 1993, rapid
advances have been made in this area. The approach
and the development of standardized methods and definitions have been published [24].
Recently, methods have been developed based on
enzymatic tests to shorten experimental times and to
obtain screening results quickly. These tests are based on
degradation of the polymer by lipase, which catalyses
hydrolytic cleavage of ester bonds. The free carboxylicacid groups generated can be determined by automatic
titration to give representative data on the degree of
degradation. This may be important, since biodegradation may in many cases be very time consuming and the
results are often only slightly outside the margins of the
test variance. Enzymatic tests may give results within a
few hours, days or weeks, using defined and controlled
test systems. These tests are not ideal in regard to natural environmental conditions, but their transferability
to those has been demonstrated and biodegradation has
been linked to degree of crystallinity [25].
To investigate the biodegradation potential of a polymeric compound, several analytical methods were used.
Table 4 gives an overview.
1064

http://www.elsevier.com/locate/trac

Low-MW metabolites of polymer degradation (e.g.,


alcohols, aldehydes, volatile fatty acids, mono-carboxylic
and di-carboxylic acids, methane, and carbon dioxide,
were assayed by gas chromatography with flame ionization detection (GC-FID) or GC with mass spectrometry
(GC-MS) [26]. Degradation intermediates (e.g., monomers, mono-esters and di-esters, and oligomers) of Ecoflex, a water-insoluble, aliphatic, aromatic copolyester,
which is synthesized from 1,4-butanediol, adipic acid
and terephthalic acid, were detected by GC-MS (Fig. 1) in
liquid cultures with a mycelium suspension of Thermomonospora fusca [89].
The residual polymer and its oligomers are often
analyzed by applying gel-permeation chromatography
(GPC) to measure a possible shift in MW distribution.
GPC can be further used to fractionate biodegradation
samples prior to analysis by matrix-assisted laser
desorption/ionization time-of-flight MS (MALDI-TOF-MS)
to lower the polydispersity of the polymer sample [27].
The fate of individual homologues of low-MW polymers was investigated by high-performance liquid
chromatography (HPLC) with subsequent ultraviolet
(UV) and fluorescent detection after derivatization [28].
Substance-specific analyses based on the parent-polymer

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Trends

Figure 2. MALDI-TOF-MS spectra of PVP 2.5 kDa in a biodegradation experiment with river water from the Rhine: (a) 0 h, (b) 5 h, (c) 8 days,
(d) 16 days after start [27].

compound and possible metabolites were performed


using LC-MS and MALDI-TOF-MS, enabling the study of
the degradation pathway at the molecular level [29].
MALDI-TOF-MS was used in a biodegradation experiment of PVP 2500 in a fixed-bed bioreactor with water
from the river Rhine to demonstrate the recalcitrance of
this water-soluble polymer [26]. Neither a shift in MW
distribution of PVP nor a rising distribution of a potential
metabolite was found (Fig. 2). Slow diminution of
higher-MW oligomers during the course of degradation
(Fig. 2ad) was explained by a decrease in concentration
due to adsorption on fixed-bed material that involved a
decrease in signal intensity.
Fig. 3 shows the formation of short-chain PEG
homologues during aerobic biodegradation of polydispersed PEG 2000 in freshwater media with inoculum of
wastewater-treatment plants (WWTPs) measured by
MALDI-TOF-MS [28]. From day 1 (Fig. 3a) to day 6
(Fig. 3b), a clear shift in the MW distribution of the
polymer and a loss of long-chain PEGs can be seen,
indicating rapid biodegradation of the PEG 2000.
2.1. Biotic degradation
It is known today that the chemical structure determines
to a great extent whether or not a substance is biode-

gradable. There are possible ways to promote biodegradation, e.g.:


I) changes in chemical molecular structure;
II) the development of MOs that are capable of using a
certain carbon source as substrate; or,
III) making substances more conductive to attack by
microbial populations, as synthetic polymers do
not occur naturally in the environment so MOs
may not have the necessary tools for many synthetic structures.
Many factors contribute to recalcitrance. MOs may
lack the necessary genetic information, which they may
acquire by plasmid transfer or de novo enzyme synthesis.
Also, the compounds may be too large to be assimilated
into the cell or no membrane transport system may exist.
Compounds can be insoluble and MOs may lack the
proper nutrients. Recalcitrant compounds can be oxidized when a readily biodegradable organic carbon (OC)
source is available, but, in its absence, the recalcitrant
compound is not degraded (e.g., alkane or lignin degradation).
Polymers are known to show mostly complex biodegradation behavior. The initial rate of degradation of a
polymer often follows a Freundlich or modified
Langmuir isotherm rather than Michaelis-Menten

http://www.elsevier.com/locate/trac

1065

Trends

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Figure 3. MALDI-TOF-MS spectra of polydispersed PEG 2000 during aerobic biodegradation in freshwater media, with matrix DCTB. (a) Sample
from day 1, and (b) sample from day 6. Numbers indicate repeating units; black and gray lines show polymeric distribution before and during
degradation.

1066

http://www.elsevier.com/locate/trac

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

kinetics [30]. The search for suitable MOs necessitates


experiments in different microcosms. Acceleration might
eventually be possible to a certain degree when changing
parameters [e.g., temperature, salinity, pH, UV light,
special enzymes/organisms or pre-adaptation of organisms (semi-continuous activated sludge (SCAS) test)] but
would lack environmental relevance.
Polymer materials may be used as potential sources of
carbon and energy for heterotrophic MOs, including
bacteria and fungi, and additives used in polymer manufacture may serve as good nutrients for degrading
organisms at ambient temperature. Microbial assimilation of polymers is limited by the cell-membrane structure of the degraders. Enzymes of degraders (e.g., lipase
and polymerase) may help in decreasing the size and the
MW of macromolecules to dimers, monomers and oligomers before uptake by a much wider range of MOs
[17,31] (Fig. 4).
In most cases of biodegradation, the first step is an
enzymatically-catalyzed hydrolysis of ester, amide or
urethane bonds. This primary depolymerization is a

Trends

surface-erosion process. It leads at least to water-soluble


intermediates, which can be assimilated by MOs and
further metabolized.
It was observed that some sort of priming effect can
occur if an easily degradable compound is added. Then,
more than the predicted amount of CO2 is evolved. The
effect decreased when the amount of substance added
was further increased.
It has also been reported that the addition of such
substances to the biodegradation test can significantly
increase the number of MOs. Addition of 5 wt% glucose
resulted in 10100 times more MOs but the increase did
not influence the final biodegradability of the polymers
tested [32].
A similar study investigated the influence of sludge
inocula in ISO 14851 Tests. MOs were taken for municipal and industrial WWTPs and the results showed that
municipal sludge degraded ST-based polymers and
polycaprolactone (PCL) to similar degrees. In some cases,
ST-based materials were degraded more easily. Degradation in industrial sludge was in both cases better, but

Figure 4. Pathways of PEG degradation.

http://www.elsevier.com/locate/trac

1067

Trends

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

degradation of cellulose (used as reference) was lower in


industrial sludge than in municipal sludge. Acclimatization of MOs to the substance did not lead to the results
expected [33].
2.2. Abiotic degradation
Synthetic hetero-chain polymers are often susceptible to
hydrolytic cleavage that can be followed by microbial
bioassimilation (e.g., polyesters), while carbon-chain
polymers undergo degradation by peroxidation and formation of low-MW carboxylic acids further used as
carbon sources by bacteria and fungi. This process can
be controlled by controlling the peroxidation step. It is
possible to control photolysis, photo-biodegradation,
hydro-biodegradation, antioxidant-controlled photo-biodegradation and thermal biodegradation using chemical
structure and additives. The change in mechanical
properties is due to chemical modification and leads to
smaller MW products that are susceptible to MOs.
Products with these special properties are very important
in medicine, agricultural technology, water and fertilizer
conservation, controlled release, forestry and packaging.
Abiotic degradation can be very easily determined,
though the experiments sometimes take a long time [34]
(e.g., hydrolysis and photo degradation are easy to
determine via weight loss [35,36]). Other physical
parameters (e.g., elongation, breaking loads, moisture
absorption, shear characterization and change in diameter) provide standard test methods in polymer research
so they are easily applicable [3739]. In addition, tracer
studies (with isotope labeling) and development of
accelerated test methods [changes in parameters (e.g.,
salinity, temperature, pressure, pH and UV light)] would
provide new challenging areas of research [17].
It is generally known that polymers with ester bonds
may be degraded by hydrolysis because hydrolytic processes under natural conditions can cleave ester bonds.
Different studies on hydrolytic degradation have been
published for hydrogels from:
 microbial poly(amino acid)s [40];
 poly(lactic acid)/poly(ethylene oxide)/poly(lactic acid)
(PLA/PEO/PLA) triblock copolymers with short PLA
blocks [41]; and,
 monofilaments of poly(butylene terephthalate)poly(tetramethylene oxide) (PBT-PTMO) along with
poly(ethylene terephthalate) (PET), poly(hexamethyleneadipamide) and PP in salt water and distilled
water and glass-epoxy composites [42].
2.3. Photo degradation/UV-oxidation
There have been studies on photo degradation or photodegraded polymers. Acceleration is generally possible
with varying light intensity but the photo-degraded
polymers are just smaller particles that are still intact
and able to act as stressors in the environment. In
general, physical parameters are measured (e.g., oxygen
1068

http://www.elsevier.com/locate/trac

and water uptake). There are only a few studies available


for marine environments [43].
In order to investigate biological degradation or oxidation of polymers after photo degradation, experimental
methods were developed using degraded poly(styrene)
(PS), PE and PP [43] and also for a photo-degraded
poly(styrene)-(PS)-vinyl-ketone copolymer [44]. UVFenton methods were tested with soil microbial communities and sewage sludge and carried out in respirometers [43,44]. However, the procedure may also be
easily adapted to aquatic systems.
The contaminated solution is passed through a
chamber, where it is exposed to intense UV radiation.
The UV-oxidation technology is a commercially available
as a water-treatment technology that has been used for
more than 10 years, mainly for contaminated groundwater. An investigation with representative samples of
commercial photodegradable PEs and an ethylenecarbon monoxide (E/CO) copolymer that were photo
degraded most rapidly but biodegraded most slowly. By
contrast, antioxidant iron dithiocarbamate photodegradable PE and a ST-filled iron-catalyzed PE showed a
higher production of carboxylic acids during photo oxidation than E/CO, therefore resulting in rapid microbial
growth.
Abiotic iron-catalyzed photo-oxidation or thermal
oxidation is the rate-limiting step in the process of bioassimilation. Abiotic degradation must therefore precede
biotic degradation. Also, the MW seems not to be the
limiting factor, since products of up to 40,000 g/mol
were readily degraded. Much more important is the
nature of the photochemical or thermochemical modification [45].
2.4. The Fenton reaction
H.J.H. Fenton discovered in 1894 that several metals
have special oxygen-transfer properties that improve the
use of hydrogen peroxide. Some metals have a strong
catalytic power to generate highly reactive hydroxyl
radicals (OH). The reaction is used to treat a large
variety of water pollutants (e.g., phenols, formaldehyde,
BTEX, pesticides and rubber chemicals). The Fenton
reaction is currently one of the most powerful oxidizing
reactions available [46]. In natural waters, the Fenton
reaction may have been underestimated, since it may
take more than a few hours for it to start [47].
The oxidative biodegradability of poly(ether urethane)
(PEU) and poly(carbonate urethane) (PCU) was investigated in an accelerated in vitro system. Changes were
consistent with those seen in long-term in vivo studies.
The accelerated method was found suitable to simulate
real environmental oxidation within a shorter time [48].
The Fenton reaction was also reported for water-soluble polymers [e.g., PEG, PVP and poly(acrylamide)
(PAM)] in wastewater treatment [49] and for cellulose
and hemicellulose [50].

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

2.5. Ultrasound degradation


The degradation of polymers using ultrasound was
investigated for PEGs. It has been demonstrated that
cleavage of a polymer chain occurs preferentially at
weak bonds incorporated in the chain and that
mechanically-induced cleavage can be localized almost
exclusively to a single weak site [51].
2.6. Modeling
Today, modeling of bioprocesses (e.g., biodegradation) or
the estimation of environmental concentrations (PECs)
plays an important role in evaluating results as well as
predicting or estimating an impact in environmental
assessment. So far, no quantitative structure-activity
relationship (QSAR) model exists. One mechanistic
model applicable for polymer biodegradation has been
reported [52]. Generally, these models were applied to
small molecules (active substances and others) from
pharmaceuticals, plant science, and PCPs.
Some recently developed models give relevant insights
on how to apply mathematical methods to biodegradation experiments [5355]. In particular, CATABOLbased models are discussed to assess biodegradation in
standardized tests using QSAR methodologies [5658].
But there are also special models used for PEC assessment [59] or specific processes (e.g., Fenton oxidation
[60] or particle-size distribution [61].
2.7. Biodegradation in the environment
The European Packaging Directive (EPD) 94/62/EC (31
December 1994) aims to harmonize national measures
concerning the management of packaging and packaging wastes in order to prevent or to reduce any impact
on the environment. First priority is given to reducing
production of packaging waste, then to re-use, recycling
and recovery, and reducing the final disposal of such
waste [62]. The EPD specifies composting as a form of
recycling for packaging wastes. CEN TC261 SC4 (European Normalization Committee) was given the mandate
to regulate and to develop norms and criteria in testing
materials for compostability. Regarding the acceptance
for composting and organic recovery, basically three
criteria are required in the following order:
(1) biodegradation of the specific material;
(2) disintegration; and,
(3) no influence on the quality of the compost.
A standard composting test based on the determination of biodegradability under dry, aerobic conditions via
measurement of evolved CO2 is described by ISO 14855
[14,15]. Relevant organic constituents with more than
1% of the dry weight have to be investigated for biodegradation. In the appropriate laboratory-scale test, not
only bacteria, but also fungi, moulds and actinomycetes
are available as active MOs [62]. The test was run in a
ring, standardized and described [63,64]. To be accepted
in composting plants, biodegradability of plastics alone is

Trends

not sufficient. Disintegration (i.e. physically falling apart


into fine, visually indistinguishable fragments at the end
of a typical composting cycle) is also an important factor.
It is measured in a pilot or full-scale composting test by
sieving the compost at the end of exposure and precise
sorting analysis.
Until the 1980s, composting was not used very much
as a solid-waste treatment (4% for Europe and less in the
USA). Nowadays, composting with control over moisture, pH, aeration, retention time, temperature and other
factors is mostly done in-vessel [65].
In compost and soil environments, the most
important factor concerning biodegradability is aerobic
biodegradation by bacteria and fungi. For standardized
composting tests, the inoculum is generally finished
compost from a composting plant [18] but degradation
environment plays an important role. Several degradable polymers were therefore tested in laboratory-scale
composting tests, exposure in thermal-hydrolytic
environment using water at 60C and exposure in
thermal-oxidative dry oven environment at 60C. It
was found that, next to chemical and biological
transformation, physical effects (e.g., restructuring or
reorganization of the macromolecules) may happen at
environmental temperatures typical of compost. It was
observed that all PE-based films behaved similarly but
different to all other polymers tested. PE-based polymers should degrade in thermal-oxidative environment
but, when applied in laboratory-scale compost tests
with sufficient oxygen available and at high temperature, no biodegradation was observed within the test
duration [66]. For determining whether mineralization
has occurred, summary parameters DOC, BOD or CO2
evolution were measured. In DOC measurement, the
initial concentration is compared with the final concentration. In measurements of BOD or CO2 evolution,
the values acquired throughout the test are accumulated and compared with the respective theoretical
data calculated from the chemical formula. DOC
measurement is not applicable if water-insoluble
polymers are to be tested; only BOD and CO2 evolution
can then be measured. For investigation of primary
degradation, additional analytical tests (e.g., GPC and/
or MALDI-TOF-MS) are necessary to check plausibility
[18].
In terrestrial test system ISO 14855 (also European
Standard Draft WI 261 085), both inoculum and test
conditions provide good conditions for polymer degradation. The purposes of the test are to demonstrate
complete biodegradation and to differentiate the latter
from abiotic deterioration. In this test, the substance
must be degraded to the maximum degree achievable
(duration 180 d), which cannot be compared to conventional treatment of biowaste by composting procedures, because the European Standard on requirements
(European Standard Draft WI 261 236 1998) clearly
http://www.elsevier.com/locate/trac

1069

Trends

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

states that degradation of test materials may be completed when using compost in the soil. The test period of
180 days opposes the mean residence time of the test
materials in biowaste-treatment plants (BWTPs) of
approximately 42 days (6 weeks) [18].
Tests on many different polymers [e.g., PE, poly(vinyl
chloride) (PVC), PP and others] were performed with
different methods in soil. Traditional methods were
compared to new methods [67]. The cumulative CO2evolution background from the soil only was typically
determined at 4 lmol during the 160-d test period. The
PE samples evolved in the range 500860 lmol CO2
with a lack of correlation between photo exposure and
CO2 evolution. The PE sample with 7.7% ST evolved
around 2 mmol CO2 and the PVC sample around
2.5 mmol CO2. It is possible that the CO2 evolved during
the 160 d might originate from the additives, colorants
or other substances, rather than the polymer resin. The
study shows only a limited correlation of techniques
used. CO2 evolution does not necessarily correlate with
MW decrease and tensile strength or elongation measurements. This especially complicates the prediction of
biodegradation of complex polymers. In particular,
residual weight determination seems to be the most
unsuitable method for determining degradation, especially for ST-containing polymers [67].
Different copolyesters with aromatic constituents were
used in biodegradation tests with compost eluate and in
soil burial tests. The polymers were synthesized with
dimethyl terephthalate and different diols (1,2-ethanediol, 1,3-propanediol, 1,4-butanediol) in certain ratios.
Results from the Sturm test performed with a compost
eluate as inoculum source showed that the ester-type
oligomers and polymers were biodegradable. The rate of
biodegradation directly depended on the chain length of
the aliphatic constituents and the degree of polymerization [68].
The microbial susceptibility of high molar mass
copolyesters of terephthalic acid and different co-monomers has been investigated in parallel with their
dependence on their composition. It has been shown
that the optimal copolymer composition of terephthalic
acid with aliphatic co-monomers with regard to biodegradation and melting point is in the range 40
50 mol% of terephthalic acid. These copolymers have
melting points above 90C when the fraction of terephthalic acid is higher than 40 mol% of the acid content.
Even polymers with melting points above 140C show
significant weight loss in composting tests within 12
weeks. At higher temperatures, biological attack is
advantageously accompanied and supported by chemical hydrolysis enhancing biodegradation of the copolyesters. Degradation at room temperature is caused by
biological attack only. Significant hydrolysis was not
observed under sterile test conditions or even in soil
burial tests [69].
1070

http://www.elsevier.com/locate/trac

The effect of inoculum treatment on biodegradation


was investigated with compost stored for different
periods and at different temperatures. Results give
independent biodegradation results for PCL and
poly(butylene succinate) (PBS) polymers but the same
results for cellulose. It was also shown that the presence
of different MOs depends on the storage temperature. It
is therefore suggested that easily biodegradable reference substances may not always be suitable as a basis
for validity criteria but more precise standards on the
preparation of the inoculum could guarantee more
reliable results [70].
Biodegradation in aqueous environments has not been
investigated much. It was reported in a recent study that
PEGs are degraded in the marine environment and
WWTPs [29], but most other studies report results on
polyhydroxy alkanoates only. Pseudoalteromonas sp.
NRRL B-30083 was isolated as predominant
poly(hydroxyabutyrate-co-hydroxyvalerate)
(PHBV)degrading bacterium from a tropical marine environment [71]. Determination of PE and PE-ST composite
degradation in marine environments and strawline of a
marsh was simulated and physical parameters (e.g.,
weight loss, tensile properties, ST loss and carbonyl
content) were analyzed. Low degradation was observed
in marine water for both control PE and PE-ST, while
deterioration could be observed in strawline of a marsh
[35].

3. Conclusions
The main focus in biodegradation research has been on
polyesters and on soil or compost biodegradation because this is a major route of entry for packaging
material especially developed for this purpose. Biodegradation in aqueous media has mostly been neglected.
There are a few guidelines and methods available that
can be adapted for wastewater, freshwater and seawater
environments and their sediment systems, but almost no
data has been published so far. As has been demonstrated [29,72], aqueous environments may be very
important because polymers used (e.g., in PCPs) may
pass through WWTPs or enter the aqueous environment
directly.
Research on the biodegradation pathways and the
conditions under which biodegradation may occur when
comparing different environmental compartments may
improve the development of polymers for these applications. It is also imperative to ensure real biodegradability
and incorporation in the geochemical life cycle rather
than to provide materials that breakdown only to small
particles or their oligomers. We also need to remember
that many products are required not to be biodegradable
and, for them, recycling is the best way to reuse the
material.

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

References
[1] N. Lucas, C. Bienaime, C. Belloy, M. Queneudec, F. Silvestre, J.E.
Nava-Saucedo, Chemosphere 73 (2008) 429.
[2] P.P. Klemchuk, Polym. Degrad. Stab. 27 (1990) 183.
[3] R. Chandra, R. Rustgi, Prog. Polym. Sci. 23 (1998) 1273.
[4] I. Kyrikou, D. Briassoulis, J. Polym. Environ. 15 (2007) 125.
[5] R.N. Tharanathan, Trends Food Sci. Technol. 14 (2003) 71.
[6] R.A. Gross, B. Kalra, Science (Washington, DC) 297 (2002) 803.
[7] S. Mecking, Angew. Chem. Int. Ed. Engl. 43 (2004) 1078.
[8] G. Scott, Degradable Polymers: Principles and Applications,
Kluwer Academic Publishers, Dordrecht, The Netherlands, 2003.
[9] G. Swift, Acc. Chem. Res. 26 (1993) 105.
[10] A. Gandini, Macromolecules 41 (2008) 9491.
[11] J.P. Eubeler, M. Bernhard, S. Zok, T.P. Knepper, Trends Anal.
Chem. (2009) (submitted for publication).
[12] J.W. Garthe, P.D. Kowal, The chemical composition of degradable plastics (http://www.age.psu.edu/extension/factsheets/c/
C17.pdf) (2001).
. Bellon-Maurel, F. SilvestreAdvances
[13] A. Calmon-Decriaud, V.A
in Polymer Science, Vol. 135, Springer, Berlin, Germany, 1998
pp. 207226.
[14] Organization for Economic Co-operation and Development
(OECD), OECD Guidelines OECD 301 A-F, OECD 302 A-C, OECD
303, OECD 304, 306 & 309, OECD, Paris, France.
[15] International Organization for Standardization (ISO), ISO Methods 7827 (1994), 9439 (1999), 9408 (1999), 9888 (1999),
10707 (1994), 10708 (1997), 14592 (2002), 14593 (1999),
15462 (1997), 16221 (2001), 14855-1 (2005), 14851 (1999),
14852 (1999), 14855 (1999), 11734 (1995), 10634 (1995),
11733 (2004), 9887(1992), ISO, Geneva, Switzerland.
[16] ASTM International, ASTM Test Methods for Testing of Chemicals: D5209-92; D5210-92; D5247-92; D5271-02; D6003-96;
D6340-98; D6691-01; D6692-01; D7081-05, ASTM International, West Conshohocken, PA, USA.
[17] J.G. Gu, J.D. Gu, J. Polym. Environ. 13 (2005) 65.
[18] U. Pagga, Appl. Microbiol. Biotechnol. 51 (1999) 125.
[19] M. van der Zee, L. Sijtsma, G.B. Tan, H. Tournois, D. De Wit,
Chemosphere 28 (1994) 1757.
[20] M. Itavaara, M. Vikman, Chemosphere 31 (1995) 4359.
[21] U.J. Strotmann, H. Schwarz, U. Pagga, Chemosphere 30 (1995)
525.
[22] U. Pagga, A. Schafer, R.J. Muller, M. Pantke, Chemosphere 42
(2001) 319.
[23] P. Reuschenbach, U. Pagga, U. Strotmann, Water Res. 37
(2003) 1571.
[24] H. Sawada, Polym. Degrad. Stab. 59 (1998) 365.
[25] K. Welzel, R.-J. Mueller, W.-D. Deckwer, Chem. Ing. Tech. 74
(2002) 1496.
[26] P. Drimal, J. Hmcirik, J. Hoffmann, J. Polym. Environ. 14 (2006)
309.
[27] S. Trimpin, P. Eichhorn, H.J. Rader, K. Mullen, T.P. Knepper, J.
Chromatogr., A 938 (2001) 67.
[28] A. Zgola-Grzeskowiak, T. Grzeskowiak, J. Zembrzuska, Z.
Lukaszewski, Chemosphere 64 (2006) 803.
[29] M. Bernhard, J.P. Eubeler, S. Zok, T.P. Knepper, Water Res. 42
(2008) 4791.
[30] M.D. Faber, Enzyme Microb. Technol. 1 (1979) 226.
[31] J.D. Gu, Int. Biodeterior. Biodegrad. 52 (2003) 69.
[32] J.C. Jang, P.K. Shin, J.S. Yoon, I.M. Lee, H.S. Lee, M.N. Kim,
Polym. Degrad. Stab. 76 (2002) 155.
[33] V. Mezzanotte, R. Bertani, F.D. Innocenti, M. Tosin, Polym.
Degrad. Stab. 87 (2005) 51.
[34] T.F.M. Ojeda, E. Dalmolin, M.M.C. Forte, R.J.S. Jacques, F.M.
Bento, F.A.O. Camargo, Polym. Degrad. Stab. 94 (2009) 965.
[35] V.T. Breslin, B. Li, J. Appl. Polym. Sci. 48 (1993) 2036.
[36] G. Scott, Polym. Degrad. Stab. 29 (1990) 135.

Trends
[37] P. Agamuthu, P.N. Faizura, Waste Manage. Res. 23 (2005) 95.
[38] A.L. Andrady, J.E. Pegram, Y. Song, J. Polym. Environ. 1 (1993)
117.
[39] L. Jiang, M.P. Wolcott, J. Zhang, Biomacromolecules 7 (2006)
199.
[40] M. Kunioka, H.J. Choi, Polym. Degrad. Stab. 59 (1998) 33.
[41] S.M. Li, I. Rashkov, J.L. Espartero, N. Manolova, M. Vert,
Macromolecules 29 (1996) 57.
[42] P. Davies, F. Pomies, L.A. Carlsson, Appl. Compos. Mater. 3
(1996) 71.
[43] P.H. Jones, D. Prasad, M. Heskins, M.H. Morgan, J.E. Guillet,
Environ. Sci. Technol. 8 (1974) 919.
[44] J.E. Guillet, T.W. Regulski, T.B. McAneney, Environ. Sci.
Technol. 8 (1974) 923.
[45] R. Arnaud, P. Dabin, J. Lemaire, S. Al-Malaika, S. Chohan, M.
Coker, G. Scott, A. Fauve, A. Maaroufi, Polym. Degrad. Stab. 46
(1994) 211.
[46] J. Ma, W. Ma, W. Song, C. Chen, Y. Tang, J. Zhao, Y. Huang, Y.
Xu, L. Zang, Environ. Sci. Technol. 40 (2006) 618.
[47] B.A. Southworth, B.M. Voelker, Environ. Sci. Technol. 37 (2003)
1130.
[48] E.M. Christenson, J.M. Anderson, A. Hiltner, J. Biomed. Mater.
Res., A 70 (2004) 245.
[49] J.A. Giroto, A.C.S.C. Teixeira, C.A.O. Nascimento, R. Guardani,
Chem. Eng. Process. 47 (2008) 2361.
[50] V. Arantes, A.M. Ferreira Milagres, J. Chem. Technol. Biotechnol.
81 (2005) 413.
[51] K.L. Berkowski, S.L. Potisek, C.R. Hickenboth, J.S. Moore,
Macromolecules 38 (2005) 8975.
[52] R. Mahadevan, L. Smith, J. Polym. Environ. 15 (2007) 75.
[53] S.H. Gordon, S.H. Imam, R.L. Shogren, N.S. Govind, R.V. Greene,
J. Appl. Polym. Sci. 76 (2000) 1767.
[54] A.T. Metters, K.S. Anseth, C.N. Bowman, J. Phys. Chem. B 105
(2001) 8069.
[55] A.T. Metters, C.N. Bowman, K.S. Anseth, J. Phys. Chem. B 104
(2000) 7043.
[56] R.M. McDowell, J.S. Jaworska, SAR QSAR Environ. Res. 13
(2002) 111.
[57] S. Dimitrov, V. Kaminski, J.D. Walker, W. Dwindle, R. Purdy, M.
Lewis, O. Me Kenyan, SAR QSAR Environ. Res. 15 (2004) 69.
[58] Y. Sakuratani, J. Yamada, K. Kasai, Y. Noguchi, T. Nishihara,
SAR QSAR Environ. Res. 16 (2005) 403.
[59] V.D.J. Keller, H.G. Rees, K.K. Fox, M.J. Whelan, Environ. Pollut.
148 (2007) 334.
[60] C.K. Duesterberg, T.D. Waite, Environ. Sci. Technol. 40 (2006)
4189.
[61] X.Y. Li, J.J. Zhang, J.H.W. Lee, Water Res. 38 (2004) 1305.
[62] U. Pagga, Polym. Degrad. Stab. 59 (1998) 371.
[63] U. Pagga, D.B. Beimborn, J. Boelens, B. De Wilde, Chemosphere
31 (1995) 4475.
[64] M. Tosin, F. Degli-Innocenti, C. Bastioli, J. Polym. Environ. 4
(1996) 55.
[65] B. De Wilde, J. Boelens, Polym. Degrad. Stab. 59 (1998) 7.
[66] M. Day, K. Shaw, D. Cooney, J. Watts, B. Harrigan, J. Polym.
Environ. 5 (1997) 137.
[67] A.V. Yabannavar, R. Bartha, Appl. Environ. Microbiol. 60
(1994) 3608.
[68] U. Witt, R.J. Mueller, W.D. Deckwer, J. Polym. Environ. 4 (1996)
9.
[69] U. Witt, R.J. Muller, W.D. Deckwer, J. Polym. Environ. 3 (1995)
215.
[70] H.S. Yang, J.S. Yoon, M.N. Kim, Polym. Degrad. Stab. 84 (2004)
411.
[71] T.D. Leathers, N.S. Govind, R.V. Greene, J. Polym. Environ. 8
(2000) 119.
[72] J. Schwarzbauer, Entwicklung einer analytischen Methode zur
Erfassung von poly(vinyl pyrrolidon) PVP in Umweltproben auf

http://www.elsevier.com/locate/trac

1071

Trends

[73]

[74]
[75]
[76]
[77]
[78]
[79]

[80]
[81]

[82]

Trends in Analytical Chemistry, Vol. 28, No. 9, 2009

Basis pyrolytischer Verfahren, Paper presented at SETAC GLB &


German Society of Chemists, Environmental Chemistry and
Ecotoxicology Group - Annual Conference, Frankfurt am Main,
Germany, 2006.
A. Hoshino, M. Tsuji, M. Momochi, A. Mizutani, H. Sawada, S.
Kohnami, H. Nakagomi, M. Ito, H. Saida, M. Ohnishi, M. Hirata,
M. Kunioka, M. Funabash, S. Uematsu, J. Polym. Environ. 15
(2007) 275.
G. Kale, R. Auras, S.P. Singh, R. Narayan, Polym. Test. 26
(2007) 1049.
R.K. Soni, S. Soam, K. Dutt, Polym. Degrad. Stab. 94 (2009)
432.
R. Solaro, A. Corti, E. Chiellini, J. Environ. Polym. Degrad. 6
(1998) 203.
Y. Orhan, H. Buyukgungor, Int. Biodeterior. Biodegrad. 45
(2000) 49.
H.P. Molitoris, S.T. Moss, G.J.M. de Koning, D. Jendrossek, Appl.
Microbiol. Biotechnol. 46 (1996) 570.
T. Nakajima-Kambe, Y. Shigeno-Akutsu, N. Nomura, F. Onuma,
T. Nakahara, Appl. Microbiol. Biotechnol. 51 (1999)
134.
J.D. Gu, D.B. Mitton, T.E. Ford, R. Mitchell, Biodegradation 9
(1998) 39.
ASTM International, ASTM D6340-98; Standard test method for
determining aerobic biodegradation of radiolabeled plastic materials in an aqueous or compost environment, ASTM International, West Conshohocken, PA, USA, 2007.
ASTM International, ASTM D6692-01; Standard test method for
determining the biodegradability of radiolabeled polymeric plastic materials in seawater, ASTM International, West Conshohocken, PA, USA, 2001.

1072

http://www.elsevier.com/locate/trac

[83] R. Benner, A.E. Maccubbin, R.E. Hodson, Appl. Environ. Microbiol. 47 (1984) 381.
[84] K. Se, T. Sakakibara, E. Ogawa, Polymer 43 (2002) 5447.
[85] S. Lee, M.A. Winnik, R.M. Whittal, L. Li, Macromolecules 29
(1996) 3060.
[86] D. Fabbri, D. Tartari, C. Trombini, Anal. Chim. Acta 413 (2000)
3.
[87] D. Fabbri, J. Anal. Appl. Pyrolysis 5859 (2001) 361.
[88] W.W. Zhu, J.R. Allaway, J. Chromatogr., A 1055 (2004) 191.
[89] P. Drimal, J. Hoffmann, M. Druzbik, Polym. Test. 26 (2007) 729.
[90] U. Witt, T. Einig, M. Yamamoto, I. Kleeberg, W.-D. Deckwer, R.-J.
Muller, Chemosphere 44 (2001) 289.
[91] G. Gallet, S. Carroccio, P. Rizzarelli, S. Karlsson, Polymer 43
(2002) 1081.
[92] X.M. Liu, E.P. Maziarz, D.J. Heiler, G.L. Grobe, J. Am. Soc. Mass
Spectrom. 14 (2003) 195.
[93] P. Terrier, W. Buchmann, G. Cheguillaume, B. Desmazieres, J.
Tortajada, Anal. Chem. 77 (2005) 3292.
[94] G. Gallet, B. Erlandsson, A.C. Albertsson, S. Karlsson, Polym.
Degrad. Stab. 77 (2002) 55.
[95] S.J. Wetzel, C.M. Guttman, K.M. Flynn, J.J. Filliben, J. Am. Soc.
Mass Spectrom. 17 (2006) 246.
[96] T. Iwata, Y. Doi, Macromolecules 31 (1998) 2461.
[97] Y. Hu, L. Zhang, Y. Cao, H. Ge, X. Jiang, C. Yang, Biomacromolecules 5 (2004) 1756.
[98] K. Numata, Y. Kikkawa, T. Tsuge, T. Iwata, Y. Doi, H. Abe,
Biomacromolecules 6 (2005) 2008.
[99] M. Heldal, S. Norland, K.M. Fagerbakke, F. Thingstad, G.
Bratbak, Mar. Pollut. Bull. 33 (1996) 3.
[100] E. Ikada, J. Polym. Environ. 7 (1999) 197.
[101] A.S. Luyt, R. Bruell, Polym. Bull. 52 (2004) 177.

Potrebbero piacerti anche