production and characterization of polyhydroxyalkanoates (phas) from burkholderia cepacia atcc 17759...

Upload: iftikhar-ali

Post on 05-Apr-2018

230 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    1/154

    PRODUCTION AND CHARACTERIZATION OF POLYHYDROXYALKANOATES (PHAS)

    FROMBURKHOLDERIA CEPACIA ATCC 17759 GROWN ON RENEWABLE

    FEEDSTOCKS

    by

    Chengjun Zhu

    A dissertation

    submitted in partial fulfillment

    of the requirements for theDoctor of Philosophy Degree

    State University of New YorkCollege of Environmental Science and Forestry

    Syracuse, New York

    August 2011

    Approved: Department of Environmental and Forest Biology

    James P. Nakas, Major Professor Emanuel J. Carter, Jr., Chair

    Examining Committee

    Donald J. Leopold, Department Chair S. Scott Shannon, Dean

    The Graduate School

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    2/154

    All rights reserved

    INFORMATION TO ALL USERSThe quality of this reproduction is dependent on the quality of the copy submitted.

    In the unlikely event that the author did not send a complete manuscriptand there are missing pages, these will be noted. Also, if material had to be removed,

    a note will indicate the deletion.

    All rights reserved. This edition of the work is protected againstunauthorized copying under Title 17, United States Code.

    ProQuest LLC.789 East Eisenhower Parkway

    P.O. Box 1346

    Ann Arbor, MI 48106 - 1346

    UMI 3494520

    Copyright 2012 by ProQuest LLC.

    UMI Number: 3494520

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    3/154

    2011

    CopyrightC. J. Zhu

    All rights reserved

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    4/154

    iii

    Acknowledgements

    I would like to thank my major professor, Dr. James P. Nakas, for his advice,

    support and knowledge throughout my Ph.D study and with the manuscripts that we

    have submitted and will submit for publication. I also would like to thank Dr. Nomura for

    his guidance and discussion of our project and his kindness and great help for use and

    maintenance of his equipment. Likewise, I am very grateful to Dr. Arthur Stipanovic for

    his expert advice and assistance regarding physical-chemical property tests. I am also

    grateful to Dr. Patrick Mather, Ms. Erika D. Rodriguez and Mrs. Xinzhu Gu for their

    advice and support regarding mechanical property tests of PHAs. I wish to thank Mr.

    David Kiemle for assistance with NMR analysis, Mr. Daniel Nicholson and Dr. Kun Cheng

    for assistance with the physical characterization of PHAs. I want to thank Mr. David Sgroi,

    Ms. Laura Mateya, Ms. Giselle Kathryn Schlegel, Mr Matthew Michael Cleere and Mr.

    Sam Kogon for PHA isolation and purification from pilot-plant scale fermentations. I am

    grateful to Mr. Joseph K. Gredder, Ms. Anna Elyse Karczewski, etc. for assistance with

    the lab-scale research for PHA production. Lastly, I would like to thank my wife Qin for

    her support of my research and my son Felix who has brought great joy into my life.

    This research was supported by a grant from the New York State Energy Research

    and Development Authority (NYSERDA), the Welch Allyn Corp. (Skaneateles, NY), the

    Blue Highway LLC. (Syracuse, NY) and Tessy Plastics Corp. (Elbridge, NY).

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    5/154

    iv

    Table of Contents

    List of Tables..viii

    List of Figures..ix

    Thesis Abstract..xii

    1. Introduction..1

    1.1 Microbial polyhydroxyalkanoates..1

    1.2 Physicochemical and mechanical properties of PHAs.3

    1.3 History and development of PHAs.8

    1.4 Metabolic pathway for biosynthesis of PHAs from various carbon sources.13

    1.5 Biodegradation and thermal degradation of PHAs.17

    1.5.1 Biodegradation of PHAs.18

    1.5.2 Thermal degradation of PHAs21

    1.6 Considerations and rationale of renewable feedstocks and downstream processing

    1.6.1 Renewable feedstocks for PHA production.23

    1.6.2 Production, economics and renewability of glycerol and levulinic acid.26

    1.6.3 Downstream processing for PHA isolation29

    2. Materials and Methods..33

    2.1Microorganism and fermentation conditions.33

    2.1.1 PHB production in shake flasks and fermentors.33

    2.1.2 PHB-co-HV production in shake flasks and fermentors.342.1.3 P3HB -co-4HB production in shake flasks and fermentors..36

    2.2Concentration determination of glycerol, xylose, lactose and levulinic acid..37

    2.3Isolation and purification of PHAs from biomass.38

    2.4Sample preparation with nucleating agents40

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    6/154

    v

    2.5GC analysis for PHAs.41

    2.6Physicochemical property and mechanical property tests of PHAs42

    2.6.1 Molecular mass determination42

    2.6.2 Thermal analysis42

    2.6.3 Tensile test.43

    2.6.4 Nuclear magnetic resonance46

    3. Results47

    3.1Renewable carbon sources for bacterial growth and PHA production by B. cepacia.47

    3.2Biodiesel-derived glycerol as a carbon source for PHA production.49

    3.3Production and characterization of PHB homopolymer using glycerol as a carbon

    source by B. cepacia.52

    3.3.1 Effects of glycerol content on bacterial growth52

    3.3.2 Effects of glycerol content on molecular mass of PHB.53

    3.3.3 Characterization of PHB end-capped with glycerol by1H NMR.55

    3.3.4 Physical properties of PHB..56

    3.3.5 Pilot scale (200-L) fermentation using biodiesel-glycerol..57

    3.3.6 Injection-molding process for conversion of PHB to biodegradable eartips58

    3.3.7 Thermal degradation of PHB..60

    3.4 Production and characterization of PHB-co-HV copolymers using glycerol and

    levulinic acid as substrates in shake flasks by B cepacia..61

    3.4.1 Bacterial growth and production of PHB-co-HV copolymers when co-feeding

    glycerol and levulinic acid in shake flasks.61

    3.4.2 1H and 13C-NMR analysis for the structure and composition of PHB-co-HV

    copolymers62

    3.4.3 Physical property test of PHB-co-HV copolymers..65

    3.4.4 Mechanical property test of PHB-co-HV copolymers..67

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    7/154

    vi

    3.5 Production and characterization of PHB-co-HV copolymers using glycerol and

    levulinic acid as substrates in fermentors by B cepacia72

    3.5.1 Bacterial growth and PHB-co-HV copolymers when co-feeding glycerol and

    levulinic acid in fermentors by B. cepacia in fermentors..72

    3.5.2 Crystallization temperatures of PHAs.74

    3.5.3 Melting temperatures of PHAs75

    3.5.4 Decomposition temperatures of PHAs..77

    3.5.5 Decompositon temperature of PHAs with nucleating agents...78

    3.5.6 Melting temperature and crystallization temperature of PHAs with

    ULTRATALC60979

    3.6 Effects of different aging periods on melting temperatures of PHAs.82

    3.7 Production of P3HB-co-4HB by B. cepaciausing -butyrolactone/1,4-butanediol83

    3.8 Confirmation of HV mol fraction in the PHB-co-HV copolymers by GC analysis and

    1H-NMR.85

    3.9 Solvent extraction of PHAs..86

    3.9.1 Determination of the volume of chloroform for maximum extraction

    efficiency 86

    3.9.2 Determination of incubation temperature for maximum extractionefficiency.88

    3.9.3 Determination of the incubation period for maximum extraction efficiency.89

    4 Discussion.91

    4.1 Renewable and inexpensive feedstocks for PHA production..91

    4.2 Bacterial growth and Properties of PHB produced from glycerol99

    4.2.1 Bacterial growth ofB. cepacia using glycerol as a carbon source..99

    4.2.2 Properties of PHB produced from glycerol as a carbon source.100

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    8/154

    vii

    4.3 Production and characterization of the PHB-co-HV copolymers produced from

    biodiesel-derived glycerol and levulinic acid103

    4.3.1 Production of the PHB-co-HV copolymers using biodiesel-derived glycerol and

    levulinic acid103

    4.3.2 Physical properties of the PHB-co-HV copolymers produced from biodiesel-

    derived glycerol and levulinic acid104

    4.4 Mechanical properties of the PHB homopolymer and the PHB-co-HV copolymer107

    4.5 Effects of nucleating agents on physical and mechanical properties of PHAs..111

    4.6 Economic considerations of PHA production..112

    4.6.1 Economic considerations of PHA production from renewable and inexpensive

    feedstocks.113

    4.6.2 Economic considerations of PHA production by selective downstream isolation

    processes114

    5 Conclusions118

    6 References.123

    7 Appendices134

    8 Curriculum Vitae139

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    9/154

    viii

    List of Tables

    Table 1. Comparison of physicochemical and mechanical properties of selected

    polyhydroxyalkanoate (PHA) polymers and petroleum-derived plastics4

    Table 2. Physical and mechanical properties of PHB-co-HV copolymers.6

    Table 3. The effect of concentrations and exposure times of glycerol on the number-

    average molecular weight (Mn) of PHB..54

    Table 4. Physical-chemical properties of PHB produced from xylose and glycerol..56

    Table 5. Comparison between the original PHB and the thermally degraded PHB..60

    Table 6. Mechanical properties of PHB, PHB-co-HV copolymers and polypropylene..71

    Table 7. Composition and molecular masses of PHB and copolymers of PHB-co-HV.73

    Table 8. Comparison of melting temperatures (Tm) of PHAs detected at different aging

    times.82

    Table 9. Comparison of HV mol% in the PHB-co-HV copolymers detected by GC and1H

    NMR.85

    Table 10. Renewable and inexpensive feedstocks used for PHA production.98

    Table 11. Comparison of dry cell mass and PHA content from different bacterial strains

    grown on various carbon sources in shake flasks..100

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    10/154

    ix

    List of Figures

    Figure 1. General structure of polyhydroxyalkanoates (a) and copolymers (b) poly-(3-

    hydroxybutyrate-co-3-hydroxyvalerate) (abbr. PHB-co-HV) and poly-(3-hydroxybutyrate-

    co-4-hydroxybutyrate) (abbr. P3HB-co-4HB)..1

    Figure 2. Pathway of PHB biosynthesis.15

    Figure 3. Proposed pathway for metabolism of various carbon sources and medium-

    chain-length PHA production.16

    Figure 4. Thermal degradation of PHB.. 22

    Figure 5. Production of biodiesel and glycerol. Catalysts include alkali and acids.27

    Figure 6. Route of levulinic acid from lignocellulosic biomass28

    Figure 7. Bacterial growth and PHA production using renewable carbon sources (tall

    oil fatty acids, biodiesel-derived glycerol and xylose) by B. cepacia in shake flask

    experiments48

    Figure 8. Bacterial growth and PHB production from various sources of biodiesel-

    derived glycerol by B. cepacia in shake flask experiments50

    Figure 9. Glycerol consumption during fermentation using pure glycerol, FutureFuel

    glycerol, Twin River glycerol and ESF glycerol by B. cepaica in shake flask experiments.51

    Figure 10. Dry biomass of B. cepacia grown in shake flasks on different

    concentrations of pure glycerol...52

    Figure 11. Number-average molecular weight (Mn) and weight-average molecular

    weight (Mw) of PHB produced by B.cepacia grown with different concentrations of

    glycerol..53

    Figure 12.1H-NMR of PHB produced by B .cepacia grown on 7% (v/v) glycerol as a

    carbon source. The expanded region indicates glycerol as the terminal end-group55

    Figure 13. Changes in biomass, PHA% and glycerol concentration during a fed-batch

    fermentation, in a 400-L fermentor, with periodic additions () of biodiesel-glycerol57

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    11/154

    x

    List of Figures (Continued)

    Figure 14. (a) Purification and vacuum drying process of PHB, (b) Schematic flowsheet

    of injection-molding process, (c) Eartips made from polypropylene (black, left) and PHB(brown, right two). 59

    Figure 15. Effects of levulinic acid concentration on bacterial growth and production of

    PHB-co-HV copolymers61

    Figure 16. (a) Chemical shift expansions of 300 MHz1H NMR spectra from PHB-co-HV

    copolymers, illustrating a progressive increase in the mol% of HV and quantified by the

    integrated areas of the HB doublet (methyl group of HB, 1.27 ppm) and the HV triplet

    (methyl group of HV, 0.90 ppm). (b) Chemical shift assignments of 300 MHz13

    C NMR

    spectrum of PHB-co-35.8 mol% HV, which was produced by B. cepacia using 3% (v/v)

    glycerol and 0.9% (w/v) levulinic acid..63

    Figure 17a. DSC curves displaying melting temperatures of PHB-co-HV copolymers66

    Figure 17b. Melting temperatures and glass transition temperatures of PHB-co-HV

    copolymers with increasing HV mol%...............................................................................67

    Figure 18a. Stress-strain curves of PHAs and polypropylene. Dot line: PHB (produced

    by B. cepacia using xylose as a carbon source); Dash dot line: PHB-co-17.6 mol% HV;

    Solid line: PHB-co-17.6 mol% HV; Dash line: polypropylene.69

    Figure 18b. Real-time images of stretching for different types of PHAs in tensile testing.

    i represents PHB homopolymer, of which the dogbone was pulled for less than 0.7 mmto break; ii displays PHB-co-29.5 mol% HV, of which the dogbone was stretched for

    approximately 35 mm to break; iii represents for PHB-co-33 mol% HV, of which the

    dogbone was pulled for around 66 mm to break.70

    Figure 19. Growth (dry cell mass, DCM, ), PHA yield () and HV content in the

    copolymer () produced by B. cepacia using glycerol and levulinic acid as carbon

    sources in a 7 L fermentor72

    Figure 20. Crystallization temperatures (Tc) of PHB () and copolymers of PHB-co-HV

    () with increasing mol% HV..75

    Figure 21. Melting temperatures (Tm) of PHB () and copolymers of PHB-co-HV ()

    with increasing mol% HV76

    Figure 22 Decomposition temperatures (Tdecomp,) and melting temperatures (Tm, )

    of PHB and copolymers of PHB-co-HV with increasing mol% HV. Bars indicate

    temperature differential (Tdecomp - Tm).77

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    12/154

    xi

    List of Figures (Continued)

    Figure 23. Effects of nucleating agents on decomposition temperatures of PHB (),

    PHB with 5% ULTRATALC609 (

    ), and PHB with 1% HPN-68L (

    ).79

    Figure 24. Effects of 5% ULTRATALC609 on melting temperature (Tm) of PHAs

    containing increasing mol% HV (--PHB,--PHB-co-5.6 mol% HV,--PHB-co-11.4 mol%

    HV,--PHB-co-14.7 mol% HV, --PHB-co-17.9 mol% HV,--PHB-co-30.5 mol% HV, --

    PHB-co-32.6 mol% HV)...80

    Figure 25. Crystallization temperatures of PHAs () and PHAs with 5% ULTRATALC609

    (). Bars indicate temperature differentials.81

    Figure 26.1H NMR spectrum of P3HB-co-3 mol% 4HB produced from 1,4-butanediol.

    Peaks 2,3,4 are typical chemical shifts for P3HB and peaks 6,7,8 (highlighted by the

    triangles) represent protons of P4HB84

    Figure 27. Relationship between PHA extraction efficiency and dosage of chloroform.

    Incubation of the mixture at 55 C overnight..87

    Figure 28. Relationship between PHA extraction efficiency and incubation temperature.

    Incubation of the mixture at room temp. when stirring or 55 C while standstill.88

    Figure 29 Relationship between PHA extraction efficiency and incubation period.

    Incubation of the mixture at room temperature when stirring.89

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    13/154

    xii

    Abstract

    C. J. Zhu. Production and Characterization of Polyhydroxyalkanoates (PHAs) by Burkholderia

    cepacia ATCC 17759 Grown on Renewable Feedstocks. 141 Pages, 11 tables, 29 figures, 2011

    This thesis describes the microbial production of polyhydroxyalkanoates (PHAs)from bench- and pilot-plant scale fermentations by B. cepacia, followed by physical-

    chemical and mechanical characterization of these polyesters. B. cepacia was evaluated

    to utilize several renewable and inexpensive feedstocks, e.g. wood hydrolysate,

    biodiesel-derived glycerol, cheese whey permeate and tall oil fatty acids from the paper

    pulping process. Glycerol was found to be the best carbon source to support bacterial

    growth and poly-3-hydroxybutyrate (PHB) production based on the highest dry cell mass

    (DCM, 5.8 g/L) and the highest PHB yield (82% of DCM). Increasing the glycerol

    concentration from 3% to 9% (v/v) resulted in a gradual reduction of biomass, PHB yield

    and molecular mass (Mn and Mw) of PHB.1H-NMR revealed that molecular masses

    decreased due to the esterification of PHB with glycerol resulting in chain termination

    (end-capping). Supplementing levulinic acid, derived from lignocellulosic materials, with

    glycerol led to production of the poly-3-hydroxybutyrate-co-3-hydroxyvalerate (PHB-co-

    HV) copolymer. Based on the concentration and timing of levulinic acid added in the

    medium, various mol fractions of HV were incorporated to form various PHB-co-HV

    copolymers. The copolymers exhibited a typical isodimorphic behavior (V-typed shape),

    where upon melting temperature decreased to a minimum point and then increased as

    mol% HV increased. Increasing mol% HV in the copolymer resulted in enhanced

    mechanical properties. Addition of heterologous nucleating agents improved industrial

    processability of PHAs by increasing crystallization temperature. However, HPN-68L was

    not used as a nucleating agent for the polyesters isolated in this study because it

    decreased the decomposition temperature of PHAs. Production of the PHB homo-polymer and the PHB-co-HV copolymers were successfully scaled up for pilot-plant scale

    fermentations. Large quantities of PHAs were isolated, purified and used in the

    fabrication of eartips, which are biodegradable and environmentally friendly, through an

    injection-molding process. These renewable and inexpensive carbon sources and

    alternative downstream PHA isolation process may greatly reduce production cost of

    PHAs, by which the market price of PHAs becomes more competitive with that of

    conventional petroleum-derived plastics.

    Key Words: renewable feedstocks, Burkholderia cepacia, biodiesel-derived glycerol, levulinic

    acid, polyhydroxyalkanoates, physical properties, mechanical properties, pilot-plant scale

    fermentation, downstream PHA recovery

    Author: Chengjun Zhu

    Candidate for the degree of Doctor of Philosophy, August 2011

    Major professor: James P. Nakas, Ph.D.

    Department of Environmental and Forest Biology

    State University of New York College of Environmental Science and Forestry, Syracyse, New York

    James P. Nakas, Ph.D.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    14/154

    1

    1. Introduction

    1.1 Microbial polyhydroxyalkanoates

    Polyhydroxyalkanoates (PHAs) are a class of polyesters which are accumulated as

    microbial intracellular carbon and energy reserves. They exist as discrete inclusions in

    the cell cytoplasm, typically 0.2-0.5 m in diameter1. A variety of PHAs is commonly

    classified into two major groups which are referred to as short-chain-length (scl, 3-5

    carbons) and medium-chain-length (mcl, 6-14 carbons)2. The scl-PHAs are semi-

    crystalline thermoplastics, whereas mcl-PHAs are more elastomeric3. A large number of

    microorganisms, encompassing Gram-positive 4, 5 and Gram-negative 2, 5-7 bacteria, have

    been reported to produce various types of PHAs.

    m=1, R=H poly-3-hydroxypropionate

    R=CH3 poly-3-hydroxybutyrate

    R=CH2CH3 poly-3-hydroxyvalerateR=(CH2)xCH3, x=2,3,,11 mcl-poly-3-hydroxyalkanoates

    m=2, R=H poly-4-hydroxybutyrate

    m=3, R=H poly-5-hydroxyvalerate

    Figure 1. General structure of polyhydroxyalkanoates (a) and copolymers (b) poly-(3-

    hydroxybutyrate-co-3-hydroxyvalerate) (abbr. PHB-co-HV) and poly-(3-hydroxybutyrate-

    co-4-hydroxybutyrate) (abbr. P3HB-co-4HB)

    O

    On

    O

    O

    m

    5

    6

    7

    8

    4

    3

    21

    O

    O

    n O

    O

    m

    a b

    O

    CH

    CH2

    C

    R O

    nm

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    15/154

    2

    Poly-3-hydroxyalkanoates are the most widely studied PHAs, which differ in the

    length of side chain (R group depicted in Figure 1a) and can also differ in composition

    (e.g. copolymers depicted in Figure 1b, terpolymers, etc.). However, poly-4-

    hydroxyalkanoates8, 9

    and poly-5-hydroxyalkanoates10

    have also been observed in the

    past several decades (Figure 1a). Over 150 different monomer units have been

    identified as constituents of these storage molecules5, 11, 12

    . The large variety of PHA

    components results in an enormous diversity of material properties, which is beneficial

    for various potential applications. PHAs have recently received increased attention due

    to their physical and mechanical properties resembling those of petroleum-derived

    plastics. These characteristics make PHAs ideal substitutes for petroleum-based plastics,

    especially when global oil prices remain at relatively high levels. Also, PHAs are

    completely degraded in the environment to CO2 and H2O13, 14

    , compared to recalcitrant,

    non-biodegradable conventional plastics, such as polypropylene, polyethylene and

    polystyrene. Since PHAs are generally biosynthesized using photosynthetically-based

    renewable carbon feedstocks and their end-use materials are biodegradable to carbon

    dioxide and water by microbial extracellular PHA depolymerases15, 16

    , production and

    disposal of these PHA biopolymers constitute a sustainable, closed life cycle process

    with much less energy consumption and greenhouse gas emission17-19

    .

    PHA polymers are synthesized as membrane-bound storage materials by a variety

    of microorganisms when exogenous carbon sources are provided in excess and their

    growth is impaired by the lack of at least one other nutrient, such as nitrogen, sulfate,

    phosphate, magnesium and oxygen6, 20

    . Therefore, microbial fermentation for PHA

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    16/154

    3

    production is regulated largely by the concentrations and contents of these nutrients in

    the medium. Optimal feeding strategies of these nutrients dramatically enhance

    microbial growth and PHA production yields. As reported by Steinbuchel et al.21

    , PHAs

    are accumulated to levels as high as 90% (w/w) of the dry cell mass.

    Furthermore, as a PHA-producing microbe enters the starvation stage due to a

    carbon deficiency, PHA polymers will function as intracellular carbon and energy

    reserves and initiate the degradation process of PHAs to provide carbon and energy for

    microbial survival in a harsh environment22

    . Therefore, PHAs, under natural rules of

    survival, have been designed and developed by a large number of microorganisms as a

    food and energy source during times of nutritional stress.

    1.2 Physicochemical and mechanical properties of PHAs

    PHAs are highly recommended as an ideal substitute for petroleum-derived

    plastics in large part due to their similar material properties (Table 1), in terms of

    processability, strength and industrial fabrication into commodity plastic products23, 24

    .

    Physicochemical and mechanical properties of PHA copolymers can be controlled and

    regulated by variation of the mole fractions of the monomeric constituents (Figure 1b),

    which could be achieved by careful selection and control of fermentation carbon

    sources.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    17/154

    4

    Poly-3-hydroxybutyrate (PHB) is the most common type of PHA produced by

    bacteria. The PHB homopolymer is a highly crystalline, stiff, yet relatively brittle,

    material. Correspondingly, as shown in Table 1, PHB exhibits high tensile strength (43

    MPa), but low elongation to break (5%)25

    . The copolymer poly-3-hydroxybutyrate-co-20

    mol%-3-hydroxyvalerate (PHB-co-20 mol% HV) exhibits lower crystallinity, less stiffness

    (20 MPa), but higher elasticity and flexibility (50% elongation to break)25

    compared to

    the homopolymer PHB. Incorporation of different monomeric subunits, such as 4-

    hydroxybutyrate (4HB)25

    , 3-hydroxyhexanoate (HHx)25

    and other mcl-

    hydroxyalkanoates26

    (e.g. 3-hydroxyoctanoate [3HO], 3-hydroxydecanoate [3HD], 3-

    hydroxydodecanoate [3HDD]), with 3HB will result in copolymers with varying material

    Table 1 Comparison of physicochemical and mechanical properties of selected

    polyhydroxyalkanoate (PHA) polymers and petroleum-derived plastics25

    PolymersCrystallinity

    a

    (%)

    Tmb

    (C)

    Tgc

    (C)

    Tensile

    strength

    (MPa)

    Elongation

    to break

    (%)

    PHB 60 177 4 43 5

    PHB-co-20%HV 56 145 -1 20 50

    PHB-co-16%4HB 45 150 -7 26 444

    PHB-co-10%HHx 34 127 -1 21 400

    Polypropylene

    50-70 176 -10 38 400

    Polyethylene (LDPE)d

    20-50 130 -36 10 620

    Notes:a degree of crystallinity b melting temperature c glass transition temperatured LDPE:low density polyethylene

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    18/154

    5

    properties to address numerous applications in medical and engineering fields.

    Compared to commercially available polylactic acid (PLA) which is also renewable and

    biodegradable27

    , diverse combinations of PHA monomeric units are superior to PLA,

    which only consists of a single monomer-lactic acid.

    In addition to regulating material properties by controlling the composition of

    PHAs, it is possible to change material properties of PHAs with the same composition by

    incorporating various fractions of the co-monomer in the copolymers. This thesis mainly

    describes the production and characterization of poly-3-hydroxybutyrate-co-3-

    hydroxyvalerate (PHB-co-HV) copolymers, biosynthesized from glycerol and levulinic

    acid as carbon sources. As the mol fraction of HV in the copolymer PHB-co-HV varies,

    the physical and mechanical properties of PHB-co-HV will correspondingly change (Table

    2). When HV ratios of PHB-co-HV increased from zero (the homopolymer of PHB) to 25%,

    the melting temperatures (Tm) and glass transition temperatures (Tg) gradually

    decreased from 179 C to 137 C and 10 C to -6 C, respectively. As shown in Table 2,

    PHB-co-HV copolymers also became more flexible (as indicated by the decrease of

    Youngs modulus) and tougher (as demonstrated by the increase in impact strength) as

    the HV ratio increased.

    Besides PHB-co-HV, similar patterns were observed with copolymers PHB-co-

    HHx

    28-30

    , P3HB-co-4HB

    20

    , PHB-co-HO

    31

    , PHB-co-HD

    31

    , when the ratios of these co-

    monomers, such as HHx, 4HB, HO or HD, in their copolymers increased.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    19/154

    6

    Table 2 Physical and mechanical properties of PHB-co-HV copolymers20

    .

    Mol fraction

    (mol%) Tm (C) Tg (C)YoungsModulus

    (GPa)

    TensileStrength

    (MPa)

    Notched IzodImpact

    Strength (J/m)3HB 3HV

    100 0 179 10 3.5 40 50

    97 3 170 8 2.9 38 60

    91 9 162 6 1.9 37 95

    86 14 150 4 1.5 35 120

    80 20 145 -1 1.2 32 200

    75 25 137 -6 0.7 30 400

    PHB-co-HV copolymers with various HV mol fractions exhibit high degrees of

    crystallinity (between 55% and 70%)32

    . PHB-co-HV is unique among the PHA family of

    copolymers in that the size and structure of HB and HV monomers are similar. Their

    similarity allows HB and HV to participate in a co-crystallization process, in which HV can

    be incorporated into the HB crystal lattice and vice versa. This phenomenon is termed

    isodimorphism32, 33

    . As a result, the melting temperatures of PHB-co-HV gradually

    decrease to a minimum point, then increase as the HV mol fraction increases. Therefore,

    isodimorphism and the transition between the HB crystal lattice and the HV crystal

    lattice typically demonstrate a V-shaped pattern (see melting temperatures of PHB-co-

    HV in results and references32, 33

    ). Incorporation of comonomers with PHB to form

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    20/154

    7

    copolymers can lower melting temperatures, by which PHA copolymers demonstrate an

    important advantage of industrial applications for melt processing at lower

    temperatures.

    However, the PHA family displays a rather slow crystallization process due to high

    purity and limited heterogeneous nuclei34

    . Basically, isolation of PHAs from cells using

    solvents excludes most of the impurities from cell components. During the melt-

    quenching process, PHAs at high purity crystallize at a relatively low rate, which causes a

    longer manufacturing cycle and a less efficient industrial fabrication process for finished

    products. PHB and PHB-co-HV have been extensively studied for their nucleation

    behaviors34, 35

    . Pure PHB exhibited a very slow nucleation, though self seeding to some

    extent, when it was cooled from a melt34

    . The nucleation density of pure PHB is too low

    to initiate efficient crystallization. At the same time, limited nuclei formed limited

    quantities of spherulites so that the size of each PHB spherulite is relatively large, which

    makes PHB somewhat brittle and subject to cracking36, 37. PHB-co-HV also exhibited a

    slow crystallization behavior, which resulted in the films made from the copolymers

    with a higher HV ratio tacky and even sticky to themselves after cooling for an extended

    period of time35, 38

    .

    Impurities usually behave as external nuclei for PHAs. They can increase not only

    the rate of crystallization, but also the quantities of spherulites corresponding to the

    numbers of nuclei. The large quantities of nuclei make the size of spherulites relatively

    small, thus improving material mechanical properties36

    . Accounting for the slow

    crystallization process of PHB and PHB-co-HV, heterogeneous nucleating agents instead

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    21/154

    8

    of impurities, which are difficult for the quantification and qualification, need to be

    supplemented with these polymer melts to speed up the crystallization process by

    increasing the crystallization temperature and forming small spherulites.

    Various nucleating agents, such as orotic acid39

    , -cyclodextrin40

    , boron nitride,

    talc, terbium oxide, lanthanum oxide35

    , saccharin and phthalimide41

    have been tested

    for enhancing the crystallization of PHAs. These nucleating agents will generally increase

    crystallization temperatures, accelerate crystallization rates and enhance PHA stability

    during heating. However, some nucleating agents have negative effects on PHAs by

    decreasing their decomposition temperatures (Tdecomp.). Hydroxyapatite, which is a

    naturally occurring mineral of calcium apatite, decreased the onset Tdecomp from 260oC

    to 225oC when the hydroxyapatite content in the hybrid PHB-hydroxyapatite composite

    increased from 0 to 10%42

    . These side effects of nucleating agents have not been

    described in many previous studies. Comprehensive physical property tests, including

    crystallization temperature, melting temperature, glass transition temperature and

    decomposition temperature, have been performed in this research (see results).

    1.3 History and development of PHAs

    Poly-3-hydroxybutyrate, as an intracellular storage material in Bacillus megaterium,

    was first identified by Lemoigne in 192643

    . At that time, PHB was discovered

    unexpectedly when Lemoigne attempted to determine the cause of acidification in an

    aqueous suspension of the bacterium Bacillus megaterium under oxygen-free conditions.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    22/154

    9

    Although methods for PHB detection and quantification were limited, Lemoigne and co-

    workers published 27 papers from 1923 to 1951. Not until the late 1950s did microbial

    physiologists finally recognize the importance of PHB in the overall metabolism of

    bacterial cells44

    .

    The rediscovery of PHB occurred simultaneously and was reported independently

    in 1958 and 1959. Williamson et al. 45 at the University of Edinburgh in Scotland treated

    the cells of various Bacillus spp. with an alkaline solution of sodium hypochlorite, and

    found large amounts of PHB (89%) with the ether-soluble lipid (11%) in some Bacillus

    species, and elucidated the function of PHB in the cell. Doudoroff and Stanier46

    at the

    University of California at Berkeley discovered that PHB was the primary product of the

    oxidative and photosynthetic assimilation of organic compounds by Pseudomonas

    saccharophila and a phototropic bacterium, Rhodospirillum rubrum, respectively. When

    external carbon sources were removed from the medium, there occurred a fairly rapid

    intracellular breakdown of the storage PHB. Therefore, PHB was putatively thought to

    play a role analogous to that of starch and glycogen in the metabolism of other

    organisms. Doudoroff et al.46

    attempted to clarify the biosynthesis and degradation

    mechanism of PHB in microbial cells. In the 1960s, these aforementioned authors47, 48

    isolated native PHB granules from a chemoheterotroph and phototroph, Bacillus

    megaterium and Rhodospirillum rubrum. The native PHB granules from R. rubrum

    retained the active PHB synthase and depolymerase that degrade PHB to the monomer,

    and the isolated granules from B. megaterium only exhibited the synthase associated

    with PHB. These authors also determined that PHB granules in bacteria actually serve as

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    23/154

    10

    an intracellular energy and carbon reserve and that PHB is produced in response to a

    nutrient limitation when at least one essential element in the environment is exhausted.

    In 1965, Marchessault and his co-workers

    49

    at SUNY-ESF collected PHB samples

    from various microorganisms and employed X-ray diffraction to explore the crystal

    structure of PHB. All PHB samples precipitated from chloroform displayed an uniformity

    of crystal structure and molecular structure due to identical X-ray diffractograms and

    infrared spectra, respectively. The molecular mass of PHB isolated by alkaline

    hypochlorite was also found to be uniformly low compared to that of PHB prepared by

    direct solvent extraction.

    In 1974, Wallen and Rohwedder50

    reported that the polyesters, isolated from

    activated sewage sludge, exhibited similar but not identical NMR and infrared spectra.

    Gas chromatographic analysis indicated a mixture of C4, C5, C6 and C7 components in

    these polyesters. The identification of a variety of PHAs in addition to PHB opened

    avenues of research for PHAs regarding their material properties.

    Since rising oil prices and an unstable oil supply became global issues and

    environmental concerns over limited fossil reserves drew more and more attention, the

    advantage of PHAs as biodegradable alternatives to petroleum-derived plastics was

    gradually recognized by researchers and much effort has been devoted to investigation

    of PHA properties. In addition, much time and effort has been dedicated to genetic

    engineering of microorganisms or plants for biosynthesis of PHAs from renewable

    feedstocks with improved material properties for different end-use applications.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    24/154

    11

    Prompted by the oil crisis in the 1970s, the first industrial production of PHB was

    introduced in 1982 by Imperial Chemical Industries (ICI) in the UK, employing a two-

    stage, fed-batch fermentation process and using a sugar-based carbon source for

    growth of Ralstonia eutropha. However, PHB was produced at a relatively high

    production cost due to expensive feedstocks and downstream processing for PHB

    isolation. Compared to conventional petroleum-derived plastics, higher cost has placed

    PHAs in a weaker economic position and mechanical properties, such as high

    crystallinity and brittleness, have resulted in a rather limited range of applications20

    .

    Based on these limitations of the homopolymer PHB, PHB-co-HV copolymer, which

    exhibits enhanced toughness and flexibility, attracted more attention and was first

    produced on a commercial scale in the late 1980s by ICI, which marketed their PHB-co-

    HV products under the trade name of Biopol6, 51

    . Biopol (PHB-co-HV copolymer) was

    produced by R. eutropha, using glucose and propionate as carbon sources, at a reported

    production cost of US$ 16/kg, which was 18-fold higher than conventional

    polypropylene52

    . The prohibitively high price of PHAs and the relatively low prices of

    commodity plastics made the commercial-scale production of PHAs unrealistic.

    Eventually Biopol was acquired by Zeneca Ltd in 1990 until it was acquired in 1996 by

    Monsanto (St. Louis, MO), which utilized their considerable expertise in plants and

    initiated research to produce PHB and related copolymers photosynthetically in plants51.

    Subsequently, Monsantos right to Biopol was sold to Metabolix (Cambridge, MA)

    in 200153

    . Recombinant Escherichia colistrains, which exhibit broad nutritional diversity

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    25/154

    12

    and relatively fast growth, have been engineered to produce PHAs from inexpensive

    carbon sources. Mirel (trademark of Telles, an affiliated company of Metabolix and

    Archer Daniels Midland Company) bioplastics were announced to be produced using

    corn starch sugars, starting in 2009 in Clinton, IA (www.mirelplastics.com). Several other

    companies are also developing and commercializing various types of PHAs, including

    Procter & Gamble (Cincinnati, USA) which introduced novel PHA copolymers with HB as

    a monomeric unit and HHx (C6), HO (C8) or HD (C10) etc. as monomeric units, all of which

    are under the trademark ofNodax54, 55

    . The projections for Nodax marketing consist

    of PHA production of 100-1000 metric ton in 2004 and the delivery plan of approximate

    $ 1/lb ($ 2.2/kg) in 2005/200656

    . Since the 1980s, BASF in Germany developed a pilot-

    scale fermentation process for production of PHB and PHB-co-HV, which were

    supplemented and blended with its biodegradable polymer Ecoflex, which are

    petroleum-derived aliphatic-aromatic copolyesters57

    . Tianan Biologic in China started to

    produce PHB-co-HV in 2000 using glucose, fructose and organic acids as carbon sources

    (www.tianan-enmat.com), and have currently scaled up to commercial production of

    2000 metric tons per year57

    .

    Most of the above-mentioned commercial PHAs have been produced by various

    microorganisms using pure carbon sources, such as glucose, sucrose, propionate and

    lauric acid, which are always more expensive than renewable feedstocks, including

    wood-based hydrolysate (xylose and levulinic acid)58

    , cheese whey permeate (including

    lactose)59

    , sugarcane molasses (sucrose, fructose, glucose and trace amount of

    maltose)60, 61

    , corn steep liquor (nitrogen sources)60

    , soybean oil62

    , and crude glycerol

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    26/154

    13

    from biodiesel-producing facilities63

    . Regarding environmental concerns of

    commercializing PHAs, the production of PHAs from renewable and biobased feedstocks,

    e.g. wood hydrolysate and biodiesel-derived glycerol, need to be taken into account at a

    higher priority.

    1.4 Metabolic pathway leading to the biosynthesis of PHAs

    from various carbon sources

    Uptake of carbon sources by microorganisms supports their growth and other

    basic vital functions. Carbon sources in the environment have to be translocated into

    the cytosol inside the cell, the location for nutrient metabolism. In this translocation

    process, transporters, which are transmembrane proteins, either actively transport the

    molecules of carbon sources by utilizing ATP (adenosine triphosphate) energy (e.g ATP-

    binding cassette transporters, ABC-transporters) or passively translocate these

    molecules without energy expenditure (e.g glycerol diffusion facilitator protein, GlpF).

    Therefore, whether a certain carbon source can be utilized for microbial growth is first

    determined by the transporters which can recognize and transport this molecule.

    Relevant to this research, xylose (a five carbon sugar) and/or glycerol were used as

    carbon sources by Burkholderia cepacia. D-xylose is transported by XylFGH transporter

    (possibly belonging to the ABC superfamily) into the cytosol and converted to D-xylulose

    by xylose isomerase (encoded by xylA gene), subsequently phosphorylated to D-

    xylulose-5-phosphate by xylulokinase (encode by xylB gene)64, 65

    (Figure 3). Glycerol is

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    27/154

    14

    translocated by GlpF into the cell and phosphorylated by glycerol kinase (encoded by

    glpK), leading to glycerol-3-phosphate. Glycerol-3-phosphate dehydrogenase (encoded

    by glpD) oxidizes glycerol-3-phosphate to dihydroxyacetone phosphate (DHAP)66

    (Figure

    3). Further metabolism of these intermediates (e.g. D-xylulose-5-phoaphate and DHAP)

    derived from the carbon sources through either glycolysis (Embden-Meyerhofpathway)

    or the pentose phosphate pathway or Entner-Doudoroff pathway results in the end

    product pyruvate, which is converted through oxidative decarboxylation by pyruvate

    dehydrogenase to acetyl-CoA67

    .

    In this research, levulinic acid (4-ketovaleric acid) was employed as the co-

    substrate to provide the precursor propionyl-CoA for biosynthesis of poly-3-

    hydroxyvalerate. Metabolism of levulinic acid is not clearly elucidated, but tentatively

    considered to form one propionyl-CoA and one acetyl-CoA through -oxidation. The

    enzymes involved are still ill-defined21

    .

    Three key enzymes -ketothiolase, acetoacetyl-CoA reductase and PHA synthase,

    encoded by phaA, phaB and phaC, respectively, are involved in the last three steps of

    PHA biosynthesis. Two acetyl-CoA moieties are condensed to acetoacetyl-CoA by -

    ketothiolase, and acetoacetyl-CoA is reduced to (R)-3-hydroxybutyryl-CoA by

    acetoacetyl-CoA reductase, followed by polymerizing the precursor 3-hydroxybutyryl-

    CoA to the polymer poly-3-hydroxybutyrate68

    . Similarly for poly-3-hydroxyvalerate

    synthesis, one acetyl-CoA and one propionyl-CoA are condensed to 3-ketovaleryl-CoA,

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    28/154

    15

    with subsequent reduction to 3-hydroxyvaleryl-CoA, followed by polymerization for PHV

    synthesis (Figure 2).

    acetyl-CoA

    acetoacetyl-CoA

    (R)-3-hydroxybutyryl-CoA

    PHB

    3-ketothiolase (PhaA)

    acetoacetyl-CoA reductase (PhaB)

    PHA synthase (PhaC)

    CoASH

    NADP+

    NADPH

    SCoA

    O

    SCoA

    O O

    SCoA

    OH O

    CoASH

    O

    O

    n

    2

    Figure 2 Pathway of PHB biosynthesis

    Biosynthesis of medium-chain-length PHAs, which are comprised of the

    constituents C6-C14 chains, recruits PHA-specific enzymes such as (R)-specific enoyl-CoA

    hydratase (PhaJ), putatively (R)-3-hydroxyacyl ACP thioesterase (PhaG) and acyl-CoA

    ligase (AlkK) (Personal communication with Qin Wang and Christopher Nomura at SUNY-

    ESF) to divert the intermediates (such as enoyl-CoA and (R)-3-hydroxyacyl acyl carrier

    protein (ACP), respectively) of fatty acid -oxidation pathway and fatty acid biosynthesis

    pathway for the formation of precursor (R)-3-hydroxyacyl-CoA (Figure 3), which leads to

    biosynthesis of medium-chain-length PHAs by PHA synthase (PhaC)69

    .

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    29/154

    16

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    30/154

    17

    Figure 3 Proposed pathway for metabolism of various carbon sources and medium-

    chain-length PHA production69

    . GAP: glyceraldehyde-3-phospahte. DHAP:

    dihydroxyacetone phosphate. (1) or (2) represents one or more steps in glycolysis

    pathway or pentose phosphate pathway, respectively. : cell membrane

    While the pathway for PHA biosynthesis has been gradually elucidated in the last

    several decades, construction of desired strains for PHA production seems to be a

    promising path and has greatly stimulated and enhanced PHA research. These

    engineered strains, such as E. coli70

    , Pseudomonas putida71

    and Cupriavidus necator62

    (formerly known as Ralstonia eutropha or Alcaligenes eutrophus) demonstrate the

    advantages of broad nutritional diversity (using various carbon sources, including

    inexpensive feedstocks), relatively rapid growth to high cell density, accumulation of

    high intracellular concentrations of PHAs and biosynthesis of novel PHAs which cannot

    be accomplished in native strains and which exhibit diverse material properties for

    various end-use applications.

    1.5 Biodegradation and thermal degradation of PHAs

    Although this thesis focuses on the production and characterization of PHAs using

    inexpensive carbon sources, some consideration must be given to the biodegradability

    of PHAs, which is a distinct advantage of PHAs over conventional plastics. Also, since

    some applications of PHAs need to be conducted at high temperature, it is reasonable

    to examine the thermal degradation of PHAs.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    31/154

    18

    1.5.1 Biodegradation of PHAs

    Petroleum-derived plastics are xenobiotic compounds, which are recalcitrant to

    degradation and take several decades, even over 100 years, to degrade in the natural

    environment72

    . One of the most important characteristics of PHAs as substitutes for

    conventional plastics is that PHAs are biodegradable. In nature, a vast consortium of

    microorganisms, by using intracellular or extracellular PHA depolymerases, will degrade

    PHAs, which either are stored inside of the cells or exist in the natural environment. It is

    noteworthy that intracellular PHA depolymerases cannot hydrolyze extracellular PHAs

    and extracellular PHA depolymerases are also unable to degrade intracellular PHA

    granules16

    . Apparently, these differences result from the biophysical structures of

    intracellular native PHA granules and extracellular denatured PHAs. The former are

    completely amorphous elastomers73

    , however, the latter are known for their high

    crystallinities2. Native PHA granules with a particular surface layer containing protein

    and phospholipids are denatured by losing this surface layer during the isolation

    processes of PHAs, leading to semicrystalline polymers that exhibit an ordered helical

    crystal structure, in which the remaining amorphous polymers are embedded13

    .

    In the remainder of this thesis, biodegradability of PHAs, if mentioned, will be

    concerned with the extracellular denatured PHAs using extracellular depolymerases in

    order to demonstrate the superiority of PHAs over petroleum-based plastics.

    Volova et al74

    studied the biodegradability patterns of PHB and PHB-co-11 mol%HV

    in a tropical marine environment. After 160 days, the loss of mass of both PHA films,

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    32/154

    19

    submerged to a depth of 120 cm, approached 50% of the original dry weight of the PHA

    films. Due to a larger surface area, PHA films demonstrated a more rapid rate of

    biodegradation than compacted PHA pellets. The polydispersity increasing in all PHA

    samples suggested that the fragments of polymers with more diverse lengths were

    growing due to random scission of the hydrolyzed polymer chains. Under their study

    conditions, no significant differences were observed for the degradation rates between

    PHB and PHB-co-11 mol%HV, and the degree of crystallinity of both PHAs remained

    unchanged.

    Mergaert et al.75

    tested microbial degradation of PHB and PHB-co-10 mol% HV in

    soils at 15 C, 28 C or 44 C for up to 200 days. These dog bone-shaped PHA samples

    were degraded at an erosion rate of 0.03% to 0.64% weight loss per day, depending on

    the polymer compositions, the types of soil and the incubation temperatures. In

    summary, the degradation was enhanced by incubation at higher temperatures, and in

    most cases the copolymer exhibited a higher erosion rate, based on weight loss, than

    the homopolymer. The degradation also resulted in the loss of mechanical properties

    (based on less elongation to break). In the studied soil (sandy soil, pH 6.5; clay soil, pH

    7.1, loamy soil, pH 6.3; hardwood forest soil, pH 3.9; pinewood forest soil, pH 3.5), 295

    dominant microbial strains capable of degrading PHBand PHB-co-HV in vitro were

    isolated and identified. This same research team76 investigated PHB, PHB-co-10 mol%

    HV and PHB-co-20 mol% HV in situ in natural waters. These polymers were degraded

    rather slowly (less than 7% weight loss after 6 months) in two freshwater ponds.

    However, after 358 days in a freshwater canal, 34% weight loss was recorded for PHB,

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    33/154

    20

    77% for PHB-co-10 mol% HV, and 100% for PHB-co-20 mol% HV. In seawater, within 270

    days, PHB lost 31% of the initial weight and the copolymers lost between 49-52%. The

    degradation rate was observed to be more rapid during the summer due to higher

    temperature. Mergaert et al.77

    also studied biodegradation of PHB, PHB-co-10 mol% HV

    and PHB-co-20 mol% HV in compost. PHB-co-20 mol% HV was degraded much faster (70%

    weight loss) than PHB (6% weight loss) and PHB-co-10 mol% HV (4% weight loss) within

    150 days. From this biodegradation study, as well as two other composts, 109 microbial

    strains capable of degrading PHAs in vitro were isolated and identified.

    Although hundreds of microorganisms are able to secrete extracellular PHA

    depolymerases to degrade PHAs in the natural environment, the biodegradation rates of

    PHAs varied dramatically from several months up to several years as a function of PHA

    composition and shapes of the test samples, and environmental factors, such as

    temperature, types of soil and water, sunlight (UV radiation)78

    , pH16

    , etc..

    In a word, the biodegradability of PHAs in the natural environment makes PHA-

    type thermoplastics ideal green substitutes for conventional plastics, which are resistant

    to degradation in the environment and therefore result in severe environmental

    pollution on a global scale.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    34/154

    21

    1.5.2 Thermal degradation of PHAs

    Considering the injection molding process for industrial fabrication of finished

    products (packaging, eartips, thermometer covers, medical implants, etc.), thermal

    degradation of PHAs must be taken into account during the heating process due to their

    potential thermal instability79

    .

    PHB was extensively studied during melt processing and found to be rather

    unstable at temperatures above or even close to its melting temperature (177 C). Doi

    and his co-workers80

    confirmed that PHB polyester suffered increasing reduction of

    molecular mass within 20 min at 175 C, even 2 C below its melting temperature. In

    addition, all copolyester samples (PHB-co-HV, HV=0-71 mol%; P3HB-co-4HB, 4HB=0-82

    mol%) tested were thermally unstable at temperatures above 170 C, based on rapid

    loss of molecular masses. Below or equal to 160 C, all copolymer samples tested

    demonstrated thermal stability and their molecular masses decreased at a very slow

    rate within 20 min. Especially for the first 2 min of heating at 180 C , most of PHA

    polymers did not show significant decrease in molecular masses. Therefore, PHAs are

    suggested to be heated to melt and kept at that temperature for a short time period (2

    min suggested in the study by Kunioka et al.80

    ).

    It was reported that thermal degradation of PHB occurred by random chain

    scission, a widely accepted six-membered ring ester decomposition process, yielding

    crotonic acid (one product from PHB pyrolysis, Figure 4), 2-pentenoic acid (one product

    from PHV pyrolysis) and other related oligomers81-83

    .

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    35/154

    22

    O CH

    R1

    H2C

    CH3

    C

    O

    H

    CH

    CH

    OCH3

    C

    O C

    O

    R2

    O CH

    R1

    H2C

    CH3

    C

    OH

    O

    +

    HC

    HC

    CH3

    C HC

    CH

    C R2

    O

    OH

    +

    H

    CH

    CH

    CH3

    CH3

    O

    Crotonic acid

    pyrolysis

    Figure 4 Thermal degradation of PHB

    A major concern of PHB and other related PHAs for use as a thermoplastic is

    thermal sensitivity during melt processing82. A potential solution is grafting of certain

    compounds with PHAs. It has been reported that none of the conventional polyolefin

    stabilizers enhanced PHB stability84

    . However, radiation grafting of methyl methacrylate

    (MMA)85

    , 2-hydroxyethyl methacrylate (HEMA)85

    , acrylic acid (AAc)86

    and styrene87, 88

    onto PHB and its copolymers was found to enhance the thermal stability of PHAs. Also,

    grafting maleic anhydride onto PHB remarkably enhanced its thermal decomposition

    temperature79

    . In most cases, a low degree of grafting not only enhanced the thermal

    stability of PHAs, but also promoted the biodegradability of PHAs due to the wettability

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    36/154

    23

    between the polymer and the enzyme solutions. It was also noted that a high degree of

    grafting steeply decreased the biodegradability of PHAs.

    Therefore, it is necessary to strictly control the temperature to melt PHAs during

    heating. Because of the loss of molecular mass of PHAs, mechanical properties of PHAs

    definitely change after heating. In addition, a low degree of grafting helps stabilize PHAs

    when the processing temperature reaches the melting temperature.

    1.6 Considerations and rationale of renewable feedstocks

    and downstream processing

    Compared to the price of conventional petrochemical plastics (less than US $ 1/kg),

    the price of PHAs was prohibitively high (ca. US $ 16/kg Biopol PHAs)89

    , which places

    PHAs at a distinct disadvantage in the market place. By examining the entire flowsheet

    of PHA production, the high price of PHAs is driven by high production costs of PHAs in

    two major areas. One is the cost of feedstocks, and secondly, the downstream recovery

    process of PHAs.

    1.6.1 Renewable feedstocks for PHA production

    A large number of microorganisms have been shown to produce various types of

    PHAs. Although it is feasible to use pure carbon sources, such as glucose, galactose,

    xylose, fatty acids, etc., for PHA production in the laboratory, high production costs for

    carbon feedstocks hinder the scale-up of PHA production towards the commercial level.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    37/154

    24

    Consequently, the cost of carbon sources accounts for as much as 50% of the overall

    production costs89, 90

    . In order to reduce the production costs of PHAs, renewable and

    inexpensive feedstocks have to be taken into consideration to substitute for pure and

    expensive carbon sources at the commercial level.

    As reviewed in 2009 by Chen57

    , the microbial production of PHAs at the industrial

    scale still uses some high purity sugars and fatty acids, such as glucose, sucrose, lauric

    acid, and propionate, during the fermentation process. Although the price of PHAs is

    expected to be approximately US $3-4/kg in the near future, it is obvious that PHAs are

    still more expensive than conventional plastics, especially when the price of oil is

    depressed.

    In order to further reduce the production costs of PHAs, some progress has been

    achieved in the last several years by using inexpensive agricultural or industrial wastes

    for PHA production.

    Khardenavis et al.91

    evaluated waste activated sludge generated from a combined

    dairy and food processing industry wastewater treatment plant for PHB production.

    Jowar or rice grain-based distillery spent wash was used as a carbon source for a PHB

    yield of 42.3% (w/w) and 40% (w/w), respectively. Addition of di-ammonium hydrogen

    phosphate further increased PHB production to 67% (w/w). Meanwhile, mixed culture

    production of PHAs from waste water was evaluated to be financially attractive in

    comparison to pure culture production of PHAs92

    .

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    38/154

    25

    Gouda et al.60

    used sugarcane molasses and corn steep liquor, two of the most

    inexpensive substrates available in Egypt, as sole carbon and nitrogen sources for PHB

    production. The best growth of Bacillus megaterium was obtained with 3% molasses,

    while the maximum yield of PHB at 46.2% (w/w) occurred with 2% molasses. Corn steep

    liquor with a concentration equivalent to 0.05% NH4Cl was the best nitrogen source for

    PHB synthesis (32.7%, w/w).

    Yellore et al.59

    isolated a strain ofMethylobacterium sp. ZP24 from a local pond,

    and tested this strain by using lactose from cheese whey, a byproduct of the dairy

    industry, for PHB production. Pure lactose at a concentration of 12 g/L resulted in

    biomass and PHB yield of 5.25 g/L and 59% (w/w) in 40 h, respectively. Cheese whey,

    used as a carbon source, led to 1.1 g/L PHB and further addition of ammonium sulphate

    increased PHB production from whey 2.5-fold. Nath et al.93

    employed a fed batch

    process to attain a PHB yield of 4.58-fold using limiting dissolved oxygen in the

    fermentor with processed cheese whey supplemented with ammonium sulfate.

    Keenan et al.58

    used aspen and maple wood hydrolysate, which contains xylose

    and glucose as major sugar components and low amounts of galactose, arabinose and

    mannose, supplemented with levulinic acid for PHB-co-HV production. These detoxified

    hydrolysates amended with 0.25%-0.5% levulinic acid were used as feedstocks, and

    resulted in PHB-co-HV yields, PHA% of dry cell mass and mol fraction of HV at 2.0 g/L, 40%

    (w/w), and 16 mol% - 52 mol%, respectively. From an economic standpoint, the

    substrate cost of hemicellulosic hydrolysate was reduced to US $ 0.34/kg PHB,

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    39/154

    26

    compared to US $ 0.58/kg PHB from hydrolyzed corn starch and US $ 1.30/kg PHB from

    pure glucose.

    1.6.2 Production, economics and renewability of glycerol and

    levulinic acid

    In this research, waste glycerol, a byproduct from the biodiesel-producing process

    (Figure 5), was tested as a carbon source for PHA production. Biodiesel-derived glycerol

    has been produced in huge quantity as the production of biodiesel has significantly

    increased. Biodiesel production increased dramatically from 500,000 gallons in 1999 to

    450 million gallons in 2007 (National Biodiesel Board, 2008). The major byproduct of the

    biodiesel industry is glycerol, which is product of approximately 10% of the final weight

    of biodiesel94

    . Consequently, in 2007, glycerol was produced in a quantity of 45 million

    gallons in the United States, and this crude glycerol is not suitable for use in the food,

    pharmaceutical, cosmetics and other industries due to low purity. It is expensive to

    refine crude glycerol to the purity needed for these applications95

    . Biodiesel production,

    as well as glycerol, are at an all-time high. Since the market is glutted, the price of

    glycerol has decreased and raw glycerol from biodiesel-producing companies is now at

    US $ 2.5 cents/lb96

    . Therefore, conversion of crude glycerol into higher-value products,

    such as PHAs, improves the economic viability of the biofuel industry by producing a

    value-added product as well as eliminating the cost of treatment for glycerol disposal.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    40/154

    27

    O

    O

    O

    R1

    R2

    R3

    O

    O

    O

    Alkali,Acids

    OH

    OH

    O

    R1

    O

    R2

    O

    R3

    OH +

    O

    O

    O

    Triacylglycerol GlycerolBiodiesel

    (Fatty acid methyl ester)

    Methanol

    Figure 5 Production of biodiesel and glycerol. Catalysts include alkali and acids.

    Besides glycerol, levulinic acid was employed as a cosubstrate for PHB-co-HV

    copolymer production by Burkholderia cepacia in this research. Levulinic acid (4-

    ketovaleric acid, 4-oxopentanoic acid), an important biomass-derived feedstock, could

    be produced cost-effectively from a wide-array of renewable, hexose-containing

    materials, including forest-based lignocellulosic biomass97, 98

    . Hexoses, such as glucose,

    fructose and mannose, from forest residues are converted under acid dehydration to 5-

    hydroxymethylfurfural (HMF) and subsequently hydrated resulting in the final products

    of levulinic acid and formic acid (Figure 6).

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    41/154

    28

    O

    HO

    HOOH

    OH

    OH

    OHO

    O

    OH

    O

    O

    H2O H2O

    +H OH

    O

    Cellulose,Hemicellulose

    Glucose 5-hydroxymethylfurfural,HMF

    levulinic acid formic acid

    Figure 6 Route of levulinic acid from lignocellulosic biomass

    Levulinic acid, in addition to glycerol, is also among the top 12 value-added

    chemicals from biomass described by the US DOE99

    . It is especially attractive because a

    variety of lignocellulosic biomass, such as rice straw, wood, pulp slurry, corn starch,

    switch grass, sugarcane, etc., can be used for the direct production of this

    thermodynamically stable molecule100. The dehydration of C6 sugars by acids was

    reported to generate levulinic acid with a theoretical maximum yield of 64.5% (w/w)

    because of the concurrent production of equimolar amounts of formic acid101

    .

    Several technologies have been developed for the industrial-scale continuous

    production of levulinic acid, with yields reported between 20% and 48% (w/w)100

    . The

    most promising technology proposed for the large-scale production of levulinic acid was

    patented by the Biofine Corporation (South Glens Falls, NY)102, 103

    . This approach used a

    double-reactor system and minimized the formation of byproducts. The carbohydrate

    feedstocks and sulfuric acid catalyst (1-5 wt% of the feedstocks) were mixed at 210-

    230 C for a short time of 13 to 25 s in the first reactor, in which C 5 and C6 sugars are

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    42/154

    29

    produced. Meanwhile, dehydration of C6 sugars in the first reactor produces HMF.

    Subsequently, HMF is continuously removed from the first reactor and fed into the

    second reactor at 195-215 C for 15 to 30 min for a final conversion to levulinic acid at a

    yield of 50%-70% (w/w). As a result, levulinic acid can be produced at a low cost (US

    $ 0.04-0.1/kg)104

    , which depends on the scale of production and the economic climate of

    the cellulosic feedstocks.

    Besides levulinic aicd which can be obtained from lignocellulosic biomass, tall oil

    fatty acids, mainly consisting of C18 (i.e. 52% oleic acid and 45% linoleic acid58

    ), are

    byproducts of the paper/Kraft pulping process, which is a technology for conversion of

    wood into almost pure cellulose fibers. These free fatty acids are isolated at a

    percentage of 30-40% from crude tall oil by distillation105, 106

    and could be potential

    inexpensive precursors for forest-based PHA production107, 108

    .

    1.6.3 Downstream processing for PHA isolation

    In addition to the production costs contributed by feedstocks, another major cost

    contributor is the recovery process of PHAs. PHAs are known to be stored inside cells

    and the native granules are surrounded by phospholipid membranes. However, only

    purified PHA polymers exhibit the desired physical and mechanical properties of

    thermoplastics. Therefore, removal of other cell components from PHAs presents a

    technical challenge and this isolation process serves to increase the production cost of

    PHAs.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    43/154

    30

    Solvent extraction is the most common PHA recovery technique. Usually,

    halogenated solvents, such as chloroform and dichloromethane, display good solubility

    of PHAs and are widely used in the laboratory. In terms of high efficiency (around 90%)

    and high purity (over 99%), solvent extraction exhibits undoubted advantages over

    other recovery methods109

    . Meanwhile, this method can also remove bacterial

    endotoxins and causes negligible degradation of PHAs. Unfortunately, solvent extraction

    is not viewed as an environmentally benign method, and high costs of these solvents

    also hamper large-scale application in the industry. In addition to the solvents for

    dissolution of PHAs, anti-solvents need to be used for precipitation of PHAs from the

    solvents. Methanol, ethanol and heptane are usually employed to precipitate and purify

    PHAs in this extraction process, in which these solvents might contribute up to 50% of

    the entire production costs of PHAs110

    and are not commercially economical if used at

    an industrial scale for PHA production. However, some non-chlorinated solvents could

    be acceptable alternatives in terms of environmental and economic concerns. Non-

    chlorinated solvents, including ethyl acetate, acetone (hot), cyclic carbonate, methyl

    tertiary butyl ether (MTBE), etc., can be used to partly reduce costs and water could also

    be used as an inexpensive anti-solvent to precipitate PHAs110, 111

    .

    Besides solvent extraction, another strategy for PHA isolation is to remove non-

    PHA cell mass (NPCM) from PHAs and keep PHAs in the solid state during the isolation

    process. In this regard, cell disruptions are necessary and can be performed in three

    ways: chemical digestion, enzyme disruption and mechanical disruption.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    44/154

    31

    Chemical agents, such as NaOH, hypochlorite and surfactants, can destroy the cells

    and dissolve cell components into the supernatant, followed by centrifugation or

    filtration to collect PHA pellets112

    . However, this method is a non-selective process and

    polymer degradation occurs simultaneously. Compared to solvent extraction, the purity

    of PHAs isolated by chemical digestion is lower and varies from 68% to 98% based on

    the digestion conditions and the types of microorganisms109, 111

    . A combination of

    heating, H2SO4, NaOH and sodium hypochlorite at proper concentrations and duration

    could improve the quality of PHAs with purity higher than 97% and a recovery yield

    higher than 95%113

    .

    Enzymatic disruption, employing proteolytic enzymes, will break cells by hydrolysis

    of proteins through cleavage of peptide bonds. The proteases help to lyse the cells

    efficiently at 50 C and pH 9.0 and release PHAs from the cell at a purity of 88.8%.

    Supplemented with chloroform or sodium dodecyl sulphate-ethylenediaminetetraacetic

    acid (SDS-EDTA), higher purity was achieved with a recovery yield of 90%109, 111

    .

    Nevertheless, the high cost of enzymes and complexity of the recovery process

    outweigh advantages of this method.

    Mechanical disruption, using a bead mill, high-pressure homogenization (including

    microfluidization) or supercritical fluid (SCF), is favored mainly due to the elimination of

    harsh chemicals, which minimizes environmental effects. The recovery yield of this

    process can reach 89%111

    . Pretreatments using SDS or NaOH help to further purify PHAs.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    45/154

    32

    The major drawbacks of this method are high capital investment cost and long

    processing time.

    Currently, a green, low-cost, highly efficient, and environmentally friendly PHA

    recovery process has not generally been accepted or implemented. Therefore, based on

    the characteristics and requirements for the end use of PHAs, a combination of several

    aforementioned recovery methods might be beneficial to reduce the production costs

    of PHAs.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    46/154

    33

    2. Materials and Methods

    2.1 Microorganism and fermentation conditions

    2.1.1 PHB production in shake flasks and fermentors

    Burkholderia cepacia ATCC 17759 was used in both shake flask as well as pilot

    scale (200 L) fermentations. The nitrogen-limited mineral salts medium used in the

    fermentations was initially described by Bertrand et al.114

    . The recipe for this medium

    was slightly changed in our experiments as follows: 1.5 g of (NH4)2SO4, 3.6 g of Na2HPO4 ,

    1.5 g of KH2PO4, 75.5 mg of CaCl2, 60 mg of NH4-Fe(III) citrate, 200 mg of MgSO4 7H2O,

    and 1 ml of trace elements solution per liter. The trace elements solution contained 100

    mg of ZnSO4 7H20, 30 mg of MnCl2 4H20, 300 mg of H3BO3, 200 mg of CoCl2 6H20, 20

    mg of CuSO4 5H20, 20 mg of NiCl2 6H20, 30 mg of NaMoO4 2H20 per liter.

    For shake-flask experiments, all cultures were shaken at 30C and 150 rpm.

    Glycerol (99.5%, EMD) and xylose (99% purity, Acros) were used to produce PHB for

    physical- chemical characterizations and were autoclaved separately in a solution of 50%

    (v/v) or 50% (w/v), respectively. All shake flask experiments were conducted in 500 mL

    baffled flasks containing 100 mL of medium with metal enclosures

    In experiments utilizing both xylose and glycerol as carbon sources, xylose (2.2%)

    was initially added into the medium. At 24 h or 48 h, 2% or 5% glycerol was added to the

    medium, and cultures were harvested at 72 h. Only 2.2% xylose was used as a carbon

    source for a control. When using the 400 L fermentor (Model No: IF 400, New Brunswick

    Scientific Co. Inc., NewBrunswick, NJ), a fed-batch method was used and glycerol (85%

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    47/154

    34

    purity) from a biodiesel-producing facility (Twin Rivers Technologies, Quincy, Mass.) was

    added as the primary carbon source for production of the homopolymer PHB.

    Tall oil fatty acids (MWV L-1A, MeadWestvaco Corporation, Richmond, VA) were

    autoclaved as received. Two milliliters of tall oil fatty acids were added as a carbon

    source into the 100 ml medium for bacterial growth and PHA production.

    Lactose (>99%, Sigma-Aldrich) was prepared as a stock solution at 20% (w/v) and

    sterilized separately. Cheese whey permeate (15%-20%, w/v) from Crowley Foods, a

    local subsidiary of HP Hood LLC in Arkport, NY, was adjusted to pH 7.0 and filtered by

    Whatman paper to remove the precipitate, followed by filter sterilization using Nalgene

    disposable filter units with 0.22 m PES membrane (ThermoFisher Scientific Inc.).

    The purities of other biodiesel-glycerol sources (from FutureFuel Corporation,

    Clayton, MO and SUNY-ESF biodiesel-producing facility, Syracuse, NY) were 93% and

    40%, respectively. The ESF crude glycerol was adjusted to neutral pH before sterilization.

    All sources of biodiesel-derived glycerol were autoclaved at 121 C for 20 min before use

    in fermentation experiments.

    2.1.2 PHB-co-HV production in shake flasks and fermentors

    PHB-co-HV copolymers were produced by B. cepacia ATCC 17759 in shake flasks, a

    7 L fermentor or a 400 L fermentor. The mineral salts-trace elements medium described

    previously was employed in these experiments, except for further addition of

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    48/154

    35

    ammonium sulfate to 10 g/L in the fermentor experiments. Reagent-grade glycerol

    (99.5%, EMD, Gibbstown, NJ) and levulinic acid (98%, Alfa Aesar, Ward Hill, MA) were

    used as carbon sources for bacterial growth and polymer production. Glycerol and

    levulinic acid were dissolved into ddH2O for a stock solution of 50% (v/v) and 50% (w/v),

    respectively. The stock solution of levulinic acid was adjusted to pH 7.0 before

    autoclaving. In shake flask experiments, Fernbach flasks (2800 ml in volume) were

    employed for copolymer production. Each Fernbach flask was prepared with 500 ml

    mineral salts medium and 3% (v/v) glycerol. Levulinic acid at a concentration of 0.1%

    (w/v) was added, except for the control group (only containing glycerol as a carbon

    source), when inoculated with a 5% (v/v) seed culture. At 24 h, the remaining levulinic

    acid at concentrations of 0.2%, 0.4%, 0.6% or 0.8% were added into the medium, to

    achieve final concentrations of levulinic acid in the medium of 0.3%, 0.5%, 0.7% and

    0.9%, respectively, along with the control group without levulinic acid and the group

    with only 0.1% levulinic aicd. All Fernbach experiments were incubated at 30 C and

    shaken at 150 rpm for 72 h.

    Copolymers of PHB-co-HV, with a ratio of HV between 5% and 32.6%, were

    produced by B. cepacia in a 7-L fermentor (BIOFLO410, New Brunswick Scientific Co.,

    Edison, NJ) with a working volume of 5 L. The concentration of glycerol was kept

    between 1% and 3% (v/v) during the fermentation and pure levulinic acid was

    continuously pumped into the fermentor at the rate of 0.5 g/Lh. Dissolved oxygen (DO)

    and pH were maintained at 40% and 7, respectively, by adjusting agitation/aeration and

    pumping 10 M sodium hydroxide. For this study, two stages of the fermentation were

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    49/154

    36

    employed: 1) growth ofB. cepacia to as high an OD540 (optical density) as possible with

    sufficient glycerol; 2) initiating levulinic acid addition at the point of highest OD540 for HV

    production and providing additional glycerol for bacterial growth and PHB production.

    Scaling up to pilot-plant level, copolymers were produced in a 400 L fermentor

    (Model No: IF 400, New Brunswick Scientific Co.,Inc. NewBrunswick, NJ), using a fed-

    batch method and glycerol (85% purity), from a biodiesel-producing facility (Twin Rivers

    Technologies, Quincy, MA), was added as the primary carbon source and levulinic aicd

    ( 97% purity, Sigma-Aldrich) was added periodically at concentrations of approximately

    1% (v/v) whenever the concentration of levulinic acisd was determined by HPLC to be

    less than 0.2%.

    2.1.3 P3HB-co-4HB production in shake flasks and

    fermentors

    P3HB-co-4HB copolymers were produced by B. cepacia ATCC 17759 in either shake

    flasks or a 7-L fermentor. In shake flask experiments, 0.1%, 0.3%, 0.5%, 0.7% and 0.9%

    (v/v) of 1,4-butanediol (99% ReagentPlus, Sigma-Aldrich) or -butyrolactone ( 99%

    ReagentPlus, Sigma-Aldrich) was co-fed with 3% (w/v) xylose or 3% (v/v) glycerol as

    major carbon sources in mineral-salts medium, described previously in 2.1.1. After

    inoculation with a 5% (v/v) seed culture shaken for 48 h, all flasks were shaken at 30 C

    and 150 rpm for 72 h. In a 7 L fermentor, Luria-Broth (LB) medium, containing 10 g/L

    BactoTM

    tryptone (BD, Sparks, MD), 5 g/L yeast extract (Sensient Technologies

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    50/154

    37

    Corporation, Junean, WI) and 5 g/L sodium chloride (EM Science, Gibbstown, NJ), was

    used for bacterial growth, supplemented with 1,4-butanediol or -butyrolactone at a

    total concentration of 2% (v/v) for 72 h at 30 C with 30% dissolved oxygen.

    Cell harvest and polymer isolation followed the procedure described in section 2.3.

    2.2 Concentration determination of glycerol, xylose, levulinic

    acid and lactose

    Glycerol concentration was measured using the Free Glycerol Reagent kit (Catalog

    No. F6428, Sigma, St,Louis, MO), basically containing glycerol kinase, glycerol phosphate

    oxidase, peroxidase, ATP, 4-aminoantipyrine (4-AAP) and sodium N-ethyl-N-(3-

    sulfopropyl) m-anisidine (ESPA). The reactions were incubated for 5 min at 37C and the

    absorbances were recorded spectrophotometrically at 540 nm (CARY 300

    Spectrophotometer, Varian Inc. USA) as described in details per the instructions of the

    manufacturer (http://www.sigmaaldrich.com/etc/medialib/docs/Sigma/Bulletin/f6428

    bul.Par.0001.File.tmp/f6428bul.pdf). Briefly, the Free Glycerol Reagent was pipetted at a

    volume of 0.8 ml into each cuvet. Water, Glycerol Standard (Catalog No. G7793, Sigma),

    and sample were added at a volume of 10 l to each cuvet labeled as blank, standard,

    and sample, respectively. The solution in each cuvet was mixed by gentle shake and all

    samples tested were incubated at 37 C for 5 minutes, and the absorbances at 540 nm

    were recorded of Blank, Standard, and Sample versus water as reference.

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    51/154

    38

    Note: the glycerol standard solution concentration is 0.26 mg glycerol/ml, with an

    equivalent triolein concentration of 2.5 mg/ml.

    Sugars, including xylose and lactose, and organic acids, including levulinic acid,

    were analyzed by HPLC (Waters Corporation, Milford, MA) equipped with a Waters 717

    plus autosampler, a 2996 photodiode array detector, a 2414 refractive index detector

    and a 1525 binary HPLC pump. For sugar analysis, a carbohydrate analysis column

    (WAT084038, 3.9 x 300 mm, Waters, Milford, MA) was employed with a mobile phase

    containing 20% Milli-Q water and 80% acetonitrile which had been filtered through a

    0.22 m membrane before use. A Jordi Gel DVB organic acid column (Cat. 17001, 250

    mm x 10 mm, The Nest Group, Inc., Southborough, MA) was used to determine the

    concentration of levulinic acid. The mobile phase (containing 0.02 M phosphoric

    acid/acetonitrile/methanol at a ratio of 90/5/5) was run at 1 ml/min and the spectrum

    was recorded at 214 nm. Empower Pro software (Waters, Milford, MA) was used to

    collect and analyze data.

    2.3 Isolation and purification of PHAs from biomass

    Cultures were harvested from shake flasks by centrifugation at 7000 x g for 10 min

    (Sorvall model SS-3). The biomass was subsequently washed with distilled H2O, if

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    52/154

    39

    necessary, using additional methanol (10 ml CH3OH per 100 ml culture) for removal of

    residual long chain fatty acids, and centrifuged again to remove the supernatant and

    stored at -20C until lyophilization. The biomass could be washed up to three times for

    further removal of residual glycerol, which could interfere with biomass drying after

    lyophilization. For cultures grown in the 400 L fermentor, the broth was centrifuged at

    16,000 rpm using a continuous flow centrifuge (CEPA High Speed Centrifuge Z81 G, New

    Brunswick Scientific, USA).

    The harvested biomass was then frozen and lyophilized at -80C and 200 millitorr

    for 24 h. The dry biomass was subsequently ground to a powder with a mortar and

    pestle, mixed with chloroform (10 ml chloroform / 1 g dry biomass), and either stirred at

    room temperature for 24 h for the biomass from the 400 L fermentor or placed in an

    incubator at 55-60 C overnight for the biomass from shake flask experiments and 7 L

    fermentor runs.

    For the purification of bench-scale quantities of PHAs, the general procedure of

    Keenan et al.115

    was used for all biomass samples. In general, after PHA extraction, the

    mixture of biomass and chloroform was filtered through Whatman #1 filter paper

    seated in a porcelain Buchner funnel apparatus to remove cell debris. Subsequently, the

    PHA-containing filtrate solution was decanted into cold methanol or ethanol (at 4 C) by

    a ratio of 1: 10 (v/v) during stirring to precipitate PHAs. The powder-like PHAs were

    collected by filtration through a Pyrex glass Buchner funnel with a fritted disc (fine-

    grade porosity with a maximum pore size of 4-5.5 m).

  • 8/2/2019 Production and Characterization of Polyhydroxyalkanoates (PHAs) From Burkholderia Cepacia ATCC 17759 Grown o

    53/154

    40

    For the extraction and purification of polymers from cells produced at the pilot-

    plant scale, lyophilized cell mass (1 kg) was stirred with 9 L of chloroform in a 10-L Pyrex

    glass container at room temperature for 24 h. This mixture was filtered as described

    previously for bench-scale PHA extraction. The polymer solution was concentrated by

    rotary evaporation using a Flash Evaporator (Buchler Instruments, Fort Lee, NJ) at 70C

    to a highly viscous solution and a distillation system (Buchler Instruments, Fort Lee, NJ)

    was used to recover chloroform. The polymer was precipitated from this solution via

    slow addition to cold methanol (1 volume of PHA solution / 10 volumes of methanol).

    Precipitated polymer samples exhibited a fibrous, noodle-shaped morphology and, if

    necessary, were washed with additional methanol for further purification. After the

    large noodle-shaped PHA polymers were screened and removed directly by a wire sieve,

    the remaining floc of PHAs was filtered through a Pyrex glass Buchner funnel with a

    fritted disc (coarse-grade porosity with a pore size of 40-60 m).

    All purified PHA samples were air dried in glass petri dishes in a hood overnight,

    followed by vacuum drying in a desiccator until use.

    2.4 Sample preparation with nucleating agents

    The purified homopolymer PHB and various copolymers of PHB-co-HV (between

    0.5 g and 1 g) were di