Polymers

您所在的位置:网站首页 halomonas boliviensis Polymers

Polymers

#Polymers | 来源: 网络整理| 查看: 265

Next Article in Journal Local Delivery of Antiproliferative Agents via Stents Next Article in Special Issue Maintaining Structural Stability of Poly(lactic acid): Effects of Multifunctional Epoxy based Reactive Oligomers Previous Article in Journal All Green Composites from Fully Renewable Biopolymers: Chitosan-Starch Reinforced with Keratin from Feathers Previous Article in Special Issue Biobased Contents of Natural Rubber Model Compound and Its Separated Constituents Journals Active Journals Find a Journal Proceedings Series Topics Information For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers Open Access Policy Institutional Open Access Program Special Issues Guidelines Editorial Process Research and Publication Ethics Article Processing Charges Awards Testimonials Author Services Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series About Overview Contact Careers News Blog Sign In / Sign Up Notice clear Notice

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

Continue Cancel clear

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess.

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Journals Active Journals Find a Journal Proceedings Series Topics Information For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers Open Access Policy Institutional Open Access Program Special Issues Guidelines Editorial Process Research and Publication Ethics Article Processing Charges Awards Testimonials Author Services Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series About Overview Contact Careers News Blog Sign In / Sign Up Submit     Journals Polymers Volume 6 Issue 3 10.3390/polym6030706 polymers-logo Submit to this Journal Review for this Journal Edit a Special Issue ► ▼ Article Menu Article Menu Subscribe SciFeed Recommended Articles Related Info Link Google Scholar More by Authors Links on DOAJ Tan, G. Amy Chen, C. Li, L. Ge, L. Wang, L. Razaad, I. Mutiara Ningtyas Li, Y. Zhao, L. Mo, Y. Wang, J. on Google Scholar Tan, G. Amy Chen, C. Li, L. Ge, L. Wang, L. Razaad, I. Mutiara Ningtyas Li, Y. Zhao, L. Mo, Y. Wang, J. on PubMed Tan, G. Amy Chen, C. Li, L. Ge, L. Wang, L. Razaad, I. Mutiara Ningtyas Li, Y. Zhao, L. Mo, Y. Wang, J. /ajax/scifeed/subscribe Article Views Citations - Table of Contents Altmetric share Share announcement Help format_quote Cite question_answer Discuss in SciProfiles thumb_up ... Endorse textsms ... Comment Need Help? Support

Find support for a specific problem in the support section of our website.

Get Support Feedback

Please let us know what you think of our products and services.

Give Feedback Information

Visit our dedicated information section to learn more about MDPI.

Get Information clear JSmol Viewer clear first_page settings Order Article Reprints Font Type: Arial Georgia Verdana Font Size: Aa Aa Aa Line Spacing:    Column Width:    Background: Open AccessReview Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review by Giin-Yu Amy Tan 1,2, Chia-Lung Chen 1,*, Ling Li 1, Liya Ge 1, Lin Wang 1, Indah Mutiara Ningtyas Razaad 1, Yanhong Li 2, Lei Zhao 1, Yu Mo 1,2 and Jing-Yuan Wang 1,2 1 Residues and Resource Reclamation Centre, Nanyang Environment and Water Research Institute, Nanyang Technological University, 1 Cleantech Loop, 637141, Singapore 2 Division of Environmental and Water Resources, School of Civil and Environmental Engineering, Nanyang Technological University, 50 Nanyang Avenue, 639798, Singapore * Author to whom correspondence should be addressed. Polymers 2014, 6(3), 706-754; https://doi.org/10.3390/polym6030706 Received: 30 January 2014 / Revised: 21 February 2014 / Accepted: 27 February 2014 / Published: 12 March 2014 (This article belongs to the Special Issue Polymers from Biomass) Download Download PDF Download PDF with Cover Download XML Download Epub Browse Figures Versions Notes

Abstract: With the impending fossil fuel crisis, the search for and development of alternative chemical/material substitutes is pivotal in reducing mankind’s dependency on fossil resources. One of the potential substitute candidates is polyhydroxyalkanoate (PHA). PHA is a carbon-neutral and valuable polymer that could be produced from many renewable carbon sources by microorganisms, making it a sustainable and environmental-friendly material. At present, PHA is not cost competitive compared to fossil-derived products. Encouraging and intensifying research work on PHA is anticipated to enhance its economic viability in the future. The development of various biomolecular and chemical techniques for PHA analysis has led to the identification of many PHA-producing microbial strains, some of which are deposited in culture collections. Research work on PHA could be rapidly initiated with these ready-to-use techniques and microbial strains. This review aims to facilitate the start-up of PHA research by providing a summary of commercially available PHA-accumulating microbial cultures, PHA biosynthetic pathways, and methods for PHA detection, extraction and analysis. Keywords: polyhydroxyalkanoate; PHA; biopolymer; bacteria; archaea; pathway; PHA detection; PHA extraction; PHA characterization 1. IntroductionPHA is a family of naturally-occurring biopolyesters synthesized by various microorganisms. First discovered by Lemogine in 1926 [1], PHA has since attracted much commercial and research interests due to its biodegradability, biocompatibility, chemical-diversity, and its manufacture from renewable carbon resources [2]. A PHA molecule is typically made up of 600 to 35,000 (R)-hydroxy fatty acid monomer units [3]. Each monomer unit harbors a side chain R group which is usually a saturated alkyl group (Figure 1) but can also take the form of unsaturated alkyl groups, branched alkyl groups, and substituted alkyl groups although these forms are less common [4]. Depending on the total number of carbon atoms within a PHA monomer, PHA can be classified as either short-chain length PHA (scl-PHA; 3 to 5 carbon atoms), medium-chain length PHA (mcl-PHA; 6 to 14 carbon atoms), or long-chain length PHA (lcl-PHA; 15 or more carbon atoms) [3]. About 150 different PHA monomers have been identified and this number continues to increase with the introduction of new types of PHA through the chemical or physical modification of naturally-occurring PHA [5], or through the creation of genetically-modified organisms (GMOs) to produce PHA with specialized functional groups [6]. These features gave rise to diverse PHA properties which can be tailored for various applications ranging from biodegradable packaging materials to medical products. PHA is also considered as pharmaceutically-active compound and currently investigated as potential anti-HIV drugs, anti-cancer drugs, antibiotics, etc. [7,8]. The production of various types of PHA material, their properties and downstream applications was recently reviewed by Philip et al. [9], Olivera et al. [10], Chen [11], and Rai et al. [7].The intense research and commercial interest in PHA is evident from the rapid increment in PHA-related publications. Web of Science citation report (Thomson Reuters, New York, NY, USA) revealed that in the last 20 years, PHA-related documents have increased by almost 10-fold while citations have increased by more than 500-fold with an average citation count of about 1100 citations per year. This has fuelled the growth of knowledge and development of techniques related to microbial PHA production. With this ready information, research work on PHA could be rapidly initiated either through using microbial strains previously deposited in culture collections or through isolating and characterizing novel PHA-producing microbes. Figure 2 is an illustration of the typical workflow processes for PHA research. This review aims to facilitate the start-up of PHA research by providing an overview of PHA-accumulating microbes currently available in culture collections, PHA biosynthetic pathways, techniques for microbial PHA detection and characterization, PHA polymer extraction, and polymer characterization. Polymers 06 00706 g001 1024 Figure 1. Polyhydroxyalkanoate (PHA) chemical structure. The nonmenclature and carbon number for PHA compounds is determined by the functional alkyl R group. Asterisk denotes chiral center of PHA-building block. Figure 1. Polyhydroxyalkanoate (PHA) chemical structure. The nonmenclature and carbon number for PHA compounds is determined by the functional alkyl R group. Asterisk denotes chiral center of PHA-building block. Polymers 06 00706 g001 Polymers 06 00706 g002 1024 Figure 2. Schematic workflow processes for PHA research. Figure 2. Schematic workflow processes for PHA research. Polymers 06 00706 g002 2. PHA Biosynthetic PathwaysPHA plays a pivotal role in priming microorganisms for stress survival. PHA promotes the long-term survival of bacteria under nutrients-scarce conditions by acting as carbon and energy reserves for both non-sporulating and sporulating bacteria. Additionally, bacteria that harbor PHA showed enhanced stress tolerance against transient environmental assaults such as ultraviolet (UV) irradiation, heat and osmotic shock [12]. PHA biosynthetic pathways are intricately linked with the bacterium’s central metabolic pathways including glycolysis, Krebs Cycle, β-oxidation, de novo fatty acids synthesis, amino acid catabolism, Calvin Cycle, and serine pathway [4,13,14,15,16,17]. Many common intermediates are also shared between PHA and these metabolic pathways, most notably being acetyl-CoA. In some PHA-producing microbes such as Cupriavidus necator, Chromatium vinosum, and Pseudomonas aeruginosa, the metabolic flux from acetyl-CoA to PHA is greatly-dependent on nutrient conditions [18]. Under nutrient-rich conditions, the production of high amounts of coenzyme A from Krebs Cycle blocks PHA synthesis by inhibiting 3-ketothiolase (PhaA) such that acetyl-CoA is channeled into the Krebs Cycle for energy production and cell growth [19] (Figure 3). Conversely, under unbalanced nutrient conditions (i.e., when an essential nutrient such as nitrogen and phosphorus is limiting in the presence of excess carbon), coenzyme A levels are non-inhibitory allowing acetyl-CoA to be directed towards PHA synthetic pathways for PHA accumulation [19,20]. This metabolic regulation strategy in turn enables the PHA-accumulating microbes to maximize nutrient resources in their adaptation to environmental conditions.To date, much insight has been gained on metabolic pathways for scl-PHA and mcl-PHA synthesis through studies using wild-type strains and heterologous expressions in recombinant strains [3]. In-depth reviews on these various PHA biosynthesis pathways and the enzymes involved have been provided by Chen [11], Lu et al. [4], Madison and Huisman [13], and Khosravi-Darani et al. [21]. Figure 3 shows the various routes of scl-PHA synthesis (pathways A to J) and mcl-PHA synthesis (pathways J to M) while Table 1 provides a summary of the enzymes involved. Although the biosynthesis of PHA from (R)-hydroxyalkyl-CoA ([R]-3-HA-CoA) precursors were most commonly reported, the diversity of PHA precursors is not restricted to (R)-3-HA-CoA alone [4]. Putative metabolic routes, such as pathways L and M (Figure 3), were recently proposed to expound for the metabolism of cyclohexanol to 6-hydroxyhexanoyl-CoA and 4,5-alkanolactone to 4,5-hydroxyacyl-CoA (4,5-HA-CoA) [11]. Nevertheless, the current knowledge on biosynthetic pathways is largely confined to (R)-3-HA-CoA precursors and falls short of accounting for the chemically-diverse PHA monomers and PHA monomers of lcl-PHA. There remains much about biosynthetic pathways waiting to be uncovered. Further studies to verify putative pathways as well as unraveling new biosynthetic pathways are anticipated to facilitate the creation of PHA materials that could be tailored for specific application needs. Polymers 06 00706 g003 1024 Figure 3. PHA biosynthetic pathways. Dotted lines represent putative pathways. Numbers represent enzymes involved in the chemical reactions that are summarized in Table 1. (3-H2MB-CoA, 3-hydroxy-2-methylbutyryl-CoA; 3-H4MV-CoA, 3-hydroxy-4-methylvaleryl-CoA; 3-HV-CoA, 3-hydroxyvaleryl-CoA; 3-H2MV-CoA, 3-hydroxy-2-methylvaleryl-CoA; 4-HB-CoA, 4-hydroxybutyryl-CoA; 3-HB-CoA, 3-hydroxybutyryl-CoA; (R)-3-HA-CoA, (R)-3-hydroxyacyl-CoA; 4,5-HA-CoA, 4,5-hydroxyacyl-CoA). Figure 3. PHA biosynthetic pathways. Dotted lines represent putative pathways. Numbers represent enzymes involved in the chemical reactions that are summarized in Table 1. (3-H2MB-CoA, 3-hydroxy-2-methylbutyryl-CoA; 3-H4MV-CoA, 3-hydroxy-4-methylvaleryl-CoA; 3-HV-CoA, 3-hydroxyvaleryl-CoA; 3-H2MV-CoA, 3-hydroxy-2-methylvaleryl-CoA; 4-HB-CoA, 4-hydroxybutyryl-CoA; 3-HB-CoA, 3-hydroxybutyryl-CoA; (R)-3-HA-CoA, (R)-3-hydroxyacyl-CoA; 4,5-HA-CoA, 4,5-hydroxyacyl-CoA). Polymers 06 00706 g003 Table Table 1. Enzymes involved in PHA biosynthesis pathways. Table 1. Enzymes involved in PHA biosynthesis pathways. No.EnzymeAbbreviationSpeciesReference1Glyceraldehyde-3-phosphate dehydrogenase-Cupriavidus necator[22]2Pyruvate dehydrogenase complex-Cupriavidus necator and Burkholderia cepacia[22]33-KetothiolasePhaACupriavidus necator[23]4NADPH-dependent acetoacetyl-CoA reductasePhaBCupriavidus necator[23]5PHA synthasePhaCCupriavidus necator and various[12,23]6Acetyl-CoA carboxylaseACCEscherichia coli K-12 MG1655[24]7Malonyl-CoA:ACP transacylaseFabDEscherichia coli K-12 MG1655[24]83-Ketoacyl carrier protein synthaseFabHEscherichia coli K-12 MG1655[24,25]9NADPH-dependent 3-Ketoacyl reductaseFabGPseudomonas aeruginosa[26]10Succinic semialdehyde dehydrogenaseSucDClostridium kluyveri[27]114-Hydroxybutyrate dehydrogenase4HbDClostridium kluyveri[27]124-Hydroxybutyrate-CoA:CoA transferaseOrfZClostridium kluyveri[27]13Alcohol dehydrogenase, putative-Aeromonas hydrophila 4AK4[28]14Hydroxyacyl-CoA synthase, putative-Mutants and recombinants of Cupriavidus necator[29]15Methylmalonyl-CoA mutaseSbmEscherichia coli W3110[30]16Methylmalonyl-CoA racemase-Nocardia corallina[31]17Methylmalonyl-CoA decarboxylaseYgfGEscherichia coli W3110[30]18Ketothiolase, putative--[32]193-KetothiolaseBktBCupriavidus necator[33]20Ketothiolase, putative--[32]21NADPH-dependent acetoacetyl-CoA reductase-Rhizobium (Cicer) sp. CC 1192[34]22Acyl-CoA synthetaseFadDPseudomonas putida CA-3 and Escherichia coli MG1655[35,36]23Acyl-CoA oxidase, putative--[37]24Enoyl-CoA hydratase I, putative--[37]25(R)-Enoyl-CoA hydratasePhaJPseudomonas putida KT2440[38]26Epidermase--[37]273-Ketoacyl-CoA thiolaseFadAPseudomonas putida KT2442[39]283-Hydroxyacyl-ACP:CoA transacylasePhaGPseudomonas mendocina[40]29Cyclohexanol dehydrogenaseChnAAcinetobacter sp. SE19 and Brevibacterium epidermidis HCU[41]30Cyclohexanone monooxygenasesChnBAcinetobacter sp. SE19 and Brevibacterium epidermidis HCU[41]31Caprolactone hydrolaseChnCAcinetobacter sp. SE19 and Brevibacterium epidermidis HCU[41]326-Hydroxyhexanoate dehydrogenaseChnDAcinetobacter sp. SE19 and Brevibacterium epidermidis HCU[41]336-Oxohexanoate dehydrogenaseChnEAcinetobacter sp. SE19 and Brevibacterium epidermidis HCU[41]34Semialdehyde dehydrogenase, putative--[11]356-hydroxyhexanoate dehydrogenase, putative--[11]36Hydroxyacyl-CoA synthase, putative--[11]37Lactonase, putative-Mutants and recombinants of Cupriavidus necator[29] 3. PHA-Producing Microbial Strains from Culture CollectionsThe PHA bioaccumulation trait is widespread among the bacterial and archaeal domains with PHA-producing microbes occurring in more than 70 bacterial and archaeal genera [4,42]. Bioaccumulated PHA is stored in the form of intracellular lipid granules in these microbes [43]. Acting as biocatalysts, these PHA-producing microorganisms enable the coupling of a myriad of carbon catabolic pathways together with PHA anabolic pathways, thereby playing a key role in the diversification of PHA production from various carbon sources. These carbon sources include saccharides (e.g., fructose, maltose, lactose, xylose, arabinose, etc.), n-alkanes (e.g., hexane, octane, dodecane, etc.), n-alkanoic acids (e.g., acetic acid, propionoic acid, butyric acids, valeric acid, lauric acid, oleic acid, etc.), n-alcohols (e.g., methanol, ethanol, octanol, glycerol, etc.), and gases (e.g., methane and carbon dioxide) [1,44]. Wastestreams, which provide a free source of carbons, have also been identified for PHA production [45]. These include waste frying oil, vinegar waste, waste fats, food waste, agricultural waste, domestic wastewater, plant oil mill effluents, crude glycerol from biodiesel production, plastic waste, landfill gas, etc. The deposition of some PHA-producing microbial strains in culture collections has made these strains commercially available. Microbial strains from culture collections are generally well-documented in terms of their genetics and biochemistry underlying carbon assimilation and PHA accumulation. With this knowledge, it enables the appropriate microbes to be selected according to the targeted carbon source, facilitating the rapid start-up of PHA-related research and/or industrial production. Table 2 provides a summary of carbon substrate utilization and PHA production by deposited bacterial and archaeal strains. Table Table 2. PHA-producing microbial strains available in culture collections. Table 2. PHA-producing microbial strains available in culture collections. MicroorganismCulture collection number bCarbon sourcePHA monomer or polymer cPHA content (%CDM)Average PHA productivity(g L−1 h−1)ReferenceGram-negative bacteria Azohydromonas australica(formerly Alcaligenes latus)ATCC 29713,DSM 1124, IAM 12664,LMG 3324Malt wasteP3HB70.10.445[46]Azohydromonas lata(formerly Alcaligenes latus)ATCC 29714,DSM 1123,IAM 12665,LMG 3325SucroseP3HB50.0–88.00.050–4.940[47,48,49]Fructose, glucoseP3HB76.5–79.40.121–0.128[50]Azotobacter beijerinckiiDSM 1041,NCIB 11292GlucoseP3HB24.80.090[51]Burkholderia cepacia (formerly Pseudomonas multivorans and Pseudomonas cepacia)ATCC 17759,DSM 50181,NCIB 9085XyloseP3HB58.4NG[52]GlycerolP3HB31.30.103[53]Fructose, glucose, sucroseP3HB50.4–59.0NG[50]Burkholderia sp. USMJCM 15050Lauric acid, myristic acid, oleic acid, palmitic acid, stearic acidP3HB1.0–69.0NG[54]Caulobacter vibrioides (formerly Caulobacter crescentus)DSM 4727GlucoseP3HB18.30.008[55]Cupriavidus necator H16 (formerly Hydrogenomonas eutropha H16, Alcaligenes eutrophus H16, Ralstonia eutropha H16 and Wautersia eutropha H16)ATCC 17699, DSM 428,KCTC 22496, NCIB 10442Fructose, glucoseP3HB67.0–70.50.052–0.067[50]4-Hydroxyhexanoic acidP3HB76.3–78.5NG[56]Corn oil, oleic acid, olive oil, palm oilP3HB79.0–82.00.041–0.047[57]Acetate, butyrate, lactic acid, propionic acid3HB, 3HV3.9–40.70.001–0.037[58]CO2P3HB88.90.230[59]Cupriavidus necator (formerly Hydrogenomonas eutropha, Alcaligenes eutrophus N9A, Ralstonia eutropha N9A and Wautersia eutropha)DSM 5184-Hydroxyhexanoic acidP3HB65.8–66.2NG[56]Cupriavidus necator (formerly Hydrogenomonas eutropha, Alcaligenes eutrophus TF93, Ralstonia eutropha TF93 and Wautersia eutropha)ATCC 17697, DSM 5314-Hydroxyhexanoic acidP3HB67.2NG[56]CO2P3HB60.00.600[60]Cupriavidus necator a (formerly Hydrogenomonas eutropha, Alcaligenes eutrophus, Ralstonia eutropha and Wautersia eutropha)CECT 4623, KCTC 2649, NCIMB 11599GlucoseP3HB76.02.420[61]Potato starch, saccharified wasteP3HB46.01.470[62]Cupriavidus necator (formerly Hydrogenomonas eutropha, Alcaligenes eutrophus, Ralstonia eutropha and Wautersia eutropha)DSM 545MolassesP3HB31.0–44.00.080–0.120[63]Glucose, propionic acidP3HB3HV80.00.820[64]Waste glycerolP3HB14.8–36.10.330–4.200[65]Halomonas boliviensis LC1ATCC BAA-759, DSM 15516Hydrolyzed starchP3HB56.0NG[66]Hydrogenophaga pseudoflavaATCC 33668, DSM 1034Lactose, sucroseP3HB3HV20.2–62.50.018–0.117[67]Hydrolyzed whey and valerateP3HB3HV40.00.050[68]Methylobacterium extorquensATCC 55366MethanolP3HB40.0–46.00.250–0.600[69]Methylobacterium extorquensATCC 8457, DSM 1340, NCIB 2879, NCTC 2879MethanolP3HB35.0–62.30.183–0.980[70,71]Methylocystis sp. GB25 aDSM 7674MethaneP3HB51.0NG[72]Novosphingobium nitrogenifigens Y88DSM 19370, ICMP 16470GlucoseP3HB81.00.014–0.021[73]Paracoccus denitrificansATCC 17741, DSM 413n-PentanolP3HV22.0–24.0NG[74]Pseudomonas aeruginosaNCIM 2948Cane molasses, fructose, glucose, glycerol, sucroseP3HB12.4–62.00.012–0.110[75]Pseudomonas aeruginosa PAO1ATCC 47085Oil and wax products from polyethylene (PE) pyrolysismcl-PHA25.0NG[76]Pseudomonas frederiksbergensis GO23 aNCIMB 41539Terephthalic acid from polyethyleneterephthalate (PET) pyrolysismcl-PHA24.00.004[77]Pseudomonas marginalisDSM 502761,3-butanediol, octanoatescl-mcl-PHA, mcl-PHA11.9–31.4NG[78]Pseudomonas mendocinaATCC 25411, DSM 500171,3-butanediol, octanoatescl-mcl-PHA13.5–19.3NG[78]Pseudomonas oleovoransATCC 8062, DSM 10454-Hydroxyhexanoic acidscl-mcl-PHA18.6NG[56]Pseudomonas putida CA-3 aNCIMB 41162Styrenemcl-PHA31.80.063[79]Styrene from polystyrene (PS) pyrolysismcl-PHA36.40.033[80]Pseudomonas putida GO16 aNCIMB 41538Terephthalic acid from polyethyleneterephthalate (PET) pyrolysismcl-PHA27.0~0.005, 0.008 d[77]Pseudomonas putida GO19 aNCIMB 41537Terephthalic acid from polyethyleneterephthalate (PET) pyrolysismcl-PHA23.0~0.005, 0.008 d[77]Pseudomonas putida GPo1 (formerly Pseudomonas oleovorans)ATCC 29347Alkenes, n-alkanesmcl-PHA2.0–28.0NG[81]n-alkanoatesscl-mcl-PHA, mcl-PHA5.0–60.0NG[82,83]Pseudomonas putida KT2440ATCC 47054Nonanoic acidmcl-PHA26.8–75.40.250–1.110[84]4-Hydroxyhexanoic acidmcl-PHA25.3–29.8NG[56]Glucosemcl-PHA32.10.006[85]Pseudomonas putida F1ATCC 700007, DSM 6899Benzene, ethylbenzene, toluenemcl-PHA1.0–22.0NG[86]Pseudomonas putida mt-2NCIMB 10432Toluene, p-xylenemcl-PHA22.0–26.0NG[86]Acetic acid, citric acid, glucose, glycerol, octanoic acid, pentanoic acid, succinic acidmcl-PHA4.0–77.0NG[87]Thermus thermophilus HB8ATCC 27634, DSM 579Wheyscl-mcl-PHA35.60.024[88] Gram-Positive bacteria Bacillus megateriumDSM 90Citric acid, glucose, glycerol, succinic acidP3HB9.0–50.0NG[87]Bacillus megateriumCCM 1464,DSM 509,IFO 12109, NBRC 12109Citric acid, glucose, glycerol, succinic acid, octanoic acidP3HB, scl-mcl-PHA, mcl-PHA3.0–48.0NG[87]Various Bacillus spp. type strainsRefer to [89]Acetate, n-alkanoate,3-Hydroxybutyrate, propionate, sucrose, valerate3HB, 3HV, 3HHx2.2–47.6NG[89]Corynebacterium glutamicumATCC 15990, DSM 20137, NCIB 10337Acetic acid, citric acid, glucose, glycerol, succinic acidP3HB, mcl-PHA4.0–32.0NG[87]Corynebacterium hydrocarboxydansATCC 21767Acetate, glucose3HB, 3HV8.0–21.0NG[90]Microlunatus phosphovorusDSM 10555,JCM 9379Glucose3HB, 3HV20.0–30.0NG[91]Nocardia lucidaNCIMB 10980Acetate, succinate3HB, 3HV7.0–20.0NG[90]Rhodococcus sp. aNCIMB 40126Acetate, 2-alkenoate,1,4-butanediol,5-chlorovalerate, fructose, glucose, hexanoate,4-Hydroxybuytrate, lactate, molasses, succinate, valerateP3HB3HV4.0–53.0NG[90]Various Streptomyces spp. type cultureRefer to [89]GlucoseP3HB1.2–82.0NG[89]Archaea Haloferax mediterraneiATCC 33500, CCM 3361,DSM 1411VinasseP3HB3HV50.0–73.00.050–0.210[92]Hydrolyzed wheyP3HB3HV72.80.090[93]Glycerol and crude glycerol from biodiesel productionP3HB3HV75.0–76.00.120[94]Various archaeal strainsRefer to [95]Fructose, glucose, glycerolP3HB, P3HB3HV0.8–22.9 98.0% Recovery: 95.0%[146]Hypochlorite digestionSodium hypochloriteCupriavidus necator(DSM 545)Biomass concentration: 10-40 g/L; pH: 8-13.6; Temperature: 0-25 °C;Digestion time: 10 min-6 h; Hypochlorite concentration: 1%-10.5% weight/volume (w/v)Purity: 90.0%-98.0% Recovery: 90.0%-95.0%[147]Sodium hypochlorite and chloroformCupriavidus necator(NCIMB 11599) andrecombinant Escherichia coliBiomass concentration: 1% (w/v); Temperature: 30 °C; Digestion time: 1 h; Hypochlorite concentration: 3%-20% volume/volume (v/v)Purity: 86.0% Recovery: NGPurity: 93.0%Recovery: NG[153]Enzyme digestionTrysin, bromelain, pancreatinCupriavidus necator (DSM 545)Digestion with 2% trypsin (50 °C, pH 9.0, 1 h) or 2% bromelain (50 °C, pH 4.75, 10 h) or 2% pancreatin (50 °C, pH 8.0, 8 h), followed by centrifugation and washing with 0.85% saline solutionPurity: 87.7%-90.3% Recovery: NG[149] NG, not given. Table Table 5. Medium-chain length PHA (mcl-PHA) recovery methods. Table 5. Medium-chain length PHA (mcl-PHA) recovery methods. MethodChemicalSpeciesConditionsPurity and recoveryReferenceSolventChloroformPseudomonas oleovorans (strains NRRL B-14682, NRRL B-14683, and NRRL B-778)30 °C overnight at 250 rpmNG[150]ChloroformPseudomonas oleovorans (NRRL B-14683), Pseudomonas resinovorans (NRRL B-2649), Pseudomonas citronellolis (NRRL B-2504), and Pseudomonas putida KT2442Soxhlet extraction for 24 hNG[154,155]ChloroformPseudomonas putida IPT 046Soxhlet extraction for 6 hNG[156]ChloroformPseudomonas aeruginosa 42A2 (NCIMB 40045)100 °C for 3 h in screw cap tubes for small quantities or in a soxhlet apparatus for large amounts of cell materialNG[157]Dichlorome-thanePseudomonas oleovorans (ATCC 29347)Soxhlet extraction at 60 °C for 5 hPurity: > 98.0%Recovery: NG[151]AcetonePseudomonas putida KT2440 (ATCC 47054)22 °C for 24 h at 170 rpmPurity: 80.0%–90.0%Recovery: 60.0%–80.0%[152]Enzyme digestionAlcalase, SDS, EDTA, lysozymePseudomonas putidaDigestion with alcalase and SDS at pH 8.5 and 55 °C followed by further treatments with EDTA and lysozyme at pH 7 and 30 °CPurity: 92.6%Recovery: nearly 90.0%[158]Pseudomonas putida KT2442Digestion with excess alcalase, EDTA and SDS at pH 8.5 and 55 °C followed by diafiltrationPurity: > 95.0%Recovery: NG[159,160] NG, not given. Table Table 6. Techniques for PHA polymer characterization. Table 6. Techniques for PHA polymer characterization. CharacteristicIndexMethodSampleSample preparationTypical conditionsReferencePHA monomeric compositionChemical derivative of PHA monomersLCRefer to Table 3GCRefer to Table 3PHA polymeric compositionTopology and functional groups of PHA molecule1D-Nuclear magnetic resonance (NMR)5–10 mg PHA for 1H-NMR and 20– 30 mg PHA for 13C-NMRDissolution of PHA polymer in 0.7 mL deuterated chloroform (CDCl3) containing 0.03% (v/v) tetramethylsilane (TMS)1H-NMR at 200 or 300 MHz and 13C-NMR measurements at 75.4 MHz at 20 °C with a sampling pulse of 3 s. Chemical shifts were referenced to the residual proton peak of CDCl3 at 7.26 ppm and to the carbon peak of CDCl3 at 77 ppm[82]2D-NMR10 mg PHA for homonuclear 2D-NMR and 40–50 mg PHA for heteronuclear 2D-NMRRefer to above “1D-NMR”For homonuclear COSY and TOCSY, 16 scans were accumulated per increment over a spectral width of 7.8 ppm. For heteronuclear HSQC, 48 scans were accumulated per increment over a spectral width of 7.8 ppm for 1H and 75 ppm for 13C. For heteronuclear HMBC spectrum, 64 scans were acquired with the long-range coupling delay set for 8 Hz[161]PHA polymeric compositionTopology and functional groups of PHA moleculeMatrix assisted laser desorption ionization-time of flight-mass spectrometry (MALDI-TOF-MS)1 µg–1 mg PHAThe matrix used was either dithranol or dihydroxybenzoic acid (DHB) at a concentration of 10 mg mL−1 in THF. 1 mg mL−1 PHA solution (in chloroform) was mixed with equal volume THF. The matrix solution and the PHA solution were subsequently mixed in a 5:2 ratio (matrix/sample). 1 µL mixture was deposited onto the stainless steel sample holder. The solvent was allowed to air-dry before loading the sample plate into the MALDI ion sourceMALDI-TOF-MS with 25 kV acceleration and detection in the positive-ion high-resolution reflection mode[162]Molecular distributionPolydispersity, molecular mass and molecular mass distributionGel permeation chromatography (GPC)0.1–1 mg PHADissolution of PHA polymer in 1 mL of THFAnalysis conducted with a refractive index detector (47 °C, 2.0 bar) and a solvent-compatible GPC column. THF, containing 250 ppm of 2,6-di-tert-butyl-4-methylphenol (BHT) as inhibitor, was used as an eluent at a flow rate of 0.5 mL min−1 and 40 °C[43]Dissolution of PHA polymer in chloroformAnalysis conducted with a differential refractive index detector (30 °C), a UV dual wavelength absorbance detector, and a combination of four GPC columns series. Chloroform was used as an eluent with a flow rate of 1.0 mL min−1[163]MALDI-TOF-MSRefer to above “PHA polymeric composition” Thermal propertiesGlass transition temperature and melting temperatureDifferential scanning calorimetry (DSC)10 mg PHA-Heat sample from −100 °C– 400 °C at a heating rate of 10 °C min−1 under purified air or nitrogen gas with a flow rate of 80 mL min−1[43]Differential thermal analysis (DTA)5 mg PHA-Crystallization was carried out isothermally by abruptly quenching the samples from melt to the crystallization temperature, at which the samples were annealed for 10 min. Melting of semicrystalline samples was performed by heating at a rate of 5 °C min−1[164]Thermodegradation temperatureThermogravimetric analysis (TGA)10 mg PHA-Heat sample from room temperature to 700 °C at a heating rate of 10 °C min−1 under purified air or nitrogen gas with a flow rate of 50 mL min−1[43]CrystallinityMelting enthalpyDSCRefer to above “Thermal properties”Infrared absorption bands correlated to crystallinityFTIR5–10 mg PHADissolve PHA in chloroform, apply onto KRS-5 window and blow dry to evaporate solvent. Alternatively, mix PHA with potassium bromide (KBr) powder and pelletizeRefer to Table 3[127,165]Place PHA sample between two pieces of barium fluoride slidesMelt sample at 100 °C for 2 min in FTIR hot stage under the protection of dry nitrogen gas. Quench the amorphous sample to 58 and 28 °C by a flow of liquid nitrogen and maintain at these temperatures for 30 min for isothermal melt-crystallization before re-heating at 1 °C min−1[166]CrystallinityDiffraction intensity correlated to crystallinityX-ray diffractionDry polymer powder-Diffractogram of the sample powder were measured at room temperature by an imaging plate diffractometer with Cu-Kα radiation (wavelength = 0.1542 nm) as an incident X-ray source emitted by a X-ray generator with a Ni filter. The scattering angle range of 2θ = 10°–40° at a scan speed of 3° min−1[156]Mechanical propertiesTensile strength, tensile stress, percent elongation, modulus of elasticityMechanical testing machine of the constant-rate-of-crosshead-movement type with extensometer and micrometersPolymer thickness 1–14 mm, width 19–29 mm, length 165–246 mmTest samples were prepared using a hydraulic press at 150 °C and conditioned at a relative humidity of 50% ± 5% for 24 h prior to measurementsPerform stress-strain test at room temperature with a strain rate of 20 mm min−1[167] For mcl-PHA recovery through enzyme digestion, enzymes alcalase and lysozyme, together with sodium dodecyl sulfate (SDS) and ethylenediaminetetraacetic acid (EDTA) were used. This combination of enzymes and chemicals were successfully applied to mcl-PHA isolation for fed-batch fermentations of up to 200 L, where the cells were first ruptured by thermal treatment and the resultant debris was exposed to excess protease (alcalase), EDTA and SDS for solubilization. After cross-flow microfiltration, the final mcl-PHA latex had a purity exceeding 95%, demonstrating potential commercial applicability [159,160]. In another study [158], the PHA granules, present in water suspension after enzymatic treatment, were recovered by removing the solubilized non-PHA cell material through ultrafiltration system and purified through continuous diafiltration process. The final purity of PHA was 92.6% and recovery was nearly 90% [158]. While enzyme digestion is a more environmental-friendly approach than solvent extraction, the purity of polymer attained is lower. Biomedical application requires a final purity of 99% or more which is currently not achievable through enzymatic mcl-PHA recovery and a second purification using solvent extraction is necessary [160]. 6. Techniques for PHA Polymer CharacterizationPurified PHA polymers are diverse in their chemical composition and material properties due to the myriad of PHA monomeric units available as well as the incorporation of these monomers at varying amounts. To identify suitable downstream applications for PHA, characterization of the biomaterial is imperative. A summary of these techniques and their execution is provided in Table 6. 6.1. Monomeric Composition and DistributionThe monomeric composition and distribution of PHA polymer could be determined from GC, LC and nuclear magnetic resonance spectroscopy (NMR). For chromatography-based methods, the analysis of PHA polymer is similar to that for intracellular PHA, requiring depolymerization of the polymer, usually combined with chemical derivatization before it can be analyzed [43,130,132], which meant that chromatography methods cannot analyze PHA as an intact polymer. NMR on the other hand, could study the chemical makeup of an intact PHA polymer and differentiate between PHA blends and PHA copolymer though providing details on the topology and functional group in molecules [150,153,154,168]. Typically, two types of NMR techniques are available and they are 1H-NMR and 13C-NMR. The high proton abundance in nature meant that 1H-NMR is more sensitive and requires shorter analytical time (within one hour). In contrast, owing to low sensitivity and natural abundance of 13C, it may require longer analysis (up to 24 h) to accumulate enough signal intensity for 13C-NMR spectrum. Despite its shortcoming, 13C-NMR performs better than 1H-NMR at the analysis of macromolecule as well as long carbon chain of monomers. As PHA polymers contain hydrogen and carbon, 1H-NMR and 13C-NMR are usually applied in combination to provide a more comprehensive analysis of the polymer. NMR is broadly used for saturated and unsaturated PHA analysis. Functional groups such as methane protons, methylene protons, –CH=CH– can be identified from both NMR spectra while microstructures like 3-hydroxypropionate (3HP) and 4-hydroxybutyrate (4HB) can be obtained by analyzing both 1H-NMR and 13C-NMR spectra [156,169]. Quantitative estimation of PHA monomers can also be performed with NMR using the intensity ratio of the signals [156]. NMR is also a powerful non-destructive tool that could be applied to the analysis of novel functionalized PHAs for which analytical standards are currently unavailable [170,171]. Typical 1H-NMR and 13C-NMR spectra of mcl-PHA are shown in Figure 5 [172]. Polymers 06 00706 g005 1024 Figure 5. (A) Typical 1H-NMR spectrum of PHA. Protons in the polymers are numbered and assigned to the peaks in the spectrum; (B) Typical 13C-NMR spectrum of PHA. Carbon atoms in polymers are numbered and assigned to peaks in the spectrum. (Adapted from [172], with permission). (3HHx, 3-hydroxyhexanoate; 3HO, 3-hydroxyoctanoate; 3HD, 3-hydroxydecanoate; 3HDD, 3-hydroxydodecanoate). Figure 5. (A) Typical 1H-NMR spectrum of PHA. Protons in the polymers are numbered and assigned to the peaks in the spectrum; (B) Typical 13C-NMR spectrum of PHA. Carbon atoms in polymers are numbered and assigned to peaks in the spectrum. (Adapted from [172], with permission). (3HHx, 3-hydroxyhexanoate; 3HO, 3-hydroxyoctanoate; 3HD, 3-hydroxydecanoate; 3HDD, 3-hydroxydodecanoate). Polymers 06 00706 g005 In addition, two-dimensional (2D) NMR methodologies (such as correlation spectroscopy [COSY], nuclear Overhauser effect spectroscopy [NOESY], heteronuclear multiple-quantum correlation [HMQC] and heteronuclear multiple bond coherence [HMBC]) are very useful in the characterization of all kinds of specialized PHA, such as unsaturated, branched, halogenated or acetylated PHA [43,56,161,173,174]. 2D-NMR provides information about the environment where each carbon/hydrogen is positioned. With the aid of 2D homonuclear or heteronuclear NMR techniques, the exact position of double bonds and the cis/trans configuration of monomers can be determined [43,173]. 1H-NMR and 2D homonuclear NMR were used to unambiguously identify the position of the hydroxyl group and the positions of the double bonds in mcl-PHA [174]; while 2D heteronuclear COSY NMR was successfully applied to show that 4-hydroxyvaleric acid was a constituent of the PHA polymer [56]. Moreover, 2D heteronuclear HMBC NMR could also clearly reveal the presence of the block microstructure in PHA [175].Besides, it is notable that MS techniques, such as fast-atom bombardment (FAB)-MS [176], pyrolysis/MS [177] and matrix assisted laser desorption ionization-time of flight-mass spectrometry (MALDI-TOF-MS) [162,178,179] have been applied for the characterization of PHA composition. Among them, MALDI-TOF-MS offers a cost-effective, straightforward and high-throughput alternative to the well-established GC-MS methods [162,179]. Compared with GC-MS, MALDI-TOF-MS is a more straightforward method, and no chemical derivatization is required during sample preparation. As the differences in monomer composition and detailed PHA structural information can easily be identified by MALDI-TOF-MS, it can be used as complementary technique to NMR [178]. The total amount of sample deposited onto the target can be in the pico- to femtomole range, which makes MALDI-TOF-MS extremely sensitive with minimal sample requirement. The significant advantages provided by MALDI-TOF-MS technology therefore could facilitate routine analysis of PHA with high accuracy and precision. 6.2. Molecular Mass (Mw), Molecular Mass Distribution (Mn), and Polydispersity Index (PDI)A PHA polymer’s average molecular mass (Mw), molecular mass distribution (Mn), and polydispersity index (PDI; Mw/Mn) could be determined through a gel permeation chromatography (GPC) system, calibrated with polystyrene standards [150]. The Mw of PHA spans over a wide range from 50 kDa to 10,000 kDa and depending on Mn value, PDI could be between 1.1 and 6.0 [13,97,180]. GPC columns such as Styragel HMW 6E (5 to 10,000 kDa) [150] and PLgel MIXED-C (0.2 to 2,000 kDa) [181] have been applied for PHA analysis. Often, the broad Mw range of PHA makes the analysis of an unknown polymer more challenging particularly when the biomaterial is a mixture of several PHA molecules with vastly different Mw and Mn values. Two or more GPC columns, connected in series, are therefore necessary to fully reveal the polymer’s Mw and Mn [153].MALDI-TOF-MS, on the other hand, is the new and promising method for PHA characterization. It could potentially be used for evaluating the Mw and Mn of PHAs and their oligomers [162,178,179]. Unlike GPC that can only be used in determining apparent Mn values, MALDI-TOF-MS can offer accurate mass measurement of PHA [182]. Because of its high molecular resolution, excellent accuracy, reproducibility, and automation properties, MALDI-TOF-MS can make a significant contribution to the study of PHA. 6.3. Thermal PropertiesThermal properties such as glass transition temperature (Tg), melting temperature (Tm), and thermodegradation temperature (Td) are commonly examined for PHA material to determine the temperature conditions at which the polymer can be processed and utilized. The Tg, Tm, and Td values for PHA are usually in the range of −52 to 4 °C, non-observable to 177 °C, and 227 to 256 °C, respectively [7,180]. Information on Tg and Tm could be obtained using differential scanning calorimetry (DSC) and differential thermal analysis (DTA). The difference which sets them apart is that DTA is able to measure mass loss and qualitatively provide thermal information [183] while direct heat flow measurement enables DSC to provide not only qualitative results, but also quantitative thermal information, making DSC the preferred method in PHA studies [153,156]. The Td value of PHA is obtained using thermogravimetric analysis (TGA), a technique where a sample is heated in a controlled atmosphere at a defined rate while sample mass loss is measured [153,156]. The development of the simultaneous thermal analysis (STA) combines TGA and DSC/DTA measurement techniques, enabling Tg, Tm, and Td values determination on a single instrument, which provides a more productive and simpler means to analyze PHA [156]. 6.4. CrystallinityPHA polymers can range from non-crystalline to highly crystalline with crystallinity values between 0% and 70% [7,97]. Crystallinity could be measured by structural analysis instruments including FTIR, DSC and X-ray diffraction. In FTIR analysis, PHA displays characteristic infrared absorption bands at certain wavenumbers which can be correlated to crystallinity. The exact band locations vary according to the chemical composition of the polymer. For scl-PHAs such as P3HB and P3HB3HV, bands around 1279, 1228, and 1185 cm−1 are sensitive to the degree of crystallinity [184,185]. Band 1725 cm−1 and bands in the range of 1500 to 1300 cm−1, 1300 to 1000 cm−1 and 1000 to 800 cm−1 are revealing of the conformational changes of mcl-PHA and scl-mcl PHA in both the crystalline phase and amorphous phase [166]. In DSC analysis, melting enthalpy (∆Hm) provides an estimated value for heat of fusion (∆Hf) under the analysis conditions, which could be related to PHA’s crystallinity. PHA polymers with very low crystallinity typically have low to non-observable ∆Hf while highly-crystalline polymers such as P3HB can have ∆Hf values up to 146 J g−1 [7,186]. On their own, both FTIR and DSC are only adequate at measuring relative crystallinities within a given material. Measurement of the absolute crystallinity using FTIR and DSC can only be performed for PHA polymer with known crystallinity [187,188]. Absolute crystallinity could be obtained using methods based on X-ray diffraction. X-ray diffraction analysis is able to shed light on the polymer’s rate of crystallinity, as well as atomic structures such as chemical bonds, their disorder [189]. Crystallinity percentage can be calculated according to semi-crystalline and amorphous polymer areas in the diffractogram using Lorentzian and Gauss functions [156]. 6.5. Mechanical PropertiesYoung’s modulus, elongation at break and tensile strength are mechanical properties commonly evaluated for PHA polymers. The Young’s modulus provides a measure of PHA’s stiffness and ranges from the very ductile mcl-PHA (0.008 MPa) to the stiffer scl-PHA (3.5 × 103 MPa) [7]. Elongation at break measures the extent that a material will stretch before it breaks and is expressed as a percentage of the material’s original length. PHA polymers can take the form of a hard rigid material or a soft elastomeric material, displaying a wide elongation at break values of between 2% and 1000% [180]. Tensile strength measures the amount of force required to pull a material until it breaks, and is typically in the range of 8.8 to 104 MPa for PHA polymers [7]. Measurement of the aforementioned mechanical properties can be performed with tensile tester instrument by standardized test methods such as the ASTM standards [167]. 7. ConclusionsThis review paper provided a summary of PHA biosynthetic pathways, PHA-producing microbial strains commercially available from culture collections and their application, as well as techniques for PHA analysis and polymer extraction. It is evident that there are many avenues through which PHA could be produced depending on the type of microorganisms employed, choice of carbon source, and cultivation conditions. These aforementioned factors also influenced the type of PHA produced, which in turn determines their downstream applications. Using this wealth of knowledge, future development in commercial PHA production could adopt a more “top-down” approach where the targeted carbon source and desired PHA product are decided a priori together with economic considerations before the appropriate microorganism or group of microoganisms is selected for the purpose as a means to achieve economic viability for the bioprocess. The formulation of microbial co-cultures for PHA production is largely considered as unexplored territories but may have the potential to produce PHA cheaply from organic waste streams. A fast-growing area in PHA research is the biosynthesis of tailored PHA for specific application needs. Existing microbial strains from culture collections serve as an excellent platform for genetic modification to produce specialized PHAs and enhancing PHA yield. The elucidation of PHA biosynthetic pathways is also likely to complement such research efforts. Many well-established methods are currently available for PHA analysis but each of them come with their own strengths and limitations. On the basis of the reports that have been gathered to date, GC-MS in conjunction with NMR remains a pre-eminent analytical tool in PHA investigations. Well-established analytical methods such as FTIR, GPC and X-ray diffraction can provide general information on the overall structures, molecular mass distribution and rate of crystallinity, respectively. However, it is important to further develop efficient technologies (e.g., LC-MS and MALDI-TOF-MS) for characterization of PHA. It is expected that the advanced analytical approaches will provide us with further insights about the physical properties and degradation mechanisms of PHA. AcknowledgmentThe authors gratefully acknowledge the financial support (ETRP 0901 161) from the National Environment Agency, Singapore.Conflicts of InterestThe authors declare no conflict of interest.ReferencesAnderson, A.J.; Dawes, E.A. Occurrence, metabolism, metabolic role, and industrial uses of bacterial polyhydroxyalkanoates. Microbiol. Rev. 1990, 54, 450–472. [Google Scholar]Shah, A.A.; Hasan, F.; Hameed, A.; Ahmed, S. Biological degradation of plastics: A comprehensive review. Biotechnol. Adv. 2008, 26, 246–265. [Google Scholar] [CrossRef]Khanna, S.; Srivastava, A.K. Recent advances in microbial polyhydroxyalkanoates. Process Biochem. 2005, 40, 607–619. [Google Scholar] [CrossRef]Lu, J.; Tappel, R.C.; Nomura, C.T. Mini-review: Biosynthesis of poly(hydroxyalkanoates). Polym. Rev. 2009, 49, 226–248. [Google Scholar] [CrossRef]Zinn, M.; Hany, R. Tailored material properties of polyhydroxyalkanoates through biosynthesis and chemical modification. Adv. Eng. Mater. 2005, 7, 408–411. [Google Scholar] [CrossRef]Escapa, I.F.; Morales, V.; Martino, V.P.; Pollet, E.; Avérous, L.; García, J.L.; Prieto, M.A. Disruption of β-oxidation pathway in Pseudomonas putida KT2442 to produce new functionalized PHAs with thioester groups. Appl. Microbiol. Biotechnol. 2011, 89, 1583–1598. [Google Scholar] [CrossRef]Rai, R.; Keshavarz, T.; Roether, J.A.; Boccaccini, A.R.; Roy, I. Medium chain length polyhydroxyalkanoates, promising new biomedical materials for the future. Mater. Sci. Eng. R. Rep. 2011, 72, 29–47. [Google Scholar] [CrossRef]De Roo, G.; Kellerhals, M.B.; Ren, Q.; Witholt, B.; Kessler, B. Production of chiral R-3-hydroxyalkanoic acids and R-3-hydroxyalkanoic acid methylesters via hydrolytic degradation of polyhydroxyalkanoate synthesized by Pseudomonads. Biotechnol. Bioeng. 2002, 77, 717–722. [Google Scholar] [CrossRef]Philip, S.; Keshavarz, T.; Roy, I. Polyhydroxyalkanoates: Biodegradable polymers with a range of applications. J. Chem. Technol. Biotechnol. 2007, 82, 233–247. [Google Scholar] [CrossRef]Olivera, E.R.; Arcos, M.; Naharro, G.; Luengo, J.M. Unusual PHA biosynthesis. In Plastics from Bacteria: Natural Functions and Applications; Chen, G.-Q., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 14, pp. 133–186. [Google Scholar]Chen, G.-Q. Plastics completely synthesized by bacteria: Polyhydroxyalkanoates. In Plastics from Bacteria: Natural Functions and Applications; Chen, G.-Q., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 14, pp. 17–37. [Google Scholar]Kadouri, D.; Jurkevitch, E.; Okon, Y.; Castro-Sowinski, S. Ecological and agricultural significance of bacterial polyhydroxyalkanoates. Crit. Rev. Microbiol. 2005, 31, 55–67. [Google Scholar] [CrossRef]Madison, L.L.; Huisman, G.W. Metabolic engineering of poly(3-hydroxyalkanoates): From DNA to plastic. Microbiol. Mol. Biol. Rev. 1999, 63, 21–53. [Google Scholar]Rothermich, M.M.; Guerrero, R.; Lenz, R.W.; Goodwin, S. Characterization, seasonal occurrence, and diel fluctuation of poly(hydroxyalkanoate) in photosynthetic microbial mats. Appl. Environ. Microbiol. 2000, 66, 4279–4291. [Google Scholar] [CrossRef]Peplinski, K.; Ehrenreich, A.; Döring, C.; Bömeke, M.; Reinecke, F.; Hutmacher, C.; Steinbüchel, A. Genome-wide transcriptome analyses of the “Knallgas” bacterium Ralstonia eutropha H16 with regard to polyhydroxyalkanoate metabolism. Microbiology 2010, 156, 2136–2152. [Google Scholar] [CrossRef]Shimizu, R.; Chou, K.; Orita, I.; Suzuki, Y.; Nakamura, S.; Fukui, T. Detection of phase-dependent transcriptomic changes and Rubisco-mediated CO2 fixation into poly (3-hydroxybutyrate) under heterotrophic condition in Ralstonia eutropha H16 based on RNA-seq and gene deletion analyses. BMC Microbiol. 2013, 13, 169. [Google Scholar] [CrossRef]Yamane, T. Yield of poly-D(-)-3-hydroxybutyrate from various carbon sources: A theoretical study. Biotechnol. Bioeng. 1993, 41, 165–170. [Google Scholar] [CrossRef]Steinbüchel, A.; Hein, S. Biochemical and molecular basis of microbial synthesis of polyhydroxyalkanoates in microorganisms. In Biopolyesters; Babel, W., Steinbüchel, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2001; Volume 71, pp. 81–123. [Google Scholar]Ratledge, C.; Kristiansen, B. Basic Biotechnology, 2nd ed.; Cambridge University Press: Cambridge, UK, 2001. [Google Scholar]Jung, Y.M.; Lee, Y.H. Utilization of oxidative pressure for enhanced production of poly-β-hydroxybutyrate and poly(3-hydroxybutyrate-3-hydroxyvalerate) in Ralstonia eutropha. J. Biosci. Bioeng. 2000, 90, 266–270. [Google Scholar]Khosravi-Darani, K.; Mokhtari, Z.-B.; Amai, T.; Tanaka, K. Microbial production of poly(hydroxybutyrate) from C1 carbon sources. Appl. Microbiol. Biotechnol. 2013, 97, 1407–1424. [Google Scholar] [CrossRef]Raberg, M.; Bechmann, J.; Brandt, U.; Schlüter, J.; Uischner, B.; Voigt, B.; Hecker, M.; Steinbüchel, A. Versatile metabolic adaptations of Ralstonia eutropha H16 to a loss of PdhL, the E3 component of the pyruvate dehydrogenase complex. Appl. Environ. Microbiol. 2011, 77, 2254–2263. [Google Scholar] [CrossRef]Peoples, O.P.; Sinskey, A.J. Poly-β-hydroxybutyrate (PHB) biosynthesis in Alcaligenes eutrophus H16. Identification and characterization of the PHB polymerase gene (phbC). J. Biol. Chem. 1989, 264, 15298–15303. [Google Scholar]Lee, S.; Jeon, E.; Yun, H.S.; Lee, J. Improvement of fatty acid biosynthesis by engineered recombinant Escherichia coli. Biotechnol. Bioprocess Eng. 2011, 16, 706–713. [Google Scholar] [CrossRef]Nomura, C.T.; Taguchi, K.; Taguchi, S.; Doi, Y. Coexpression of genetically engineered 3-ketoacyl-ACP synthase III (fabH) and polyhydroxyalkanoate synthase (phaC) genes leads to short-chain-length-medium-chain-length polyhydroxyalkanoate copolymer production from glucose in Escherichia coli JM109. Appl. Environ. Microbiol. 2004, 70, 999–1007. [Google Scholar] [CrossRef]Ren, Q.; Sierro, N.; Witholt, B.; Kessler, B. FabG, an NADPH-dependent 3-ketoacyl reductase of Pseudomonas aeruginosa, provides precursors for medium-chain-length poly-3-hydroxyalkanoate biosynthesis in Escherichia coli. J. Bacteriol. 2000, 182, 2978–2981. [Google Scholar] [CrossRef]Valentin, H.E.; Dennis, D. Production of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) in recombinant Escherichia coli grown on glucose. J. Biotechnol. 1997, 58, 33–38. [Google Scholar] [CrossRef]Xie, W.P.; Chen, G.-Q. Production and characterization of terpolyester poly(3-hydroxybutyrate-co-4-hydroxybutyrate-co-3-hydroxyhexanoate) by recombinant Aeromonas hydrophila 4AK4 harboring genes phaPCJ. Biochem. Eng. J. 2008, 38, 384–389. [Google Scholar] [CrossRef]Valentin, H.E.; Steinbüchel, A. Accumulation of poly(3-hydroxybutyric acid-co-3-hydroxyvaleric acid-co-4-hydroxyvaleric acid) by mutants and recombinant strains of Alcaligenes eutrophus. J. Polym. Environ. 1995, 3, 169–175. [Google Scholar] [CrossRef]Aldor, I.S.; Kim, S.W.; Jones Prather, K.L.; Keasling, J.D. Metabolic engineering of a novel propionate-independent pathway for the production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) in recombinant Salmonella enterica serovar Typhimurium. Appl. Environ. Microbiol. 2002, 68, 3848–3854. [Google Scholar] [CrossRef]Valentin, H.E.; Dennis, D. Metabolic pathway for poly(3-hydroxybutyrate-co-3-hydroxyvalerate) formation in Nocardia corallina: Inactivation of mutB by chromosomal integration of a kanamycin resistance gene. Appl. Environ. Microbiol. 1996, 62, 372–379. [Google Scholar]Satoh, H.; Mino, T.; Matsuo, T. PHA production by activated sludge. Int. J. Biol. Macromol. 1999, 25, 105–109. [Google Scholar] [CrossRef]Slater, S.; Houmiel, K.L.; Tran, M.; Mitsky, T.A.; Taylor, N.B.; Padgette, S.R.; Gruys, K.J. Multiple β-ketothiolases mediate poly(β-hydroxyalkanoate) copolymer synthesis in Ralstonia eutropha. J. Bacteriol. 1998, 180, 1979–1987. [Google Scholar]Chohan, S.N.; Copeland, L. Acetoacetyl coenzyme a reductase and polyhydroxybutyrate synthesis in Rhizobium (Cicer) sp. strain CC 1192. Appl. Environ. Microbiol. 1998, 64, 2859–2863. [Google Scholar]Hume, A.R.; Nikodinovic-Runic, J.; O’Connor, K.E. FadD from Pseudomonas putida CA-3 is a true long-chain fatty acyl coenzyme A synthetase that activates phenylalkanoic and alkanoic acids. J. Bacteriol. 2009, 191, 7554–7565. [Google Scholar] [CrossRef]Yuan, M.-Q.; Shi, Z.-Y.; Wei, X.-X.; Wu, Q.; Chen, S.-F.; Chen, G.-Q. Microbial production of medium-chain-length 3-hydroxyalkanoic acids by recombinant Pseudomonas putida KT2442 harboring genes fadL, fadD and phaZ. FEMS Microbiol. Lett. 2008, 283, 167–175. [Google Scholar] [CrossRef]Mittendorf, V.; Robertson, E.J.; Leech, R.M.; Krüger, N.; Steinbüchel, A.; Poirier, Y. Synthesis of medium-chain-length polyhydroxyalkanoates in Arabidopsis thaliana using intermediates of peroxisomal fatty acid β-oxidation. Proc. Natl. Acad. Sci. USA 1998, 95, 13397–13402. [Google Scholar] [CrossRef]Sato, S.; Kanazawa, H.; Tsuge, T. Expression and characterization of (R)-specific enoyl coenzyme A hydratases making a channeling route to polyhydroxyalkanoate biosynthesis in Pseudomonas putida. Appl. Microbiol. Biotechnol. 2011, 90, 951–959. [Google Scholar] [CrossRef]Ouyang, S.-P.; Luo, R.C.; Chen, S.-S.; Liu, Q.; Chung, A.; Wu, Q.; Chen, G.-Q. Production of polyhydroxyalkanoates with high 3-hydroxydodecanoate monomer content by fadB and fadA knockout mutant of Pseudomonas putida KT2442. Biomacromolecules 2007, 8, 2504–2511. [Google Scholar] [CrossRef]Zheng, L.Z.; Li, Z.; Tian, H.-L.; Li, M.; Chen, G.-Q. Molecular cloning and functional analysis of (R)-3-hydroxyacyl-acyl carrier protein:coenzyme A transacylase from Pseudomonas mendocina LZ. FEMS Microbiol. Lett. 2005, 252, 299–307. [Google Scholar] [CrossRef]Brzostowicz, P.B.; Blasko, M.B.; Rouvière, P.R. Identification of two gene clusters involved in cyclohexanone oxidation in Brevibacterium epidermidis strain HCU. Appl. Microbiol. Biotechnol. 2002, 58, 781–789. [Google Scholar] [CrossRef]Poli, A.; Di Donato, P.; Abbamondi, G.R.; Nicolaus, B. Synthesis, production, and biotechnological applications of exopolysaccharides and polyhydroxyalkanoates by archaea. Archaea 2011, 2011, 1–13. [Google Scholar]Galia, M.B. Isolation and analysis of storage compounds. In Handbook of Hydrocarbon and Lipid Microbiology; Timmis, K.N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 3725–3741. [Google Scholar]Verlinden, R.A.J.; Hill, D.J.; Kenward, M.A.; Williams, C.D.; Radecka, I. Bacterial synthesis of biodegradable polyhydroxyalkanoates. J. Appl. Microbiol. 2007, 102, 1437–1449. [Google Scholar] [CrossRef]Koller, M.; Atlić, A.; Dias, M.; Reiterer, A.; Braunegg, G. Microbial PHA production from waste raw materials. In Plastics from Bacteria: Natural Functions and Applications; Chen, G.-Q., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 14, pp. 85–119. [Google Scholar]Yu, P.H.; Chua, H.; Huang, A.L.; Lo, W.; Chen, G.-Q. Conversion of food industrial wastes into bioplastics. Appl. Biochem. Biotechnol. 1998, 70, 603–614. [Google Scholar]Wang, F.; Lee, S.Y. Poly(3-hydroxybutyrate) production with high productivity and high polymer content by a fed-batch culture of Alcaligenes latus under nitrogen limitation. Appl. Environ. Microbiol. 1997, 63, 3703–3706. [Google Scholar]Grothe, E.; Moo-Young, M.; Chisti, Y. Fermentation optimization for the production of poly(beta-hydroxybutyric acid) microbial thermoplastic. Enzyme Microb. Technol. 1999, 25, 132–141. [Google Scholar] [CrossRef]Yamane, T.; Fukunaga, M.; Lee, Y.W. Increased PHB productivity by high-cell-density fed-batch culture of Alcaligenes latus, a growth-associated PHB producer. Biotechnol. Bioeng. 1996, 50, 197–202. [Google Scholar] [CrossRef]Gomez, J.; Rodrigues, M.; Alli, R.; Torres, B.; Netto, C.B.; Oliveira, M.; Da Silva, L. Evaluation of soil Gram-negative bacteria yielding polyhydroxyalkanoic acids from carbohydrates and propionic acid. Appl. Microbiol. Biotechnol. 1996, 45, 785–791. [Google Scholar] [CrossRef]Lasemi, Z.; Darzi, G.N.; Baei, M.S. Media optimization for poly(β-hydroxybutyrate) production using Azotobacter Beijerinckii. Int. J. Polym. Mater. 2012, 62, 265–269. [Google Scholar] [CrossRef]Pan, W.; Perrotta, J.A.; Stipanovic, A.J.; Nomura, C.T.; Nakas, J.P. Production of polyhydroxyalkanoates by Burkholderia cepacia ATCC 17759 using a detoxified sugar maple hemicellulosic hydrolysate. J. Ind. Microbiol. Biotechnol. 2012, 39, 459–469. [Google Scholar] [CrossRef]Zhu, C.; Nomura, C.T.; Perrotta, J.A.; Stipanovic, A.J.; Nakas, J.P. Production and characterization of poly-3-hydroxybutyrate from biodiesel-glycerol by Burkholderia cepacia ATCC 17759. Biotechnol. Prog. 2010, 26, 424–430. [Google Scholar]Chee, J.-Y.; Tan, Y.; Samian, M.-R.; Sudesh, K. Isolation and characterization of a Burkholderia sp. USM (JCM15050) capable of producing polyhydroxyalkanoate (PHA) from triglycerides, fatty acids and glycerols. J. Polym. Environ. 2010, 18, 584–592. [Google Scholar] [CrossRef]Qi, Q.S.; Rehm, B.H.A. Polyhydroxybutyrate biosynthesis in Caulobacter crescentus: Molecular characterization of the polyhydroxybutyrate synthase. Microbiology 2001, 147, 3353–3358. [Google Scholar]Valentin, H.E.; Lee, E.Y.; Choi, C.Y.; Steinbüchel, A. Identification of 4-hydroxyhexanoic acid as a new constituent of biosynthetic polyhydroxyalkanoic acids from bacteria. Appl. Microbiol. Biotechnol. 1994, 40, 710–716. [Google Scholar] [CrossRef]Fukui, T.; Doi, Y. Efficient production of polyhydroxyalkanoates from plant oils by Alcaligenes eutrophus and its recombinant strain. Appl. Microbiol. Biotechnol. 1998, 49, 333–336. [Google Scholar] [CrossRef]Chakraborty, P.; Gibbons, W.; Muthukumarappan, K. Conversion of volatile fatty acids into polyhydroxyalkanoate by Ralstonia eutropha. J. Appl. Microbiol. 2009, 106, 1996–2005. [Google Scholar]Sonnleitner, B.; Heinzle, E.; Braunegg, G.; Lafferty, R.M. Formal kinetics of poly-β-hydroxybutyric acid (PHB) production in Alcaligenes eutrophus H 16 and Mycoplana rubra R 14 with respect to the dissolved oxygen tension in ammonium-limited batch cultures. Eur. J. Appl. Microbiol. 1979, 7, 1–10. [Google Scholar] [CrossRef]Ishizaki, A.; Tanaka, K. Production of poly-β-hydroxybutyric acid from carbon dioxide by Alcaligenes eutrophus ATCC 17697T. J. Ferment. Bioeng. 1991, 71, 254–257. [Google Scholar] [CrossRef]Kim, B.S.; Lee, S.C.; Lee, S.Y.; Chang, H.N.; Chang, Y.K.; Woo, S.I. Production of poly(3-hydroxybutyric acid) by fed-batch culture of Alcaligenes eutrophus with glucose concentration control. Biotechnol. Bioeng. 1994, 43, 892–898. [Google Scholar] [CrossRef]Haas, R.; Jin, B.; Zepf, F.T. Production of poly(3-hydroxybutyrate) from waste potato starch. Biosci. Biotechnol. Biochem. 2008, 72, 253–256. [Google Scholar]Beaulieu, M.; Beaulieu, Y.; Melinard, J.; Pandian, S.; Goulet, J. Influence of ammonium salts and cane molasses on growth of Alcaligenes eutrophus and production of polyhydroxybutyrate. Appl. Environ. Microbiol. 1995, 61, 165–169. [Google Scholar]Du, G.C.C.; Chen, J.; Yu, J.; Lun, S.Y. Feeding strategy of propionic acid for production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with Ralstonia eutropha. Biochem. Eng. J. 2001, 8, 103–110. [Google Scholar] [CrossRef]Cavalheiro, J.M.B.T.; de Almeida, M.C.M.D.; Grandfils, C.; da Fonseca, M.M.R. Poly(3-hydroxybutyrate) production by Cupriavidus necator using waste glycerol. Process Biochem. 2009, 44, 509–515. [Google Scholar] [CrossRef]Quillaguamán, J.; Hashim, S.; Bento, F.; Mattiasson, B.; Hatti-Kaul, R. Poly(β-hydroxybutyrate) production by a moderate halophile, Halomonas boliviensis LC1 using starch hydrolysate as substrate. J. Appl. Microbiol. 2005, 99, 151–157. [Google Scholar] [CrossRef]Povolo, S.; Romanelli, M.G.; Basaglia, M.; Ilieva, V.I.; Corti, A.; Morelli, A.; Chiellini, E.; Casella, S. Polyhydroxyalkanoate biosynthesis by Hydrogenophaga pseudoflava DSM1034 from structurally unrelated carbon sources. New Biotechnol. 2013, 30, 629–634. [Google Scholar] [CrossRef]Koller, M.; Hesse, P.; Bona, R.; Kutschera, C.; Atlic, A.; Braunegg, G. Potential of various archae- and eubacterial strains as industrial polyhydroxyalkanoate producers from whey. Macromol. Biosci. 2007, 7, 218–226. [Google Scholar] [CrossRef]Bourque, D.; Pomerleau, Y.; Groleau, D. High-cell-density production of poly-β-hydroxybutyrate (PHB) from methanol by Methylobacterium extorquens: Production of high-molecular-mass PHB. Appl. Microbiol. Biotechnol. 1995, 44, 367–376. [Google Scholar] [CrossRef]Mokhtari-Hosseini, Z.B.; Vasheghani-Farahani, E.; Heidarzadeh-Vazifekhoran, A.; Shojaosadati, S.A.; Karimzadeh, R.; Darani, K.K. Statistical media optimization for growth and PHB production from methanol by a methylotrophic bacterium. Bioresour. Technol. 2009, 100, 2436–2443. [Google Scholar] [CrossRef]Mokhtari-Hosseini, Z.B.; Vasheghani-Farahani, E.; Shojaosadati, S.A.; Karimzadeh, R.; Heidarzadeh-Vazifekhoran, A. Effect of feed composition on PHB production from methanol by HCDC of Methylobacterium extorquens (DSMZ 1340). J. Chem. Technol. Biotechnol. 2009, 84, 1136–1139. [Google Scholar] [CrossRef]Wendlandt, K.D.; Jechorek, M.; Helm, J.; Stottmeister, U. Production of PHB with a high molecular mass from methane. Polym. Degrad. Stabil. 1998, 59, 191–194. [Google Scholar] [CrossRef]Smit, A.M.; Strabala, T.J.; Peng, L.; Rawson, P.; Lloyd-Jones, G.; Jordan, T.W. Proteomic phenotyping of Novosphingobium nitrogenifigens reveals a robust capacity for simultaneous nitrogen fixation, polyhydroxyalkanoate production, and resistance to reactive oxygen species. Appl. Environ. Microbiol. 2012, 78, 4802–4815. [Google Scholar] [CrossRef]Yamane, T.; Chen, X.; Ueda, S. Growth-associated production of poly(3-hydroxyvalerate) from n-pentanol by a methylotrophic bacterium, Paracoccus denitrificans. Appl. Environ. Microbiol. 1996, 62, 380–384. [Google Scholar]Tripathi, A.D.; Yadav, A.; Jha, A.; Srivastava, S.K. Utilizing of sugar refinery waste (cane molasses) for production of bio-plastic under submerged fermentation process. J. Polym. Environ. 2012, 20, 446–453. [Google Scholar] [CrossRef]Guzik, M.W.; Kenny, S.T.; Duane, G.F.; Casey, E.; Woods, T.; Babu, R.P.; Nikodinovic-Runic, J.; Murray, M.; O’Connor, K.E. Conversion of post consumer polyethylene to the biodegradable polymer polyhydroxyalkanoate. Appl. Microbiol. Biotechnol. 2014. In Press. [Google Scholar]Kenny, S.T.; Nikodinovic-Runic, J.; Kaminsky, W.; Woods, T.; Babu, R.P.; Keely, C.M.; Blau, W.; O’Connor, K.E. Up-cycling of PET (polyethylene terephthalate) to the biodegradable plastic PHA (polyhydroxyalkanoate). Environ. Sci. Technol. 2008, 42, 7696–7701. [Google Scholar] [CrossRef]Lee, E.; Jendrossek, D.; Schirmer, A.; Choi, C.; Steinbüchel, A. Biosynthesis of copolyesters consisting of 3-hydroxybutyric acid and medium-chain-length 3-hydroxyalkanoic acids from 1,3-butanediol or from 3-hydroxybutyrate by Pseudomonas sp. A33. Appl. Microbiol. Biotechnol. 1995, 42, 901–909. [Google Scholar] [CrossRef]Nikodinovic-Runic, J.; Casey, E.; Duane, G.F.; Mitic, D.; Hume, A.R.; Kenny, S.T.; O’Connor, K.E. Process analysis of the conversion of styrene to biomass and medium chain length polyhydroxyalkanoate in a two-phase bioreactor. Biotechnol. Bioeng. 2011, 108, 2447–2455. [Google Scholar] [CrossRef]Ward, P.G.; Goff, M.; Donner, M.; Kaminsky, W.; O'Connor, K.E. A two step chemo-biotechnological conversion of polystyrene to a biodegradable thermoplastic. Environ. Sci. Technol. 2006, 40, 2433–2437. [Google Scholar] [CrossRef]Lageveen, R.G.; Huisman, G.W.; Preusting, H.; Ketelaar, P.; Eggink, G.; Witholt, B. Formation of polyesters by Pseudomonas oleovorans: Effect of substrates on formation and composition of poly-(R)-3-hydroxyalkanoates and poly-(R)-3-hydroxyalkenoates. Appl. Environ. Microbiol. 1988, 54, 2924–2932. [Google Scholar]Gross, R.A.; DeMello, C.; Lenz, R.W.; Brandl, H.; Fuller, R.C. The biosynthesis and characterization of poly(β-hydroxyalkanoates) produced by Pseudomonas oleovorans. Macromolecules 1989, 22, 1106–1115. [Google Scholar] [CrossRef]Elbahloul, Y.; Steinbuchel, A. Large-scale production of poly(3-hydroxyoctanoic acid) by Pseudomonas putida GPo1 and a simplified downstream process. Appl. Environ. Microbiol. 2009, 75, 643–651. [Google Scholar] [CrossRef]Sun, Z.; Ramsay, J.A.; Guay, M.; Ramsay, B.A. Carbon-limited fed-batch production of medium-chain-length polyhydroxyalkanoates from nonanoic acid by Pseudomonas putida KT2440. Appl. Microbiol. Biotechnol. 2007, 74, 69–77. [Google Scholar] [CrossRef]Davis, R.; Kataria, R.; Cerrone, F.; Woods, T.; Kenny, S.; O’Donovan, A.; Guzik, M.; Shaikh, H.; Duane, G.; Gupta, V.K.; et al. Conversion of grass biomass into fermentable sugars and its utilization for medium chain length polyhydroxyalkanoate (mcl-PHA) production by Pseudomonas strains. Bioresour. Technol. 2013, 150, 202–209. [Google Scholar] [CrossRef]Nikodinovic, J.; Kenny, S.T.; Babu, R.P.; Woods, T.; Blau, W.; O’Connor, K.E. The conversion of BTEX compounds by single and defined mixed cultures to medium-chain-length polyhydroxyalkanoate. Appl. Microbiol. Biotechnol. 2008, 80, 665–673. [Google Scholar] [CrossRef]Shahid, S.; Mosrati, R.; Ledauphin, J.; Amiel, C.; Fontaine, P.; Gaillard, J.-L.; Corroler, D. Impact of carbon source and variable nitrogen conditions on bacterial biosynthesis of polyhydroxyalkanoates: Evidence of an atypical metabolism in Bacillus megaterium DSM 509. J. Biosci. Bioeng. 2013, 116, 302–308. [Google Scholar] [CrossRef]Pantazaki, A.A.; Papaneophytou, C.P.; Pritsa, A.G.; Liakopoulou-Kyriakides, M.; Kyriakidis, D.A. Production of polyhydroxyalkanoates from whey by Thermus thermophilus HB8. Process Biochem. 2009, 44, 847–853. [Google Scholar] [CrossRef]Valappil, S.P.; Boccaccini, A.R.; Bucke, C.; Roy, I. Polyhydroxyalkanoates in Gram-positive bacteria: Insights from the genera Bacillus and Streptomyces. Antonie Van Leeuwenhoek 2007, 91, 1–17. [Google Scholar]Haywood, G.W.; Anderson, A.J.; Roger Williams, D.; Dawes, E.A.; Ewing, D.F. Accumulation of a poly(hydroxyalkanoate) copolymer containing primarily 3-hydroxyvalerate from simple carbohydrate substrates by Rhodococcus sp. NCIMB 40126. Int. J. Biol. Macromol. 1991, 13, 83–88. [Google Scholar] [CrossRef]Akar, A.; Akkaya, E.U.; Yesiladali, S.K.; Celikyilmaz, G.; Cokgor, E.U.; Tamerler, C.; Orhon, D.; Cakar, Z.P. Accumulation of polyhydroxyalkanoates by Microlunatus phosphovorus under various growth conditions. J. Ind. Microbiol. Biotechnol. 2006, 33, 215–220. [Google Scholar] [CrossRef]Bhattacharyya, A.; Pramanik, A.; Maji, S.K.; Haldar, S.; Mukhopadhyay, U.K.; Mukherjee, J. Utilization of vinasse for production of poly-3-(hydroxybutyrate-co-hydroxyvalerate) by Haloferax mediterranei. AMB Express 2012, 2, 1–10. [Google Scholar] [CrossRef]Koller, M.; Atlić, A.; Gonzalez‐Garcia, Y.; Kutschera, C.; Braunegg, G. Polyhydroxyalkanoate (PHA) biosynthesis from whey lactose. Macromol. Symp. 2008, 272, 87–92. [Google Scholar] [CrossRef]Hermann-Krauss, C.; Koller, M.; Muhr, A.; Fasl, H.; Stelzer, F.; Braunegg, G. Archaeal production of polyhydroxyalkanoate (PHA) co-and terpolyesters from biodiesel industry-derived by-products. Archaea 2013, 2013. [Google Scholar]Han, J.; Hou, J.; Liu, H.; Cai, S.; Feng, B.; Zhou, J.; Xiang, H. Wide distribution among halophilic archaea of a novel polyhydroxyalkanoate synthase subtype with homology to bacterial type III synthases. Appl. Environ. Microbiol. 2010, 76, 7811–7819. [Google Scholar] [CrossRef]Chen, G.-Q. Industrial production of PHA. In Plastics from Bacteria: Natural Functions and Applications; Chen, G.-Q., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 14, pp. 121–132. [Google Scholar]Chanprateep, S. Current trends in biodegradable polyhydroxyalkanoates. J. Biosci. Bioeng. 2010, 110, 621–632. [Google Scholar] [CrossRef]Poblete-Castro, I.; Becker, J.; Dohnt, K.; dos Santos, V.M.; Wittmann, C. Industrial biotechnology of Pseudomonas putida and related species. Appl. Microbiol. Biotechnol. 2012, 93, 2279–2290. [Google Scholar] [CrossRef]Greene, E.A.; Voordouw, G. Biodegradation of C5+ hydrocarbons by a mixed bacterial consortium from a C5+-contaminated site. Environ. Technol. 2004, 25, 355–363. [Google Scholar] [CrossRef]Jung, I.-G.; Park, C.-H. Characteristics of styrene degradation by Rhodococcus pyridinovorans isolated from a biofilter. Chemosphere 2005, 61, 451–456. [Google Scholar] [CrossRef]Gąszczak, A.; Bartelmus, G.; Greń, I. Kinetics of styrene biodegradation by Pseudomonas sp. E-93486. Appl. Microbiol. Biotechnol. 2012, 93, 565–573. [Google Scholar] [CrossRef]Ray, A.; Cot, M.; Puzo, G.; Gilleron, M.; Nigou, J. Bacterial cell wall macroamphiphiles: Pathogen-/microbe-associated molecular patterns detected by mammalian innate immune system. Biochimie 2013, 95, 33–42. [Google Scholar] [CrossRef]Chen, G.-Q.; Wu, Q. The application of polyhydroxyalkanoates as tissue engineering materials. Biomaterials 2005, 26, 6565–6578. [Google Scholar] [CrossRef]Wampfler, B.; Ramsauer, T.; Rezzonico, S.; Hischier, R.; Kohling, R.; Thony-Meyer, L.; Zinn, M. Isolation and purification of medium chain length poly(3-hydroxyalkanoates) (mcl-PHA) for medical applications using nonchlorinated solvents. Biomacromolecules 2010, 11, 2716–2723. [Google Scholar] [CrossRef]Karr, D.B.; Waters, J.K.; Emerich, D.W. Analysis of poly-β-hydroxybutyrate in Rhizobium japonicum bacteroids by ion-exclusion high-pressure liquid chromatography and UV detection. Appl. Environ. Microbiol. 1983, 46, 1339–1344. [Google Scholar]Misaki, A.; Azuma, I.; Yamamura, Y. Structural and immunochemical studies on D-arabino-D-mannans and D-mannans of Mycobacterium tuberculosis and other Mycobacterium species. J. Biochem. 1977, 82, 1759–1770. [Google Scholar]Nigou, J.; Gilleron, M.; Puzo, G. Lipoarabinomannans: From structure to biosynthesis. Biochimie 2003, 85, 153–166. [Google Scholar] [CrossRef]Sutcliffe, I.; Brown, A.; Dover, L. The Rhodococcal cell envelope: Composition, organisation and biosynthesis. In Biology of Rhodococcus; Alvarez, H.M., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 16, pp. 29–71. [Google Scholar]Ruhland, G.J.; Fiedler, F. Occurrence and structure of lipoteichoic acids in the genus Staphylococcus. Arch. Microbiol. 1990, 154, 375–379. [Google Scholar]Sutcliffe, I.C. The lipoteichoic acids and lipoglycans of Gram-positive bacteria: A chemotaxonomic perspective. Syst. Appl. Microbiol. 1995, 17, 467–480. [Google Scholar] [CrossRef]Iwasaki, H.; Shimada, A.; Yokoyama, K.; Ito, E. Structure and glycosylation of lipoteichoic acids in Bacillus strains. J. Bacteriol. 1989, 171, 424–429. [Google Scholar]Sutcliffe, I.C.; Shaw, N. Atypical lipoteichoic acids of Gram-positive bacteria. J. Bacteriol. 1991, 173, 7065–7069. [Google Scholar]Danson, M.J.; Hough, D.W. The structural basis of protein halophilicity. Comp. Biochem. Physiol. Part. A Physiol. 1997, 117, 307–312. [Google Scholar] [CrossRef]Chen, G.-Q.; Zhang, G.; Park, S.; Lee, S. Industrial scale production of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Appl. Microbiol. Biotechnol. 2001, 57, 50–55. [Google Scholar] [CrossRef]Choi, J.; Lee, S.Y. Factors affecting the economics of polyhydroxyalkanoate production by bacterial fermentation. Appl. Microbiol. Biotechnol. 1999, 51, 13–21. [Google Scholar] [CrossRef]Ganduri, V.; Ghosh, S.; Patnaik, P. Mixing control as a device to increase PHB production in batch fermentations with co-cultures of Lactobacillus delbrueckii and Ralstonia eutropha. Process Biochem. 2005, 40, 257–264. [Google Scholar] [CrossRef]Shi, H.; Shiraishi, M.; Shimizu, K. Metabolic flux analysis for biosynthesis of poly(β-hydroxybutyric acid) in Alcaligenes eutrophus from various carbon sources. J. Ferment. Bioeng. 1997, 84, 579–587. [Google Scholar] [CrossRef]Tanaka, K.; Katamune, K.; Ishizaki, A. Fermentative production of poly(β-hydroxybutyric acid) from xylose via L-lactate by a two-stage culture method employing Lactococcus lactis IO-1 and Alcaligenes eutrophus. Can. J. Microbiol. 1995, 41, 257–261. [Google Scholar] [CrossRef]van der Ha, D.; Nachtergaele, L.; Kerckhof, F.-M.; Rameiyanti, D.; Bossier, P.; Verstraete, W.; Boon, N. Conversion of biogas to bioproducts by algae and methane oxidizing bacteria. Environ. Sci. Technol. 2012, 46, 13425–13431. [Google Scholar] [CrossRef]Romo, D.M.R.; Grosso, M.V.; Solano, N.C.M.; Castaño, D.M. A most effective method for selecting a broad range of short and medium-chain-length polyhidroxyalcanoate producing microorganisms. Electron. J. Biotechnol. 2007, 10, 348–357. [Google Scholar]Solaiman, D.K.; Ashby, R.D. Rapid genetic characterization of poly(hydroxyalkanoate) synthase and its applications. Biomacromolecules 2005, 6, 532–537. [Google Scholar] [CrossRef]Spiekermann, P.; Rehm, B.H.; Kalscheuer, R.; Baumeister, D.; Steinbüchel, A. A sensitive, viable-colony staining method using Nile red for direct screening of bacteria that accumulate polyhydroxyalkanoic acids and other lipid storage compounds. Arch. Microbiol. 1999, 171, 73–80. [Google Scholar] [CrossRef]Melnicki, M.R.; Eroglu, E.; Melis, A. Changes in hydrogen production and polymer accumulation upon sulfur-deprivation in purple photosynthetic bacteria. Int. J. Hydrog. Energy 2009, 34, 6157–6170. [Google Scholar] [CrossRef]Wu, H.A.; Sheu, D.S.; Lee, C.Y. Rapid differentiation between short-chain-length and medium-chain-length polyhydroxyalkanoate-accumulating bacteria with spectrofluorometry. J. Microbiol. Meth. 2003, 53, 131–135. [Google Scholar] [CrossRef]Ostle, A.G.; Holt, J. Nile blue A as a fluorescent stain for poly-beta-hydroxybutyrate. Appl. Environ. Microbiol. 1982, 44, 238–241. [Google Scholar]Law, J.H.; Slepecky, R.A. Assay of poly-β-hydroxybutyric acid. J. Bacteriol. 1961, 82, 33–36. [Google Scholar]Hong, K.; Sun, S.; Tian, W.; Chen, G.-Q.; Huang, W. A rapid method for detecting bacterial polyhydroxyalkanoates in intact cells by fourier transform infrared spectroscopy. Appl. Microbiol. Biotechnol. 1999, 51, 523–526. [Google Scholar] [CrossRef]Gumel, A.; Annuar, M.; Heidelberg, T. Effects of carbon substrates on biodegradable polymer composition and stability produced by Delftia tsuruhatensis Bet002 isolated from palm oil mill effluent. Polym. Degrad. Stabil. 2012, 97, 1224–1231. [Google Scholar] [CrossRef]Hesselmann, R.P.; Fleischmann, T.; Hany, R.; Zehnder, A.J. Determination of polyhydroxyalkanoates in activated sludge by ion chromatographic and enzymatic methods. J. Microbiol. Meth. 1999, 35, 111–119. [Google Scholar] [CrossRef]Grubelnik, A.; Wiesli, L.; Furrer, P.; Rentsch, D.; Hany, R.; Meyer, V.R. A simple HPLC-MS method for the quantitative determination of the composition of bacterial medium chain-length polyhydroxyalkanoates. J. Sep. Sci. 2008, 31, 1739–1744. [Google Scholar] [CrossRef]Furrer, P.; Hany, R.; Rentsch, D.; Grubelnik, A.; Ruth, K.; Panke, S.; Zinn, M. Quantitative analysis of bacterial medium-chain-length poly([R]-3-hydroxyalkanoates) by gas chromatography. J. Chromatogr. A 2007, 1143, 199–206. [Google Scholar] [CrossRef]Tan, G.-Y.A.; Chen, C.-L.; Ge, L.; Li, L.; Wang, L.; Zhao, L.; Mo, Y.; Tan, S.N.; Wang, J.-Y. Enhanced gas chromatography-mass spectrometry method for bacterial polyhydroxyalkanoates (PHAs) analysis. J. Biosci. Bioeng. 2014, 117, 379–382. [Google Scholar] [CrossRef]Lee, E.Y.; Choi, C.Y. Gas chromatography-mass spectrometric analysis and its application to a screening procedure for novel bacterial polyhydroxyalkanoic acids containing long chain saturated and unsaturated monomers. J. Ferment. Bioeng. 1995, 80, 408–414. [Google Scholar] [CrossRef]Braunegg, G.; Sonnleitner, B.; Lafferty, R. A rapid gas chromatographic method for the determination of poly-β-hydroxybutyric acid in microbial biomass. Eur. J. Appl. Microbiol. 1978, 6, 29–37. [Google Scholar] [CrossRef]Ward, A.C.; Dawes, E.A. A disk assay for poly-β-hydroxybutyrate. Anal. Biochem. 1973, 52, 607–613. [Google Scholar] [CrossRef]Slepecky, R.A.; Law, J.H. A rapid spectrophotometric assay of alpha, beta-unsaturated acids and beta-hydroxy acids. Anal. Chem. 1960, 32, 1697–1699. [Google Scholar] [CrossRef]Arcos-Hernandez, M.V.; Gurieff, N.; Pratt, S.; Magnusson, P.; Werker, A.; Vargas, A.; Lant, P. Rapid quantification of intracellular PHA using infrared spectroscopy: An application in mixed cultures. J. Biotechnol. 2010, 150, 372–379. [Google Scholar]de Rijk, T.C.; van de Meer, P.; Eggink, G.; Weusthuis, R.A. Methods for analysis of poly(3-hydroxyalkanoate) (PHA) composition. In Biopolymers Online; Doi, Y., Steinbüchel, A., Eds.; Wiley-VCH: Weinheim, Germany, 2005; Volume 3b, pp. 1–12. [Google Scholar]Korotkova, N.A.; Ashin, V.; Doronina, N.V.; Trotsenko, Y.A. A new method for quantitative determination of poly-3-hydroxybutyrate and 3-hydroxybutyrate-3-hydroxyvalerate copolymer in microbial biomass by reversed-phase high-performance liquid chromatography. Appl. Biochem. Microbiol. 1997, 33, 302–305. [Google Scholar]Zhang, S.; Norrlöw, O.; Wawrzynczyk, J.; Dey, E.S. Poly(3-hydroxybutyrate) biosynthesis in the biofilm of Alcaligenes eutrophus, using glucose enzymatically released from pulp fiber sludge. Appl. Environ. Microbiol. 2004, 70, 6776–6782. [Google Scholar] [CrossRef]Comeau, Y.; Hall, K.J.; Oldham, W.K. Determination of poly-β-hydroxybutyrate and poly-β-hydroxyvalerate in activated sludge by gas-liquid chromatography. Appl. Environ. Microbiol. 1988, 54, 2325–2327. [Google Scholar]Gumel, A.M.; Annuar, M.S.M.; Heidelberg, T. Biosynthesis and characterization of polyhydroxyalkanoates copolymers produced by Pseudomonas putida Bet001 isolated from palm oil mill effluent. PLoS One 2012, 7, e45214. [Google Scholar]Kunasundari, B.; Sudesh, K. Isolation and recovery of microbial polyhydroxyalkanoates. Express Polym. Lett. 2011, 5, 620–634. [Google Scholar] [CrossRef]Ramsay, J.A.; Berger, E.; Voyer, R.; Chavarie, C.; Ramsay, B.A. Extraction of poly-3-hydroxybutyrate using chlorinated solvents. Biotechnol. Tech. 1994, 8, 589–594. [Google Scholar] [CrossRef]Koller, M.; Bona, R.; Chiellini, E.; Braunegg, G. Extraction of short-chain-length poly-[(R)-hydroxyalkanoates] (scl-PHA) by the “anti-solvent” acetone under elevated temperature and pressure. Biotechnol. Lett. 2013, 35, 1023–1028. [Google Scholar] [CrossRef]Nonato, R.; Mantelatto, P.; Rossell, C. Integrated production of biodegradable plastic, sugar and ethanol. Appl. Microbiol. Biotechnol. 2001, 57, 1–5. [Google Scholar] [CrossRef]Berger, E.; Ramsay, B.A.; Ramsay, J.A.; Chavarie, C.; Braunegg, G. PHB recovery by hypochlorite digestion of non-PHB biomass. Biotechnol. Tech. 1989, 3, 227–232. [Google Scholar] [CrossRef]Hahn, S.K.; Chang, Y.K.; Kim, B.S.; Chang, H.N. Optimization of microbial poly(3-hydroxybutyrate) recover using dispersions of sodium hypochlorite solution and chloroform. Biotechnol. Bioeng. 1994, 44, 256–261. [Google Scholar] [CrossRef]Kapritchkoff, F.M.; Viotti, A.P.; Alli, R.C.P.; Zuccolo, M.; Pradella, J.G.C.; Maiorano, A.E.; Miranda, E.A.; Bonomi, A. Enzymatic recovery and purification of polyhydroxybutyrate produced by Ralstonia eutropha. J. Biotechnol. 2006, 122, 453–462. [Google Scholar] [CrossRef]Ashby, R.; Solaiman, D.; Foglia, T. Poly(ethylene glycol)-mediated molar mass control of short-chain- and medium-chain-length poly(hydroxyalkanoates) from Pseudomonas oleovorans. Appl. Microbiol. Biotechnol. 2002, 60, 154–159. [Google Scholar] [CrossRef]Durner, R.; Zinn, M.; Witholt, B.; Egli, T. Accumulation of poly[(R)-3-hydroxyalkanoates] in Pseudomonas oleovorans during growth in batch and chemostat culture with different carbon sources. Biotechnol. Bioeng. 2000, 72, 278–288. [Google Scholar]Jiang, X.; Ramsay, J.A.; Ramsay, B.A. Acetone extraction of mcl-PHA from Pseudomonas putida KT2440. J. Microbiol. Meth. 2006, 67, 212–219. [Google Scholar] [CrossRef]Hahn, S.K.; Chang, Y.K.; Lee, S.Y. Recovery and characterization of poly(3-hydroxybutyric acid) synthesized in Alcaligenes eutrophus and recombinant Escherichia coli. Appl. Environ. Microbiol. 1995, 61, 34–39. [Google Scholar]Cromwick, A.M.; Foglia, T.; Lenz, R.W. The microbial production of poly(hydroxyalkanoates) from tallow. Appl. Microbiol. Biotechnol. 1996, 46, 464–469. [Google Scholar] [CrossRef]Ashby, R.D.; Foglia, T.A. Poly(hydroxyalkanoate) biosynthesis from triglyceride substrates. Appl. Microbiol. Biotechnol. 1998, 49, 431–437. [Google Scholar] [CrossRef]Sánchez, R.J.; Schripsema, J.; da Silva, L.F.; Taciro, M.K.; Pradella, J.G.C.; Gomez, J.G.C. Medium-chain-length polyhydroxyalkanoic acids (PHAmcl) produced by Pseudomonas putida IPT 046 from renewable sources. European Polymer Journal 2003, 39, 1385–1394. [Google Scholar] [CrossRef]Fernández, D.; Rodríguez, E.; Bassas, M.; Viñas, M.; Solanas, A.M.; Llorens, J.; Marqués, A.M.; Manresa, A. Agro-industrial oily wastes as substrates for PHA production by the new strain Pseudomonas aeruginosa NCIB 40045: Effect of culture conditions. Biochem. Eng. J. 2005, 26, 159–167. [Google Scholar] [CrossRef]Yasotha, K.; Aroua, M.K.; Ramachandran, K.B.; Tan, I.K.P. Recovery of medium-chain-length polyhydroxyalkanoates (PHAs) through enzymatic digestion treatments and ultrafiltration. Biochem. Eng. J. 2006, 30, 260–268. [Google Scholar] [CrossRef]De Koning, G.J.M.; Kellerhals, M.; van Meurs, C.; Witholt, B. A process for the recovery of poly(hydroxyalkanoates) from Pseudomonads-Part 2. Bioprocess Eng. 1997, 17, 15–21. [Google Scholar] [CrossRef]De Koning, G.J.M.; Witholt, B. A process for the recovery of poly(hydroxyalkanoates) from Pseudomonads-Part 1. Bioprocess Eng. 1997, 17, 7–13. [Google Scholar] [CrossRef]Dai, Y.; Lambert, L.; Yuan, Z.; Keller, J. Characterisation of polyhydroxyalkanoate copolymers with controllable four-monomer composition. J. Biotechnol. 2008, 134, 137–145. [Google Scholar] [CrossRef]Saeed, K.A.; Ayorinde, F.O.; Eribo, B.E.; Gordon, M.; Collier, L. Characterization of partially transesterified poly(β-hydroxyalkanoate)s using matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. Rapid Commun. Mass Spectrom. 1999, 13, 1951–1957. [Google Scholar]Li, Z.; Yang, X.; Wu, L.; Chen, Z.; Lin, Y.; Xu, K.; Chen, G.-Q. Synthesis, characterization and biocompatibility of biodegradable elastomeric poly(ether-ester urethane)s based on poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) and poly(ethylene glycol) via melting polymerization. J. Biomater. Sci., Polym. Ed. 2009, 20, 1179–1202. [Google Scholar]Katime, I.; Cadenato, A. Compatibility of peo/poly(iso-butyl methacrylate) and peo/poly(tert-butyl methacrylate) blends by DTA. Mater. Lett. 1995, 22, 303–308. [Google Scholar] [CrossRef]Shamala, T.; Divyashree, M.; Davis, R.; Kumari, K.L.; Vijayendra, S.; Raj, B. Production and characterization of bacterial polyhydroxyalkanoate copolymers and evaluation of their blends by fourier transform infrared spectroscopy and scanning electron microscopy. Indian J. Microbiol. 2009, 49, 251–258. [Google Scholar] [CrossRef]Chen, S.; Liu, Q.; Wang, H.; Zhu, B.; Yu, F.; Chen, G.-Q.; Inoue, Y. Polymorphic crystallization of fractionated microbial medium-chain-length polyhydroxyalkanoates. Polymer 2009, 50, 4378–4388. [Google Scholar] [CrossRef]Wu, C.-S.; Liao, H.-T. The mechanical properties, biocompatibility and biodegradability of chestnut shell fibre and polyhydroxyalkanoate composites. Polym. Degrad. Stabil. 2014, 99, 274–282. [Google Scholar] [CrossRef]Teeka, J.; Imai, T.; Reungsang, A.; Cheng, X.; Yuliani, E.; Thiantanankul, J.; Poomipuk, N.; Yamaguchi, J.; Jeenanong, A.; Higuchi, T.; et al. Characterization of polyhydroxyalkanoates (PHAs) biosynthesis by isolated Novosphingobium sp. THA_AIK7 using crude glycerol. J. Ind. Microbiol. Biotechnol. 2012, 39, 749–758. [Google Scholar] [CrossRef]Meng, D.-C.; Shi, Z.-Y.; Wu, L.-P.; Zhou, Q.; Wu, Q.; Chen, J.-C.; Chen, G.-Q. Production and characterization of poly(3-hydroxypropionate-co-4-hydroxybutyrate) with fully controllable structures by recombinant Escherichia coli containing an engineered pathway. Metab. Eng. 2012, 14, 317–324. [Google Scholar] [CrossRef]Hartmann, R.; Hany, R.; Geiger, T.; Egli, T.; Witholt, B.; Zinn, M. Tailored biosynthesis of olefinic medium-chain-length poly[(R)-3-hydroxyalkanoates] in Pseudomonas putida GPo1 with improved thermal properties. Macromolecules 2004, 37, 6780–6785. [Google Scholar] [CrossRef]Hany, R.; Hartmann, R.; Böhlen, C.; Brandenberger, S.; Kawada, J.; Löwe, C.; Zinn, M.; Witholt, B.; Marchessault, R.H. Chemical synthesis and characterization of POSS-functionalized poly[3-hydroxyalkanoates]. Polymer 2005, 46, 5025–5031. [Google Scholar] [CrossRef]Wang, Q.; Tappel, R.C.; Zhu, C.; Nomura, C.T. Development of a new strategy for production of medium-chain-length polyhydroxyalkanoates by recombinant Escherichia coli via inexpensive non-fatty acid feedstocks. Appl. Environ. Microbiol. 2012, 78, 519–527. [Google Scholar] [CrossRef]De Waard, P.; van der Wal, H.; Huijberts, G.; Eggink, G. Heteronuclear NMR analysis of unsaturated fatty acids in poly(3-hydroxyalkanoates). Study of beta-oxidation in Pseudomonas putida. J. Biol. Chem. 1993, 268, 315–319. [Google Scholar]Eggink, G.; de Waard, P.; Huijberts, G.N. Formation of novel poly(hydroxyalkanoates) from long-chain fatty acids. Can. J. Microbiol. 1995, 41, 14–21. [Google Scholar] [CrossRef]Tripathi, L.; Wu, L.-P.; Dechuan, M.; Chen, J.; Wu, Q.; Chen, G.-Q. Pseudomonas putida KT2442 as a platform for the biosynthesis of polyhydroxyalkanoates with adjustable monomer contents and compositions. Bioresour. Technol. 2013, 142, 225–231. [Google Scholar] [CrossRef]Ballistreri, A.; Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Montaudo, G. Sequencing bacterial poly(β-hydroxybutyrate-co-β-hydroxyvalerate) by partial methanolysis, HPLC fractionation, and fast-atom-bombardment mass spectrometry analysis. Macromolecules 1989, 22, 2107–2111. [Google Scholar] [CrossRef]Ballistreri, A.; Montaudo, G.; Garozzo, D.; Giuffrida, M.; Montaudo, M.S. Microstructure of bacterial poly(β-hydroxybutyrate-co-β-hydroxyvalerate) by fast atom bombardment mass spectrometry analysis of the partial pyrolysis products. Macromolecules 1991, 24, 1231–1236. [Google Scholar]Abate, R.; Ballistreri, A.; Montaudo, G.; Garozzo, D.; Impallomeni, G.; Critchley, G.; Tanaka, K. Quantitative applications of matrix-assisted laser desorption/ionization with time-of-flight mass spectrometry: Determination of copolymer composition in bacterial copolyesters. Rapid Commun. Mass Spectrom. 1993, 7, 1033–1036. [Google Scholar] [CrossRef]Saeed, K.A.; Ayorinde, F.O.; Eribo, B.E.; Gordon, M.; Collier, L. Characterization of partially transesterified poly(beta-hydroxyalkanoate)s by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. J. AOAC Int. 2001, 84, 1109–11015. [Google Scholar]Chen, G.-Q. Introduction of bacterial plastics PHA, PLA, PBS, PE, PTT, and PPP. In Plastics from Bacteria: Natural Functions and Applications; Chen, G.-Q., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 14, pp. 1–16. [Google Scholar]Koller, M.; Bona, R.; Braunegg, G.; Hermann, C.; Horvat, P.; Kroutil, M.; Martinz, J.; Neto, J.; Pereira, L.; Varila, P. Production of polyhydroxyalkanoates from agricultural waste and surplus materials. Biomacromolecules 2005, 6, 561–565. [Google Scholar] [CrossRef]Li, L. MALDI-MS for polymer characterization. In MALDI MS: A Practical Guide to Instrumentation, Methods, and Applications, 2nd ed.; Hillenkamp, F., Peter-Katalinic, J., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; pp. 313–365. [Google Scholar]Aoyagi, Y.; Yamashita, K.; Doi, Y. Thermal degradation of poly[(R)-3-hydroxybutyrate], poly[ε-caprolactone], and poly[(S)-lactide]. Polym. Degrad. Stabil. 2001, 76, 53–59. [Google Scholar] [CrossRef]Bloembergen, S.; Holden, D.A.; Hamer, G.K.; Bluhm, T.L.; Marchessault, R.H. Studies of composition and crystallinity of bacterial poly(β-hydroxybutyrate-co-β-hydroxyvalerate). Macromolecules 1986, 19, 2865–2871. [Google Scholar] [CrossRef]Porter, M.; Yu, J. Monitoring the in situ crystallization of native biopolyester granules in Ralstonia eutropha via infrared spectroscopy. J. Microbiol. Meth. 2011, 87, 49–55. [Google Scholar] [CrossRef]Barham, P.J.; Keller, A.; Otun, E.L.; Holmes, P.A. Crystallization and morphology of a bacterial thermoplastic: Poly-3-hydroxybutyrate. J. Mater. Sci. 1984, 19, 2781–2794. [Google Scholar] [CrossRef]Simon-Colin, C.; Raguénès, G.; Crassous, P.; Moppert, X.; Guezennec, J. A novel mcl-PHA produced on coprah oil by Pseudomonas guezennei biovar. tikehau, isolated from a “kopara” mat of French Polynesia. Int. J. Biol. Macromol. 2008, 43, 176–181. [Google Scholar] [CrossRef]Cheng, S.-T.; Chen, Z.-F.; Chen, G.-Q. The expression of cross-linked elastin by rabbit blood vessel smooth muscle cells cultured in polyhydroxyalkanoate scaffolds. Biomaterials 2008, 29, 4187–4194. [Google Scholar] [CrossRef]Dufresne, A.; Kellerhas, M.B.; Witholt, B. Transcrystallization in mcl-PHAs: Cellulose whiskers composites. Macromolecules 1999, 32, 7396–7401. [Google Scholar] [CrossRef] Share and Cite MDPI and ACS Style

Tan, G.-Y.A.; Chen, C.-L.; Li, L.; Ge, L.; Wang, L.; Razaad, I.M.N.; Li, Y.; Zhao, L.; Mo, Y.; Wang, J.-Y. Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review. Polymers 2014, 6, 706-754. https://doi.org/10.3390/polym6030706

AMA Style

Tan G-YA, Chen C-L, Li L, Ge L, Wang L, Razaad IMN, Li Y, Zhao L, Mo Y, Wang J-Y. Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review. Polymers. 2014; 6(3):706-754. https://doi.org/10.3390/polym6030706

Chicago/Turabian Style

Tan, Giin-Yu Amy, Chia-Lung Chen, Ling Li, Liya Ge, Lin Wang, Indah Mutiara Ningtyas Razaad, Yanhong Li, Lei Zhao, Yu Mo, and Jing-Yuan Wang. 2014. "Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review" Polymers 6, no. 3: 706-754. https://doi.org/10.3390/polym6030706

Find Other Styles Article Metrics No No Article Access Statistics For more information on the journal statistics, click here. Multiple requests from the same IP address are counted as one view. Zoom | Orient | As Lines | As Sticks | As Cartoon | As Surface | Previous Scene | Next Scene Cite Export citation file: BibTeX | EndNote | RIS MDPI and ACS Style

Tan, G.-Y.A.; Chen, C.-L.; Li, L.; Ge, L.; Wang, L.; Razaad, I.M.N.; Li, Y.; Zhao, L.; Mo, Y.; Wang, J.-Y. Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review. Polymers 2014, 6, 706-754. https://doi.org/10.3390/polym6030706

AMA Style

Tan G-YA, Chen C-L, Li L, Ge L, Wang L, Razaad IMN, Li Y, Zhao L, Mo Y, Wang J-Y. Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review. Polymers. 2014; 6(3):706-754. https://doi.org/10.3390/polym6030706

Chicago/Turabian Style

Tan, Giin-Yu Amy, Chia-Lung Chen, Ling Li, Liya Ge, Lin Wang, Indah Mutiara Ningtyas Razaad, Yanhong Li, Lei Zhao, Yu Mo, and Jing-Yuan Wang. 2014. "Start a Research on Biopolymer Polyhydroxyalkanoate (PHA): A Review" Polymers 6, no. 3: 706-754. https://doi.org/10.3390/polym6030706

Find Other Styles clear Polymers, EISSN 2073-4360, Published by MDPI RSS Content Alert Further Information Article Processing Charges Pay an Invoice Open Access Policy Contact MDPI Jobs at MDPI Guidelines For Authors For Reviewers For Editors For Librarians For Publishers For Societies For Conference Organizers MDPI Initiatives Sciforum MDPI Books Preprints.org Scilit SciProfiles Encyclopedia JAMS Proceedings Series Follow MDPI LinkedIn Facebook Twitter MDPI © 1996-2023 MDPI (Basel, Switzerland) unless otherwise stated Disclaimer Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. Terms and Conditions Privacy Policy We use cookies on our website to ensure you get the best experience. Read more about our cookies here. Accept We have just recently launched a new version of our website. Help us to further improve by taking part in this short 5 minute survey here. here. Never show this again Share Link Copy clear Share https://www.mdpi.com/67904 clear Back to TopTop


【本文地址】


今日新闻


推荐新闻


CopyRight 2018-2019 办公设备维修网 版权所有 豫ICP备15022753号-3