Critical Review

Subscriber access provided by Otterbein University

Methane as a resource: can the methanotrophs add value? PJ Strong, Sihuang Xie, and William P. Clarke Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es504242n • Publication Date (Web): 27 Feb 2015 Downloaded from http://pubs.acs.org on March 4, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 51

Environmental Science & Technology

Methane as a resource: can the methanotrophs add value?

Authors P.J. Strong* S. Xie W.P. Clarke

Affiliation Centre for Solid Waste Bioprocessing, School of Civil Engineering, School of Chemical Engineering, The University of Queensland, St Lucia, 4072, Australia.

*Corresponding author. P.J. Strong [email protected] Tel: +61 459652099 Centre for Solid Waste Bioprocessing, School of Civil Engineering, School of Chemical Engineering, The University of Queensland, St Lucia, 4072, Australia.

ACS Paragon Plus Environment

Environmental Science & Technology

Rationale This review considers all of the biological options for either converting methane into a product or using it directly for purposes other the production of electricity or heat. Methane is cheap and abundant, does not compete with food demand, and in the case of anaerobic digestion is renewable carbon source. It is therefore a suitable substrate by which to generate products or drive processes. To our knowledge there is currently no review that comprehensively covers the potential products or processes that could be generated using methane-consuming bacteria. Generally review papers regard methane in terms of energy capture or emissions prevention. The methanotrophs have rarely been regarded for more than their remediative abilities or for single-cell protein production. Biological reviews of the methanotrophs often cover a few individual biotechnological applications or just one in detail. In this review we explore all the avenues for biological applications, thereby providing a single source for the readers that will be readily citable. Some of the options discussed are nascent and even unproven, but represent exciting potential research avenues. It will appeal to both a popular audience that is interested in applied alternatives to mitigating anthropogenic methane emissions, as well as a scientific audience and applied researchers interested in current developments and an assessment of the available options.

The Table of Contents graphic below was created by the corresponding author.

Table of Contents graphic

ACS Paragon Plus Environment

Page 2 of 51

Page 3 of 51

Environmental Science & Technology

Abstract Methane is an abundant gas used in energy recovery systems, heating and transport. Methanotrophs are bacteria capable of using methane as their sole carbon source. Although intensively-researched, the myriad of potential biotechnological applications of methanotrophic bacteria has not been comprehensively discussed in a single review.

Methanotrophs can generate single-cell protein, biopolymers, components for nanotechnology applications (surface layers), soluble metabolites (methanol, formaldehyde, organic acids and ectoine), lipids (biodiesel and health supplements), growth media and vitamin B12 using methane as their carbon source. They may be genetically engineered to produce new compounds such as carotenoids or farnesene. Some enzymes (dehydrogenases, oxidase and catalase) are valuable products with high conversion efficiencies and can generate methanol or sequester CO2 as formic acid ex vivo. Live cultures can be used for bioremediation, chemical transformation (propene to propylene oxide), wastewater denitrification, as components of biosensors, or possibly for directly generating electricity.

This review demonstrates the potential for methanotrophs and their consortia to generate value while using methane as a carbon source. While there are notable challenges using a low solubility gas as a carbon source, the massive methane resource, and the potential cost savings while sequestering a greenhouse gas, keeps interest piqued in these unique bacteria.

Keywords biofuel; bioremediation; coal seam gas; natural gas; proteobacteria; methane monooxygenase

ACS Paragon Plus Environment

Environmental Science & Technology

1

1. Methane

2

Methane is a colourless, odourless gas that is emitted from both natural and anthropogenic sources.

3

It provides energy or heat via combustion. However, its emission into the atmosphere has negative

4

consequences as it is a greenhouse gas with approximately twenty times the impact of carbon

5

dioxide. Anthropogenic activity accounts for the majority of global methane emissions (63 %, or

6

566 Tg CH4/year), with natural biological emissions accounting for the remainder (208 Tg CH4/year)1.

7

Anthropogenic methane emissions are generated by the use of fossil fuels, livestock farming,

8

landfilling and biomass burning. Natural sources of methane are wetlands, oceans, estuaries, rivers,

9

lakes, permafrost, gas hydrates, geological sources (terrestrial and marine), wildfires, vegetation,

10

terrestrial arthropods and wild animals 1-2,2b. The ratio of anthropogenic:natural methane production

11

has increased steadily since the advent of the industrial revolution. With increased food

12

requirements, greater waste generation and greater use of fossil fuels by an increasing human

13

population, it will in all likelihood increase further.

14 15

The principal use of methane is as a fuel, as its combustion is highly exothermic:

16

CH4(g) + 2 O2(g) →CO2(g) + 2 H2O(l) where ∆H = -891 kJ.

17

Methane is primarily used for generating electricity in gas turbines or steam boilers, but is also piped

18

into homes for domestic heating and cooking, or used as a vehicle fuel as compressed natural gas3.

19

The reciprocating engine, gas turbine or steam turbine technologies are available over a wide range,

20

from modular units of several hundred kW up to 250 MW for commercial steam turbines4. Although

21

methane can be used as a transport fuel, there is currently a relative lack of infrastructure for

22

fuelling vehicles compatible with natural gas. However, new approaches whereby methane is

23

converted into liquid transportation fuels can take advantage of existing engines and delivery

24

infrastructure5.

25

ACS Paragon Plus Environment

Page 4 of 51

Page 5 of 51

Environmental Science & Technology

26

From a biological perspective, methane represents a carbon and energy source for a group of

27

bacteria known as methanotrophs. Methanotrophs use methane as their sole carbon source and

28

directly convert methane into cellular compounds, or transform it into a substrate that drives

29

processes via methanotrophs or their syntrophic interaction of other microbes. The biological

30

monetisation of methane has become a topic of intense interest: in 2013 the Advanced Research

31

Projects Agency (ARPA-E) within the U.S. Department of Energy granted US$34 million worth of

32

funding directed towards research converting methane into liquid fuels. There are notable barriers

33

to overcome, but the size of the methane resource is enormous, and in the case of anaerobic

34

digestion is sustainably produced; this justifies the continued interest in these bacteria that can

35

oxidise methane under ambient conditions.

36 37

2. Methanotrophs

38

Methanotrophs, a subset of the methylotrophs, can assimilate methane as their sole carbon source.

39

Methanotrophs fall under proteobacteria; a major phylum of gram-negative bacteria that includes

40

genera such as Escherichia, Salmonella, Vibrio, Helicobacter and Yersinia. The proteobacteria phylum

41

is divided into six classes according to ribosomal RNA sequences. Two classes contain

42

methanotrophs: the alpha-proteobacteria and the gamma-proteobacteria. Methanotrophs are

43

traditionally classified as Type I (γ-proteobacteria) or Type II (α-proteobacteria). The Type I and II

44

distinctions were primarily based on the metabolic pathway used to assimilate formaldehyde, and

45

related attributes such as cell membrane composition and arrangement and cell morphology6.

46

Typically, Type I methanotrophs are γ-proteobacteria that assimilate formaldehyde via the ribulose

47

monophosphate pathway6b, 6d, 7, while Type II are α-proteobacteria that use the serine pathway. The

48

generalised differences between Type I and Type II methanotrophs are summarised in Table 1, but

49

exceptions are common and metabolic pathway flexibility may be greater than previously thought8.

50

This is evident for Type I methanotrophs, which contain a sub-division denoted as Type X

51

(Methylocaldum and Methylococcus species) that express enzymes associated with the serine

ACS Paragon Plus Environment

Environmental Science & Technology

52

pathway typical of the α-proteobacteria6b. Methane can also be oxidised by a phylum of

53

thermoacidophilic bacteria known as Verrucomicrobia (which can grow at a pH below 1)9, as well as

54

certain Archaea in consortia with sulfate reducing bacteria10. The ability of certain Archea to

55

anaerobically oxidise methane (where NOx11 or SOx12 replace O2) corroborates research supporting

56

the concept of reverse methanogenesis13.

57 58

Methane consumption is made possible by an enzyme known as methane mono-oxygenase (MMO),

59

which occurs in a particulate form (pMMO) within an intracellular membrane or a soluble form

60

(sMMO) within the cytoplasm14. Copper availability plays a defining regulatory role with regard to

61

sMMO expression15, where sMMO synthesis is inhibited by higher Cu2+ concentrations16. The sMMO

62

has a di-iron catalytic site, while research suggests that pMMO requires both copper and iron to be

63

catalytically active7b. The sMMO can be produced by various α- or γ-proteobacteria and has a much

64

broader substrate range than the pMMO. It is capable of oxidising numerous compounds, including

65

propene, butane, cyclohexane, chlorotrifluoroethylene, toluene, naphthalene, chloroform, diethyl

66

ether and CO (substrates tabled in 14b). Although pMMO is expressed by most methanotrophs, some

67

Methylocella spp. express sMMO exclusively17.

68 69

The metabolic activity and growth of methanotrophs are influenced factors that typically affect

70

microbial cultures, including nutrient supply, temperature, pH, macronutrients and trace metals.

71

Traditionally, the growth of γ-proteobacteria was reported to favour higher O2 concentrations and

72

low CH4 concentrations, while α-proteobacteria methanotrophs preferred low O2 and higher CH4

73

concentrations6d,

74

Similarly, facultative methanotrophy was initially discounted, but has been demonstrated in certain

75

genera of α-proteobacteria7b, 21, 21b. Methanotrophs are able to assimilate CO2 and are thus affected

76

by CO2 concentrations. The α-proteobacteria can assimilate up to 50 % of their biomass from CO2,

77

while the γ-proteobacteria can assimilate up to 15 %22. The N source in media commonly used to

18, 19

, but γ-proteobacteria can grow efficiently under a low oxygen tension20.

ACS Paragon Plus Environment

Page 6 of 51

Page 7 of 51

Environmental Science & Technology

78

culture methanotrophs is usually nitrate (e.g. nitrate minimal salts) or ammonia. Ammonia affects

79

different methanotrophs to differing extents and can be a competitive inhibitor of methane

80

monooxygenase or toxic when in the intermediate forms of hydroxylamine and nitrite23. However,

81

ammonium acts more as a nutrient than an inhibitor in the presence of sufficient methane.

82

Nitrogenase occurs in a broad range of methanotrophs (α- and γ-proteobacteria) allowing for use of

83

N2 as a nitrogen source; nitrogen-fixing may be widespread and important to nitrogen cycling in

84

many environments23.

85 86

3. Methanotrophs: prior reviews

87

Many aspects of these ubiquitous bacteria have been reviewed. This includes the natural occurrence

88

of methanotrophs24, their use in methane mitigation and environmental remediation6d, 14b, 25, their

89

physiology, biochemical pathways for N metabolism26, C1 metabolism and assimilation22,

90

enzymes involved6c,

91

facultative methanotrophy7b, 21b, 30.

28

and the influence of copper on their metabolic capabilities27a,

27

, the

28c, 29

and

92 93

Although the biotechnological applications have been reviewed in part, no single review has

94

comprehensively covered all prospective products or processes. Specific topics such as biodiesel

95

generation5b, epoxide production31, Polyhydroxybutyrate accumulation32 and denitrification33 have

96

been covered individually, while some reviews cover many facets of methanotrophy and some

97

biotechnological applications6b,

98

development of various aspects of methanotrophic physiology and biochemistry, as well as views on

99

the development of single cell protein production, pollutant bioremediation, whole cell catalytic

100

production of propylene oxide (epoxypropane) and alludes to the possibility for electrochemically

101

driving reactions that require reducing agents such as NADH. Trotsenko, et al. 34 reviewed single cell

102

protein, biopolymer, ectoine, surface layers, vitamins, pollutant bioremediation and the role of

103

methanotroph as plant growth-promoting bacteria. These topics, as well as extracellular

27a

. A review by Dalton

14a

discusses much of the history and

ACS Paragon Plus Environment

Environmental Science & Technology

104

polysaccharides, lipids for biodiesel and human health supplements, growth media, methanol,

105

formaldehyde and organic acids, enzymes, enzymatic transformation of CH4 and CO2 (incorporating

106

electrochemical reductant recycling), and products resulting from genetic engineering are covered in

107

the current review.

108 109

3. Methanotrophs: products and processes

110

3.1. Products

111

3.1.1. Single-cell protein

112

Single-cell protein (SCP) production from microbes such as yeasts, fungi, algae, and bacteria surged

113

in the 1950s and 1960s due to a lack of protein source and predictions of impending global

114

shortages. The advent of low cost soya production in the 1970s negatively impacted microbial SCP

115

research, although the production of a fungal protein was commercialised in 1985. Currently, SCP is

116

the closest example of a commercially-successful biologically-generated product using methane as

117

the carbon source. The commercial production of methanotrophic SCP originated as research in

118

Denmark by E.B. Larsen in the 1980s and has culminated the company now known as UniBio A/S.

119

The fermentation uses natural gas, technically pure oxygen, ammonia as the N source, phosphoric

120

acid as the P source and supplemented with other minerals. The process is controlled at pH 6.5 and

121

run at a temperature of 45°C. The concentrated biomass is sterilised by rapid heating to 140°C and

122

then slowly cooled, allowing cells lysis and accessibility to the protein. The bacterial biomass consists

123

predominantly of a strain of Methylococcus capsulatus (Bath) and is a promising protein source,

124

based on criteria such as amino acid composition, digestibility, and animal performance and

125

health35. The methanotrophic protein has been used as a protein source for several mono-gastric

126

species, including pigs, broiler chickens, mink, fox, Atlantic salmon, rainbow trout, and Atlantic

127

halibut35.

128

ACS Paragon Plus Environment

Page 8 of 51

Page 9 of 51

Environmental Science & Technology

129

Although bacteria possess a high protein content (ranging from 50 to 65 %) relative to other micro-

130

organisms (30 to 60 %), they also possess the highest nucleic acid content (8 to 12 %)36, which can

131

adversely affect consumers. A report regarding the safety of the UniBio protein (then known as

132

Dansk BioProtein®) noted significant immune effects in rats during a toxicity study (NCSF, 2006). The

133

protein was only approved for animal nutrition and was recommended for animals with a short life

134

span, as the nucleic acids could cause kidney and bladder stones in longer-lived species. However,

135

the risk associated with human consumption of products from animals fed on BioProtein® was

136

considered negligible. Heat treatment or hydrolysis can lower RNA content. At UniBio A/S nucleic

137

acids are neutralised by hydrolysis, generating a product fit for human consumption. Heat treatment

138

at 60 to 65°C for 10 to 20 minutes has been used to remove nearly 90 % of the RNA from a

139

Methylomonas sp. cultured for single-cell protein use37.

140 141

Although using a C1 gas as the substrate is highly selective, large-scale continuous industrial

142

fermentation is always subject to infiltration by other micro-organisms as cell lysis products or

143

metabolic byproducts serve as C sources, which is problematic in a continuous process. The SCP

144

strain was repeatedly contaminated by three different bacteria: a member of the Aneurinibacillus

145

group, a Brevibacillus agri strain and an acetate-oxidising Ralstonia species. Fortunately, the three

146

contaminant bacteria were non-toxic. Their presence was beneficial as they stabilised the culture by

147

consuming metabolic byproducts that would have inhibited growth38.

148 149

3.1.2a. Biopolymers: Internal storage polymers

150

Polyhydroxy-alkanoates (PHA) such as poly(3-hydroxybutyrate), or PHB, are widespread and

151

intensively-researched bacterial storage polymers seen as potential substitutes for plastics derived

152

from the petroleum industry. They have beneficial properties such as biodegradability,

153

biocompatibility and thermoplasticity. These biopolymers are synthesised and deposited intra-

154

cellularly as granules that serve as a source of carbon, energy or reducing-power and, in exceptional

ACS Paragon Plus Environment

Environmental Science & Technology

155

circumstances, may comprise up to 90 % of a microbe’s dry weight39. PHB accumulation is induced

156

by exposing an active culture to excess carbon while under a nutrient limitation of some sort. PHA

157

yield and quality, is affected by pH, temperature and the availability of other carbon sources,

158

methane, oxygen, carbon dioxide, macronutrients (nitrogen, phosphorus, sulfur, potassium,

159

magnesium sodium) and trace metals (copper, iron, zinc, manganese cobalt)40. Although PHB

160

production by methanotrophs and methylotrophs has been of interest for decades41, it has been re-

161

evaluated in recent times5b, 40c. Cheap, or zero cost, substrates such as CH4 and CO2 are receiving

162

considerable attention because of their potential to lower production costs32, 42.

163 164

Microbe strain selection, or genetic engineering of existing strains, is critical because of the impact

165

on the maximum production rate and yield. In the 1990s, heterotrophically-grown Alcaligenes

166

eutrophus became the organism of choice for industrial PHB production as it produced high yields of

167

high molecular weight PHBs using various economically-acceptable substrates. It was benchmarked

168

against a methylotroph (Methylobacterium sp.), which only managed moderate polymers yields that

169

had a low molecular weight (Mw) and were difficult to extract43. To date, many methylotrophs and

170

methanotrophs have been assessed for PHA production and are tabled in a reviews by Khosravi-

171

Darani, et al. 32 and Karthikeyan, et al. 40c. Polymer yields and Mw have improved among the newly-

172

isolated bacteria as well as mixed cultures. Zhang, et al.

173

using methane with methanol and citric acid and obtained a high quality PHB (Mw: 1.5×106 Da) at a

174

yield 40 %. Wendlandt, et al. 45 also produced a high quality PHB with a high molecular mass of (up

175

to 2.5 × 106 Da), using a Methylocystis sp. in a rapid, non-sterile process. They obtained a PHB

176

content of up to 51 % using a two-stage process consisting of a continuous growth stage (dilution

177

rate: 0.17 h−1) and a PHB accumulation stage under P-deficient conditions14c. Shah et al.46 enriched

178

their methanotrophs for pMMO or sMMO and compared PHB accumulation under different batch

179

conditions. The pMMO-rich cells displayed greater and more rapid PHB accumulation (up to 50 %

180

PHB content within 120 hours) and attained a much greater biomass yield (18 g/l). The pMMO-rich

44

cultured a Methylosinus trichosporium

ACS Paragon Plus Environment

Page 10 of 51

Page 11 of 51

Environmental Science & Technology

181

cells also. Recently, researchers included silicone oil (10 % v/v) in the two-phase partitioning

182

bioreactor (using 1 % methane in an air stream) with a co-culture of Methylobacterium

183

organophilum and reported up to 57 % PHB under nitrogen limitation47 and improved methane

184

consumption by up to 45 % in the growth stage. Methylotrophs have improved PHB accumulation

185

when ammonium was the sole N source under potassium limitation48. A complex nitrogen source

186

can also improve biopolymer yields when using a defined medium49. Helm, et al.

187

polymer with an ultra-high average Mw of 3.1 MDa under potassium-limited conditions using

188

methane-utilizing mixed culture where a Methylocystis sp. was dominant. A maximum specific PHB

189

formation rate 0.08 g.g-1.h-1 and a yield coefficient of 0.45 g PHB.g-1 CH4 were obtained in further K-

190

deficiency experiments. The Mw was lower when sulfur (21 % lower) or iron (42 % lower) were

191

limiting.

50

produced a

192 193

Even after decades of considerable effort to commercialise microbial PHB production, the high

194

production cost compared to traditional petrochemical-based plastics (such as polyethylene and

195

polypropylene) still limits commercial application. Choi and Lee 51 performed a sensitivity analysis of

196

several factors (productivity, content and yield, the cost of the carbon substrate, and the recovery

197

method) that affect PHA production with a view to scale-up and found biopolymer content had

198

multiple effects on the process economics. Productivity only affected equipment-related costs, but

199

the yield per cell had multiple effects on process economics. Other research has also placed a firm

200

emphasis on the importance of yield52. Although substrate costs can account for a major portion

201

(30 %) of the production cost39b, 51-53, the costs associated with downstream processing would still

202

render microbial polymer production uncompetitive with the petroleum-based polymers. An initial

203

bridge to commercial production is to target higher-value polymers used in biomedical applications

204

5b

205

medical devices, tissue repair, artificial organ construction and nutritional/therapeutic uses55. This

, as the biocompatibility and biodegradation of PHA has shown potential for drug delivery54,

ACS Paragon Plus Environment

Environmental Science & Technology

206

field of research has received a promising boost: the FDA has recently approved the use of poly-4-

207

hydroxybutyrate for a clinical application56.

208 209

3.1.2b. Biopolymers: Extra-cellular polysaccharides

210

The colloid and adhesive properties of extra-cellular polysaccharides (EPS) and their effects on liquid

211

rheology are used in the food and non-food industries such as the pharmaceutical, textile and oil

212

industries. Traditionally, industrial polysaccharides have been derived from algal and plant sources57.

213

There has been an active search for producers that may be cultured using non-edible raw

214

materials58. Methanotrophs satisfy this requirement as they can synthesise EPS using methane as

215

their sole carbon source58-59. The genes responsible for elements of the EPS synthesis have been

216

isolated from a Methylomonas sp. with the intention of genetically engineering them into other C1-

217

utilising microorganisms to alter levels of EPS production for commercial production57. Malashenko

218

et al. 58 studied the dynamics of EPS production by methanotrophs in chemostats and observed EPS

219

synthesis rates ranging from 0.03 to 0.43 g.g-1 dry biomass. Highest production was achieved using

220

Type I isolates (Methylomonas and Methylobacter spp.). The viscosity of 0.1 % aqueous solutions of

221

EPS synthesised by the mesophilic methanotrophs varied from 2.2 to 4.0 mm2/s, which was

222

comparable to the viscosity of EPS solutions synthesised by known microbial producers as well as the

223

benchmark equivalent composed of 0.1 % xanthane (3.5 to 4.5 mm2/s).

224 225

However, EPS production may inhibit further synthesis as it negatively affects gas uptake.

226

Chiemchaisri and Visvanathan59a studied methanotrophic EPS production in a bench-scale soil

227

reactor. Methane oxidation rates (regulated by temperature and soil water content) were correlated

228

to EPS production and highest EPS production occurred at 30 °C. They observed that a high oxygen

229

content accelerated EPS production, but subsequently limited gas diffusion and inhibited

230

production.

231

ACS Paragon Plus Environment

Page 12 of 51

Page 13 of 51

Environmental Science & Technology

232

3.1.3. Internal osmo-protectants: Ectoine

233

Halo-tolerant bacteria employ two primary survival strategies: they synthesise and accumulate

234

intracellular organic osmo-protectants (such as ectoine, glutamate and sucrose), or they make

235

structural and functional changes to their cell envelopes by altering the phospholipids composition

236

of membranes and forming glycoprotein surface layers14b, 60. Ectoine is a cyclic imino acid that is one

237

of the most widespread microbial protective measures against osmotic dehydration. It is also an

238

efficient stabiliser of enzymes, nucleic acids, and DNA-protein complexes, and can be used as

239

a moisturiser in the cosmetic industry. Cosmetic formulations comprising ectoine or its derivatives

240

offer excellent protection against UV-induced damage to the DNA of skin cells61. Ectoine is produced

241

annually on a scale of tons in an industrial process using the halophilic γ-proteobacterium

242

Halomonas elongata DSM 2581T as producer strain62.

243 244

Because synthesis and purification is expensive, new microbial producer strains and their enzymes

245

are constantly assessed to improve the economics of microbial ectoine production. Halo-tolerant

246

methanotrophs are known and isolated, with the best-known belonging to the genus

247

Methylomicrobium. Trotsenko and his colleagues demonstrated that moderately halophilic

248

methanotrophs and methylotrophs were able to accumulate up to 20 % of their dry mass as

249

ectoine34, 60. This is an area well worth further exploration, as purified ectoine has one of the highest

250

retail values of the methanotroph products (approximately $1300 kg-1). Additionally, the bacteria

251

may potentially be reused multiple times: Halomonas elongata has been reused 9 times, yielding

252

ectoine at an average of 15.5% g.g-1 biomass63. However, downstream processing is complex and

253

expensive and represents a challenge to process economics.

254 255

3.1.4. External osmo-protectants: Surface layers

256

Bacterial cell surface layers are one of the most commonly observed outermost structures of

257

prokaryotes. These regular para-crystalline structures cover the entire surface of a cell and consist of

ACS Paragon Plus Environment

Environmental Science & Technology

258

a single layer of identical proteins or glycoproteins. Isolated surface layer glycoproteins possess the

259

intrinsic property of self-assembly and recrystallise into isoporous lattices in suspension, onto

260

various surfaces (polymer, silicon and metal) and interfaces (air-liquid, lipid films and liposomes).

261

These characteristics and subsequent functionalising of surfaces has led to new types of

262

ultrafiltration membranes, affinity structures, enzyme membranes, micro-carriers, biosensors,

263

diagnostic devices, biocompatible surfaces and vaccines, as well as targeting, delivery, and

264

encapsulation systems34, 64.

265 266

As with the de novo synthesis of internal solutes such as ectoine, the osmo-adaptation of

267

methanotrophs also involves structural and functional changes to cell envelopes and changes in the

268

chemical composition of membranes60. These changes in the surface layers are potentially of

269

industrial interest. Regularly arranged glycoprotein surface layers of hexagonal and linear symmetry

270

have been observed on the outer cell walls of two halo-tolerant Methylobacter spp. Interestingly,

271

the surface layer in a Methylomicrobium sp. (consisting of tightly packed, cup-shaped sub-units) was

272

negligible when the bacteria were cultured at a neutral pH with no salt in the media 60, indicating the

273

stimulatory effect of osmotic stress. With methanotrophs, it is likely that proteins associated with

274

surface layers facilitate copper ion transport to pMMO and provide an additional mechanism to

275

maintain copper homeostasis in the cells65. Surface layers have prospective nanotechnology

276

applications in ultrafiltration because they form porous semi-permeable membranes; their

277

components undergo self-assembly and in vitro cross-linking at the surface of membranes. The

278

characteristics of isolated surface layers have allowed for various applications in biotechnology,

279

vaccine development, diagnostics, biomimetics and molecular nanotechnology34, 64b. Methanotrophs

280

may be in contention for commercial production if highly desired properties are discovered or

281

engineered for their specific proteins or glycoproteins.

282

ACS Paragon Plus Environment

Page 14 of 51

Page 15 of 51

Environmental Science & Technology

283

3.1.5a. Lipids: biodiesel

284

Biodiesel is another possible methane-generated commodity with a potentially enormous market. It

285

may be used as a transport fuel or an energy fuel, and is advantageous compared to ethanol as it can

286

be used without any modifications to engines. Typically, animal or plant fats (triglycerides) are

287

extracted and converted via trans-esterification into biodiesel. The triglyceride consists of three long

288

chain fatty acids attached to a glycerine molecule. The triglyceride reacts with alcohol (typically

289

methanol or ethanol) in the presence of a catalyst (usually a strong base: NaOH or KOH) to yield

290

esters and glycerol (Figure 2).

291 292

Oleaginous microbes, including yeasts and various microalgae, have been researched intensively as a

293

biological means to produce diesel due to their ability to accumulate a lipid content greater than

294

20 % of their dry mass. Microbially-produced lipids are advantageous to animal or plant-based lipids

295

as they are produced in a short life cycle, may be less labour-intensive, may be less affected by

296

location, season and climate, and in certain cases may be easier to scale-up. However, there are

297

difficulties. The scale-up of dense cultures of autotrophic microalgae is difficult due to light

298

penetration66, difficulties obtaining a high lipid content with dense microbial67 and the high degree

299

of processing associated with lipid extraction and transformation5b.

300 301

Essentially, lipids from microorganisms may be divided into two types: those associated with storage

302

(non-polar lipids) and those that form structures (polar lipids). The non-polar storage lipids are

303

mainly in the form of triacylglycerides and can be trans-esterified to produce biodiesel. The

304

structural polar lipids (phospholipids) and sterols are important components of cell membranes that

305

typically have a high content of polyunsaturated fatty acids (PUFAs) and are essential nutrients for

306

humans 67. In microalgae, one of the most intensively researched oleaginous microbes, the types of

307

lipids produced range from neutral lipids to polar lipids, wax esters, sterols and hydrocarbons, as

308

well as prenyl derivatives such as tocopherols, carotenoids, terpenes, quinines and pyrrole

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 51

309

derivatives. In methanotrophs, the two major classes of phospholipids are phosphatidyl-glycerol (PG)

310

and phosphatidyl-ethanolamine (PE), which has two derivatives: phosphatidyl methyl ethanolamine

311

(PME) and phosphatidyl dimethyl ethanolamine (PDME). Gamma-proteobacteria (Methylomonas

312

methanica, Methylomonas rubra and Methylomicrobium album BG8) contain PE and PG

313

phospholipids with predominantly C16:1 fatty acids, while alpha-proteobacteria methanotrophs

314

(Methylosinus trichosporium OB3b and CSC1) contain PG, PME and PDME with predominantly C18:1

315

fatty acids

316

production, the sugars, P and S contents are problematic to subsequent catalysts used in the

317

processing. The high hetero-atom content (specifically P and N) in the lipid fraction makes

318

downstream extraction and conversion difficult5b. Undesirable components such as sugars,

319

phosphorous and sulphur exacerbate processing problems due to gumming or catalyst inactivation.

320

This will require additional research and development of a solvent-based extraction process to

321

minimise these negative impacts5b.

68

. Although the C14 to C18, saturated or mono-unsaturated fatty acids suit diesel

322 323

Fei, et al. 5b recently reviewed the use of methanotrophs to generate lipids for biodiesel production,

324

thoroughly covering aspects from microbial lipid production to the pitfalls associated with

325

downstream processing. Lipid accumulation exhibits similarities to PHB accumulation, as both may

326

be induced by limiting the oxygen supply, nitrogen or phosphate concentrations to the microbes.

327

Similarly, a two-stage culture process is promoted as the most plausible production method; where

328

the first stage produces healthy biomass and the second stage limits a nutrient source

329

advantage of using a gas for producing a fuel is significant as it avoids the food vs. fuel debate that

330

plagues first generation biofuels. Silverman, et al.

331

methanotrophic/methylotrophic biomass into oil that is refined into a fuel, where the oil is derived

332

from the cell membrane of microorganism. Their preferred production method involves supercritical

333

CO2 extraction of the oil, which may be refined by a cracking, trans-esterification, reforming,

334

distilling, hydro-processing, and/or isomerisation. While they are focusing on using the structural

69

5b

. The

patented the conversion of lipid-containing

ACS Paragon Plus Environment

Page 17 of 51

Environmental Science & Technology

335

lipids, ARPA-E funded research at the University of Washington is focussed on manipulating

336

methanotrophs to produce storage lipids. Here, methanotrophs will be genetically modified to

337

increase their lipid production and enhance the fraction of non-phosphorous-based lipids, thereby

338

aiding trans-esterification into biodiesel.

339 340

3.1.5b. Lipids: human health supplements

341

As the lipids produced by natural methanotrophic isolates are predominantly membrane-derived,

342

they are not ideal for catalytic conversion to biodiesel. However, they may have an alternative

343

higher-value application as a health supplement. There is a current patent for using methanotrophic

344

lipids in the manufacture of an oral administration for use in the treatment of animal subjects to

345

reduce plasma cholesterol levels or lower the ratio of LDL to HDL cholesterol in the plasma. The

346

composition is apparently also useful for increasing docosahexaenoic acid concentration in the

347

plasma, which acts as an immuno-protectant70. The patent is based on research the effects of three

348

different high-lipid diets on plasma lipoproteins and phospholipids in mink (Mustela vison).

349

Phospholipids from natural gas-utilising bacteria in the diet decreased plasma lipoprotein levels, the

350

LDL/HDL cholesterol ratio, and plasma phospholipid levels compared with the highly unsaturated

351

soybean oil. The decrease of plasma cholesterol was attributed to a specific mixture of phospholipids

352

containing a high level of phosphatidyl-ethanolamine, and not the dietary fatty acid composition71.

353

These phospholipids could comprise part of a formulation taken as a health supplement, or

354

potentially used as part of complimentary treatment programme.

355 356

3.1.6. Growth media and vitamin B12

357

Analogous to yeast, beef or potato dextrose extracts, soluble compounds within the methanotrophic

358

biomass may contribute to the bulk of an extract that can provide nutrients for a growth media.

359

There is a patent for a microbial growth medium derived from a microbial culture composed

360

primarily of Methylococcus capsulatus (Bath) and containing Ralstonia sp. DB3 and Brevibacillus agri

ACS Paragon Plus Environment

Environmental Science & Technology

361

DB5 and optionally Aneurinibacillus sp. DB472. The nutrient medium is proposed to consist of a

362

hydrolysate, homogenate or an autolysate of the biomass, with an autolysate being preferred. The

363

addition of further compounds and nutrients such as glucose and/or nitrate and minerals (e.g. K, Ca,

364

Mg, Na, Mb, Fe, Zn, B, Co, Mn and Ni salts) further enhances the scope of this growth medium.

365

Methanotrophs and methylotrophs utilising the serine or RUMP pathways are able to produce

366

vitamin B12 (up to 800 ng/g wet mass using a Methylobacterium species), an essential vitamin to

367

many organisms73.

368 369

3.1.7. Soluble metabolic products: methanol, formaldehyde and organic acids

370

Soluble metabolic intermediates such as methanol, formaldehyde and organic acids are all potential

371

products from methanotrophs with multiple industrial uses and are required in large quantities

372

annually. Viably converting methane to methanol at ambient temperature and pressure is of great

373

interest as methanol is a more easily transportable than methane. Methanol is the first intermediate

374

formed during the conversion of methane to carbon dioxide. In the metabolic pathway (Figure 1) of

375

native bacteria, methanol is rapidly converted to formaldehyde, implying that genetic manipulation

376

or fermentation process control would be vital to ensure accumulation. Tabata and Okura 74 initially

377

failed to produce extracellular methanol from methanotrophs because it was rapidly oxidised

378

internally by methanol dehydrogenase. Extracellular methanol was detected when they selectively

379

inhibited methanol dehydrogenase using cyclopropanol. Han et al. 75 also achieved partial oxidation

380

of methane into methanol through selective inhibition of methanol dehydrogenase by using

381

relatively common media components such as phosphate, NaCl, NH4Cl or EDTA. Alternatively,

382

researchers used a thermophilic methanotroph (Methylocaludum) in a process where methanol was

383

constantly removed from the fermentation. Methanol remained in the gas phase at 50 °C and was

384

successfully condensed from the reactor headspace76. Careful selection of the production strain is

385

also vital for methanol production as production rates and yields from different isolates vary by

386

orders of magnitude. Mehta et al. 77 reported specific production of 1 mmol.hr-1.mg-1 dry mass under

ACS Paragon Plus Environment

Page 18 of 51

Page 19 of 51

Environmental Science & Technology

78

387

continuous conditions, while Xin et al.

388

lower. The low productivity was attributed to the MMO specific activity of their Methylosinus

389

trichosporium strain, which was approximately 1 % of the strain used by used by Mehta et al. 77.

reported methanol production three orders of magnitude

390 391

Low methanol yields are still problematic. Although earlier research had claimed methanol

392

production up to 1 g/l79, very low yields (≤mg/l) are generally reported for methanol (and other

393

soluble metabolites). Recently, 1 g/l was reported using a high-density culture80, but the process

394

required a culture concentrated via centrifugation in 0.4 M phosphate buffer and 20 mM formic

395

acid. This process would be difficult to scale up and even then, the methanol concentration would

396

have to be improved by up to two orders of magnitude.

397 398

Formaldehyde and formic acid represent alternative products formed as intermediates during the

399

mineralisation process, while acetate, lactate, and succinate are produced under oxygen-limited

400

conditions20. Current yields according to literature are very low (mM). While a product such as

401

formaldehyde has a large global demand, it is toxic to cells as it crosslinks proteins (it is often used as

402

a fixing agent because of this quality). This quality may render a cellular and even a purely enzymatic

403

system inactive. One potential alternative is to use small molecule catalysts that mimic the methanol

404

dehydrogenase active site to convert methanol to formaldehyde. It is worth noting that there are a

405

number of companies actively pursuing these compounds. Various companies and research

406

institutions (e.g. Kiverdi, Coskata, CALYSTA Energy, NatureWorks, National Renewable Energy

407

Laboratory, Intrexon) are trying to commercialise biological production of lactate, succinate,

408

muconic acid, butanol and propanol. It will be of interest to see what yields and production

409

efficiencies can be obtained by these researchers, and how significantly production can be improved

410

by bioreactor engineering, genetic engineering and the inclusion of synthetic biology approaches.

411

ACS Paragon Plus Environment

Environmental Science & Technology

412

3.1.8. Methanotroph enzymes, cell-free catalysis and electrochemistry

413

Enzymes produced by methylotrophs are themselves a valuable product. Some dehydrogenases and

414

other enzymes (including glucose-6-phosphate dehydrogenase, glutamate dehydrogenase, malate

415

dehydrogenase, alcohol oxidase and catalase) exhibit high activities34. Enzymes such as diamino-

416

butyrate acetyl-transferase from a halo-tolerant methanotroph have been cloned and expressed in

417

heterologous expression systems; this enzyme catalyses one of the key reactions of biosynthesis of

418

the bacterial osmo-protectant ectoine81.

419 420

Methane may be enzymatically converted to methanol, while the complete enzymatic oxidation of

421

methanol to CO2 has been demonstrated using the various dehydrogenases82. The conversion of

422

methane to methanol is of great industrial interest. Unfortunately, both the particulate and soluble

423

MMO require reducing equivalents, which has limited the potential for a commercial methanol

424

production from these enzymes. The sMMO requires NAD(P)H and O2 catalyses to convert methane

425

to methanol: CH4 + NAD(P)H + H+ + O2 = CH3OH + NAD(P)+ + H2O, while the pMMO requires

426

cytochromes b559/569 or c553 artificial reductants such as duroquinol and NADH83 to complete the

427

reaction. The cost of providing external reducing equivalents for the MMOs renders the economics

428

of the reaction unfeasible for producing a low-value commodity such as methanol. Normally, within

429

the cell the NAD(P)H would be regenerated in a subsequent enzymatic oxidation step. However,

430

developments in NAD(P)H regeneration may be overcome using electrochemical techniques,

431

mediators or combinations with NAD(P) reducing and NAD(P)H-oxidising enzymes.

432 433

Besides the cost associated with the enzyme co-factor /reducing equivalent, it is also difficult to

434

heterologously produce MMOs84. MMO genes have been cloned into bacterial expression systems

435

(which could enable faster and greater enzyme production), because of the relatively slow growth

436

rate of methanotrophs. However, MMOs are complex proteins (consisting of a reductase,

437

hydroxylase and regulatory protein) and a heterologous expression system has proved elusive.

ACS Paragon Plus Environment

Page 20 of 51

Page 21 of 51

Environmental Science & Technology

438

Complete, active MMOs have not been produced thus far, only a partially-active sMMO with a

439

functional hydroxylase22. Active pMMO85 or sMMO86 can be obtained directly from methanotrophic

440

cultures. Although the pMMO is commonly inactive when stringently purified from native bacteria

441

(as it is removed from its lipid matrix and the closely associated proteins and co-factors), active

442

MMO can be purified from methanotrophic biomass87.

443 444

The other three key enzymes in the mineralisation of methanol to CO2 (methanol dehydrogenase,

445

formaldehyde dehydrogenase and formate dehydrogenase) may be used to generate methanol,

446

formaldehyde and formic acid. There is strong interest in enzyme-catalysed redox reactions to

447

produce electricity, fuels as well as chemical commodities from immobilised enzymes on

448

electrodes82, 88. Recently, research has not only focused on forward reaction of CH4 conversion to

449

methanol, but on the reverse process where CO2 is converted to formic acid, formaldehyde and then

450

methanol. Formate dehydrogenase alone catalyses a very useful reaction: converting CO2 into

451

formate. This was achieved by Srikanth et al.

452

demonstrated the ability to electrochemically regenerate the co-factor (NAD+). The three other key

453

enzymes that are involved in the mineralisation of methanol to CO2 have recently been

454

demonstrated capable of the complete reverse enzymatic mineralisation of CH4 displayed in

455

Figure 290. Carbon dioxide was electro-catalytically converted to formate, formaldehyde and then

456

methanol at the cathode. Reactive red was used as an electron mediator and the co-factor (NAD+)

457

was regenerated via cathodic reduction. They further enhanced the process by including carbonic

458

anhydrase, an enzyme that facilitates the hydration of CO2(g). Enzyme-electrochemical techniques

459

are of interest as the ability to transform methane into useful chemicals without the requiring a live

460

culture are highly beneficial, but there are technical, cost and process hurdles to overcome.

89

, in an electrochemical system where they also

461

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 51

462

3.1.9 Propylene oxide production (whole cell catalysis)

463

The broad substrate range of MMOs allow these enzymes to degrade soil contaminants as well as

464

generate products, such as 1- and 2-alcohols from C1-C8 n-alkanes, 1,2-epoxides from terminal

465

alkenes and ethanol/ethanal from diethyl ether91. Alkanes are hydroxylated mostly at the terminal

466

and sub-terminal positions, while ring hydroxylation of aromatics occurs primarily at the meta

467

position. The sMMO oxygenates alkenes to epoxides with retention of stereochemistry around the

468

C=C double bond92. The range for sMMO (n-alkanes, n-alkenes, aromatic and alicyclic compounds) is

469

significantly greater than that of pMMO (n-alkanes and n-alkenes). The broad substrate range of

470

MMOs nearly enabled the commercialisation of propylene oxide production in the 1990s.

471 472

Dalton and colleagues studied the transformation of propylene (propene/methyl-ethylene) to

473

propylene oxide (epoxy-propylene/epoxy-propane) using a methanotroph31,

474

studies to establish the epoxidation rates of propylene to propylene oxide by Methylococcus

475

capsulatus (Bath) were optimised in shake-flask cultures, where adding electron donors such as

476

methanol, formaldehyde, formate or hydrogen stimulated the endogenous rate of propylene oxide

477

formation up to 50 times. Specific production rates as high as 500 mol.min-1.g-1 of cells (dry mass)

478

were obtained with methanol as the electron donor, but were only sustained for short periods of up

479

to 4 minutes94. The loss of the MMO activity and subsequent declining epoxidation rates were

480

studied further95 and subsequently demonstrated as a result of reversible product inhibition96.

481

Reactivation of the bacterial propylene oxidation mechanism could occur without growth, but the

482

process required the presence of an energy source (methane or methanol), sulfur, nitrogen and

483

oxygen. In the presence of growth substrates, cells could be reactivated after the removal of

484

propylene oxide. De novo protein synthesis was also required for reactivation of activity and cultures

485

possessing sMMO took twice as long to recover compared to cells containing pMMO. In the pilot

486

process, methanol was used as the carbon source and MMO inhibition was circumvented by

487

operating a two-stage process system that allowed the epoxide-inhibited culture to recover in a

ACS Paragon Plus Environment

93

. Initial laboratory

Page 23 of 51

Environmental Science & Technology

488

separate bioreactor in the presence of methane and other nutrients. Conducting the

489

biotransformation at the optimal growth temperature for the methanotroph (45 °C) was doubly

490

advantageous, because the boiling point of propylene oxide is 34 °C: allowing simpler product

491

recovery from the gas phase. The process was run at a high cell density (30 g.l−1) and propylene

492

oxide was produced at 250 g.l−1.day−1. The authors not only demonstrated continuous pilot-scale

493

production of propylene oxide from propylene using a methanotroph, but also evaluated the

494

process for producing 1,2-epoxybutane from but-1-ene and acetaldehyde from ethane. The process

495

could unfortunately not be commercialised, as the production cost was already at parity to

496

established commercial chemical technology at the time before inclusions for transport, storage and

497

profit97. More information about propylene oxide production is available in other review articles31, 34.

498 98

499

Xin et al.

500

with methane as the electron-donating co-substrate. They circulated an optimised gas mixture

501

(methane: 35 %; propene: 20 %; oxygen: 45 %) continuously, which removed the product. In this

502

manner they were able to operate a bioreactor continuously for 25 days without any obvious loss of

503

propylene oxide productivity. More recently, Su et al.

504

production using high cell density cultures of Methylosinus trichosporim OB3b and claimed a

505

propylene oxide productivity nearly four-fold greater than the highest reported productivity by

506

optimising temperature, initial propene concentration, sodium formate and MgCl2 concentrations .

recently used a Methylomonas sp. to catalyse the epoxidation in a continuous fashion

99

optimised conditions for propylene oxide

507 508

The combination of greater productivities and the consolidated bioprocessing approach (where the

509

product is generated and removed from a single reactor maintained at a temperature above the

510

products boiling point) is worth investigating again. However, there are new challenges as the

511

traditional methods for propylene oxide production have advanced since the 1990s. The older

512

chlorohydrin process was gradually replaced by the oxidation of isobutane and its catalytic

513

conversion to propylene oxide, which in turn has been superseded by ethylbenzene oxidation, which

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 51

514

also yields styrene as a co-product100. Until recently, commercial propene peroxidation was

515

unfeasible due to the cost of H2O2. BASF and DOW chemicals have commercialised a process that

516

converts O2 into H2O2 (continuously via the oxidation and reduction of 2-ethylanthrahydroquinone),

517

which in conjunction with propene and a catalyst (titanium silicalite-1) produces propylene oxide

518

with no byproducts other than water100-101. This is an elegant process that would be challenging to

519

improve upon. Enzymatic MMO catalysis (as opposed to live cell catalysis) could be worth pursuing,

520

but would require sufficient MMO, a stable enzyme process and cheap regeneration of the reducing

521

equivalents.

522 523

3.1.10. New products and improved efficiency of genetically engineered methanotrophs

524

There has been considerable research over the past three decades regarding the genetics of

525

methanotrophs102. Various methanotrophs have been sequenced6a,

526

genetic engineering of these methanotrophs to over-produce metabolites, or even compounds not

527

naturally synthesised by these bacteria. The proof-of-concept for metabolic engineering of

528

methanotrophs to heterologously synthesise compounds has been validated. Sharpe

529

colleagues genetically modified (GM) methanotrophs to produce high-value carotenoids, while using

530

methane or methanol as a carbon source. Carotenoids are a family of yellow to orange-red

531

terpenoid pigments that protect against oxidative damage and are desired by food, medical, and

532

cosmetic industries

533

associated enzymes to convert it to astaxanthin in a Methylomonas species106. Various C40

534

carotenoids were accumulated in the intracytoplasmic membrane system in high concentrations104.

535

Although astaxanthin is widely used as a feed supplement in poultry and aquaculture industries, it is

536

a challenge to produce in bacteria as astaxanthin generally forms a small percentage of the total

537

carotenoids. A yield of 1 to 2.4 mg.g-1 dry mass was obtained, where 90 % of the total carotenoid

538

was astaxanthin (primarily the E-isomer) by engineering a methanotroph to contain two complete

539

sets of carotenoid biosynthetic genes, proving that astaxanthin with desirable properties could be

105

103

and this has allowed the

104

and

. The research group initially expressed a canthaxanthin gene cluster and

ACS Paragon Plus Environment

Page 25 of 51

Environmental Science & Technology

540

produced in methanotrophs through genetic engineering106. As astaxanthin production was strongly

541

affected by oxygen availability, bacterial haemoglobins were incorporated into the bacteria by

542

further genetic engineering. Co-expression of the haemoglobin and astaxanthin-encoding genes

543

significantly increased astaxanthin expression - as the haemoglobins likely improved the activity of

544

the oxygen-requiring enzymes. A plasmid-free production Methylomonas strain produced more

545

astaxanthin than the parent strain107. This research demonstrated the ability to engineer non-

546

traditional microbial hosts that could use methane or methanol as alternative feedstocks for

547

microbial processes, as well as improve production by metabolic engineering104, 106-108.

548 549

Recently, Intrexon Corporation announced that they had genetically modified a methanotroph to

550

produce farnesene using methane as the carbon source109. Farnesene represents an enormous

551

global commodity as it is a basic precursor for diesel, lubricants and specialty products (cosmetics,

552

rubber and plastics). While this serves as another example of successful genetic engineering, there

553

was no data substantiating any appreciable yield. Moving from initial laboratory proof-of-concept

554

(where specific production may be as below µM.g-1) to pilot and commercial scale may require

555

production to be increased by several orders of magnitude, in addition to overcoming numerous

556

biological and engineering hurdles.

557 558

Methane-oxidising bacteria have served as hosts for producing recombinant and heterologous

559

proteins, including β-glucuronidase and genetically engineered MMOs27c. There are now a number of

560

genetic tools allowing mutagenesis and expression studies with methanotrophs and these

561

techniques allow introduction of broad-host-range plasmids carrying homologous and heterologous

562

genes into methanotrophs, promoter probe fusions, transposon mutagenesis and mutagenesis by

563

marker-exchange27b. There has been a substantial advances in genetic engineering and modification

564

of the capabilities of methanotrophs since the 1980s110, and this will continue to be a challenging,

ACS Paragon Plus Environment

Environmental Science & Technology

565

but rewarding field. Synthetic biology may also contribute greatly to improving the production

566

efficiencies and yields5a.

567 568

3.2 Processes

569

3.2.1. Methane mitigation

570

The remediative abilities of methanotrophs are documented and their potential for mitigating

571

methane emissions from landfills and coal mines and their ability to degrade other hazardous

572

organic compounds have been reviewed6b,

573

technologies have been demonstrated beyond the laboratory scale as adaptable field-scale systems

574

that may be engineered to meet site-specific climatic variations and ensure minimal atmospheric

575

methane emission111, where methane oxidation efficiencies as high as 100 % have been reported18,

576

112

577

passive drainage and biofiltration of landfill gas as a means of managing landfill gas emissions from

578

low to moderate gas generation landfill sites. Passively aerated biofilters operating in a temperate

579

climate achieved maximum methane oxidation efficiencies greater than 90 % and average oxidation

580

efficiencies greater than 50 % over four years of operation. Although temperature and moisture

581

within the biofilter were affected by local climatic conditions, their effect on biofilter performance

582

was overshadowed by landfill gas loading. Interestingly, microbial methane oxidation was limited by

583

outflowing biogas as it prevented diffusion of atmospheric oxygen into the biofilter. Methanotrophic

584

systems have also been combined with algae, thereby sequestering both methane and CO2 and has

585

the potential to generate additional biological products114.

. Dever et al.

113

14b, 14c, 27a, 31

. Methanotrophic methane mitigation

conducted a field scale trial at a landfill site (Sydney, Australia) investigating

586 587

3.2.2 Contaminant bioremediation

588

Methanotrophs are useful bioremediation agents because of the broad substrate range of their

589

MMO enzymes (in particular sMMO), which allows them to remove heavy metals115 and transform

590

organic pollutants116. The sMMO enzymes can transform a variety of hydrocarbons, including

ACS Paragon Plus Environment

Page 26 of 51

Page 27 of 51

Environmental Science & Technology

591

alkanes, alkenes, alicyclic hydrocarbons, aromatic compounds and halogenated aliphatics31,

91, 117

592

Chlorinated compounds degradable by sMMOs include chloroform118, dichloro-ethene 119, trichloro-

593

ethylene118, 120, tetrachloro-ethene121, hydrochloro-fluorocarbons122 and even vinyl-chloride 123.

.

594 595

Methane or nutrients may be added to stimulate methanotrophs and enhance biodegradation and

596

biotransformation of contaminants. Biostimulation of methanotrophs according to the site-specific

597

needs has even been demonstrated at a field scale in situ within contaminated aquifers and soils,

598

and ex situ in bioreactors14b, 25c, 124. Their remediative capacities have also been improved by genetic

599

modification125.

600 601

Plant-microbe associations are important relationships benefiting both partners. Enhanced

602

methanotroph-plant associations may be worth pursuing to create a more stable spread of

603

methanotrophs in a soil environment in a symbiotic relationship with plant roots. Even if the

604

methanotrophs do not greatly benefit the host (as is normally the case with endophytes that provide

605

nutrients or secrete plant growth promoting factors), as long as they are actively present it is

606

environmentally beneficial. Alternatively, transgenic plants can also mobilise or degrade chlorinated

607

solvent, xenobiotic compounds, explosives and phenolic substances. A symbiotic relationship

608

between GM methanotrophs and transgenic plants could significantly enhance the bioremediation

609

of contaminated sites125d.

610 611

3.2.4 Denitrification

612

Biological wastewater denitrification systems require organic carbon to facilitate the reduction of

613

nitrate to nitrogen. This carbon requirement may be partially met by the acidogenic fermentation of

614

a portion of the organic waste entering wastewater treatment systems, which provides volatile fatty

615

acids such as acetate. However, modern wastewater treatment plants frequently need to

616

supplement their systems with a costly external carbon source such as ethanol to achieve more

ACS Paragon Plus Environment

Environmental Science & Technology

617

stringent discharge limits126. Using methane as a low-cost carbon source to facilitate denitrification

618

would be highly beneficial. Incorporating methane into the denitrification process was suggested by

619

various researchers in the 1970s127, but denitrification by pure methanotrophic isolates had still not

620

been validated four decades later128.

621 622

Although methanotrophs can catalyse nitrogen cycle processes such as nitrification and nitrogen

623

fixation, they cannot perform complete denitrification

624

hypothesised that the responsible agent in the mixed methanotrophic culture was a denitrifying

625

methanol-consuming bacteria that was using a methanotrophic byproduct to perform the initial

626

reduction of nitrate to nitrite. The premise was one of syntrophy, where one organism lives off the

627

products of another organism, and this was later verified128. Methane is used by the methanotrophs

628

and they in turn provide an electron donor (such as acetate) for the denitrifying bacteria. As cheaper

629

wastewater denitrification alternatives are being investigated, there has been a recent increase in

630

research published regarding methane oxidation coupled to denitrification129. Various consortia of

631

microorganisms are capable using methane as the sole carbon source for denitrification both

632

aerobically33, 128, 130 and anaerobically33, 129, 131.

128

. Even as early as the 1970s, it was

633 634

3.2.5. Electricity generation in microbial fuel cells

635

In a microbial fuel cell (MFC) micro-organisms are attached to electrodes that harvest the electrical

636

current produced from the spontaneous oxidation of organic substrates132. Theoretically,

637

methanotrophs could be employed in a MFC using methane as the carbon source, thereby coupling

638

biological methane oxidation to electricity generation. The concept of using methane as a carbon

639

source in a microbial fuel cell was patented by Girguis and Reimers133, but, to date, no electrogenic

640

activity has been demonstrated by a methanotroph. However, it is not essential that they are

641

electrogenic as they may serve as an intermediary at the surface of a biofilm and provide organic

642

metabolites to sustain electrogenic bacteria in contact with the anode. The potential

ACS Paragon Plus Environment

Page 28 of 51

Page 29 of 51

Environmental Science & Technology

643

methanotrophic metabolites that have been used as sole carbon sources in MFCs include

644

methanol134 formic acid135, or fermentation products produced under oxygen-limited conditions20

645

such as acetate136 or lactate137. However, there are significant problems associated with using

646

oxygen in the anodic chamber and the materials required to construct MFCs are expensive

647

(considering the poor power densities achieved) making industrial application unlikely.

648 649

3.2.6 Biosensors

650

Biosensors are analytical devices that use a transducer coupled to biological material that elicits a

651

signal in response to an analyte. They can be extremely sensitive as well as highly specific, but may

652

suffer from instability138 or be affected by the external environment. The ability of MMO to react

653

with methane allows the bacteria containing the enzyme, or the isolated enzyme, to be used as the

654

biological component of a biosensor. A Methylomonas culture was exploited early in the

655

development of environmental biosensors by Okada, et al. 139, where methane concentrations were

656

determined in 3 min at 30 °C at pH of 7.2. It allowed for a minimum methane measurement of

657

13 μM and was reproducible to within 5 % for more than 20 days over which more than 500 assays

658

were performed. The bacteria were also incorporated into sensors developed by Daamgard and

659

colleagues140 where oxygen consumption was measured using an internal oxygen amperometric

660

microsensor, acting as a proxy for methane presence. Wen, et al.

661

using a mixed culture of methane-oxidizing bacteria. With both biosensors the sensitivity and the

662

response time were improved by increasing the number of bacteria. Unfortunately, both biosensors

663

were restricted by environmental factors (e.g., pH and temperature) that affected the physiological

664

state of bacteria142. The use of MMO enzymes coupled to an electrode was also investigated, but this

665

was unstable and produced unusual results – presumably due to methane and oxygen retention143.

141

666

ACS Paragon Plus Environment

developed a similar biosensor

Environmental Science & Technology

667

4. Overview of potential products

668

The potential products, their reported yields and production efficiencies are summarised in Table 2,

669

along with their potential global demand, potential nearness to market and relevant references.

670 671

The increasing demand for protein-rich feed by the expanding global aquaculture industry bodes

672

well for single-cell protein production. Further improvements to strains via genetic modification (to

673

generate carotenoids or improved vitamin B12 production) could enhance the nutritional value of

674

the product. Although SCP production appears feasible (as evidenced by the recent construction of a

675

new facility in Trinidad and Tobago), government initiatives/incentives may be necessary to facilitate

676

private investment and further increase this application.

677 678

Propylene oxide and ectoine production are particularly interesting as they represent extracellular

679

products. Unlike most cellular products, this enables cell reuse, which is useful for relatively slow

680

growing biomass. Once the optimal biomass concentration is attained the key constraints are the

681

rates of production and tolerance towards the product inhibition. The commercial value of ectoine

682

makes it a compound worth pursuing via improved downstream processing, new production strains,

683

and genetic engineering to improve yields and yield efficiencies. Propylene oxide has an enormous

684

market and production could be improved via improved enzyme efficiencies and product tolerance,

685

or enhanced simultaneous product removal using either new thermo-tolerant species, or improved

686

production strains engineered to exhibit thermophilic traits. With the near-commercialisation of

687

propylene oxide production by methanotrophs in the 1990s, it is worth revisiting this process,

688

although new production processes will be difficult to compete against.

689 690

Biopolymer production (polyhydroxy-alkanoates and extracellular polysaccharides) and external

691

surface layers may be limited by the production rate of the relatively slow-growing methanotrophs.

692

It can take days to weeks as opposed to hours to days when using conventional bacteria. Although

ACS Paragon Plus Environment

Page 30 of 51

Page 31 of 51

Environmental Science & Technology

693

the price of petroleum precursors will have to increase before biopolymer production becomes

694

viable, a potential bridge towards economic feasibility may be focussing on producing a consistent,

695

high-quality PHB for medical applications. However, for PHAs and EPS, the low solubility of methane

696

is a great challenge, especially with high-density cultures.

697 698

Biodiesel also represents an enormous market, but polar membrane lipids with a high heteroatom

699

content from methanotrophs may not be the best microbial alternative. If attempts to genetically

700

enhance the methanotrophs to improve lipid yields and produce desired storage lipids

701

(triacyglycerols) are successful, it will be an important step towards commercialising biofuels from

702

these bacteria, as the quality of the lipid intermediates have been regarded as playing a major role in

703

the overall fuel production cost and will have a strong impact on the catalytic upgrading steps5b.

704

However, considering the lipid composition in native bacteria, use as a health supplement for

705

lowering cholesterol currently seems most appropriate.

706 707

Products that currently appear unfeasible to produce include farnesene and soluble metabolites

708

such as methanol, formaldehyde and organic acids. There is no data available that indicates that

709

farnesene is produced in any appreciable quantities in methanotrophs. Soluble metabolites such as

710

methanol, formaldehyde and organic acids are produced in low concentrations according to

711

literature. Even methanol titres of 1 g/l will not be commercially viable considering rates and

712

processing conditions; it also negatively impacts culture maintenance as reducing equivalents are

713

not regenerated. However, methanol is an important global commodity and the potential to

714

generate it under ambient conditions using cells modified via traditional genetic manipulation or

715

synthetic biology, or via enzymatic catalysis, is still a prominent topic.

716 717

Although some methanotroph enzymes possess high catalytic activities, it is unlikely that the native

718

methanotroph will be used for production. Cloning the enzymes’ genes into common expression

ACS Paragon Plus Environment

Environmental Science & Technology

719

systems that use conventional carbon sources is more feasible, but this would use a traditional

720

carbon source rather than methane. It is not inconceivable that enzymes could be harvested from

721

spent methanotrophic biomass though, thereby adding value and potentially improving the

722

economic viability of another process.

723 724

Incorporating methanotrophs into biosensors and electricity generation via microbial fuel cells seem

725

unlikely options. The biosensor have a limited functional pH range and require oxygen and reducing

726

equivalents, in addition to general biosensor problems such as background noise, interference and

727

inhibition. Although patented, electricity generation in microbial fuel cells with methane as the sole

728

carbon source has not actually been demonstrated. Electricity generation seems improbable as the

729

presence of oxygen in the anodic chamber will negate electron transfer.

730 731

Regarding processes, partial denitrification aided by methanotrophs could be worth incorporating

732

into wastewater treatment plants where possible, as this could lessen the cost associated with

733

providing an external carbon source for denitrification. Other processes, such as mitigating methane

734

releases into the atmosphere and bioremediation have been demonstrated at a field scale11, 93 and

735

are technically possible, but will only be applied when there is a sufficient financial incentive to limit

736

carbon emissions or remove contaminants. GMOs with improved traits (e.g. tolerance and

737

degradation efficiency) could further enhance these applications, but the perceived of risks related

738

to genetic transfer requires extensive assessment before a GMO may be used in the environment116,

739

125a, 144

.

740 741

5. Methanotroph applications: synthetic biology, scale-up and commercialisation

742

The future of all of the bioprocesses mentioned in this review must yield a profit, or be of greater

743

value when benchmarked against conventional methane applications such as energy recovery or

744

heating, and must be achievable at scale. To ensure a consistent and optimal product or robust

ACS Paragon Plus Environment

Page 32 of 51

Page 33 of 51

Environmental Science & Technology

745

process, fermentation and downstream processing must be optimised, integrated and functional at

746

scale.

747 748

The greatest technical challenge for gas fermentation at scale remains the efficient mass transfer of

749

poorly-soluble gaseous substrates into the aqueous phase145. Other major concerns are the need to

750

regenerate reducing equivalents and the cost of downstream processing. Many reactor types

751

(continuous stirred tank reactors, bubble columns, airlift reactors, trickle beds and numerous

752

variations) have been applied to gas fermentation145, and microbubble generation

753

immobilized hollow fibre membranes146 have been investigated to improve gas transfer efficiency.

754

Increasing the headspace pressure is another means of improving mass transfer. Paraffin and

755

nanoparticle addition can improve gas transfer, but at an increased cost. Cathodic reduction (directly

756

to the cells or an immobilised MMO complex) may provide the electrons required or regenerate the

757

reducing equivalents, but feasibility depends on the cost of the reduction and whether it can be

758

scaled up in an economically-viable manner.

145b

or

759 760

Another significant problem at scale with gas fermentation is foaming (especially with high density

761

cultures) and may be compounded by operating secondary reactors to the extreme of the microbe’s

762

capacity to maintain structural integrity or survive exposure to high product or metabolite

763

concentrations. This may be further aggravated by microaeration, where the small gas bubbles that

764

are required to improve gas-liquid transfer may be more easily stabilised by bacterial lysis products

765

or proteins and glycoproteins.

766 767

However, downstream processing costs may single-handedly prevent the commercialisation of these

768

products. Generally, if the product is intracellular, cells must be concentrated, lysed, and then the

769

product must be extracted, purified, concentrated and processed. There are considerable costs

770

associated with cell pretreatment147, solvent recovery, impurity removal, increases in viscosity,

ACS Paragon Plus Environment

Environmental Science & Technology

771

subsequent requirements for product upgrading or modification, waste treatment and

772

environmental consequences42,

773

increases in liquid viscosity.

51

as well as technical issues dealing with particular solvents and

774 775

The capabilities and yield efficiencies of production strains may be enhanced with cloned genes, or

776

potentially even with completely synthetic additions in the future. There is hope of potentially

777

doubling the energy efficiency of MMOs by engineering a dioxygenase-like enzyme to activate

778

methane, that would allow two methane molecules to be activated for the same energy input5a.

779

Genetic engineering and synthetic biology have a large role in the future application of

780

methanotrophs and their enzymes, as indicated by the strong emphasis on these techniques in many

781

of the ARPA-E funded projects. Some projects involve completely synthetic enzymes for methane

782

activation (Arzeda Corp.), while others will re-engineer enzymes for methylation (Northwestern

783

University, Lawrence Berkeley National Laboratory), pathways (University of California Davis) or

784

methanotrophs (University of California Los Angeles), or even use phototrophic organisms (MOgene

785

Green Chemicals LLC) to produce a liquid fuels such as butanol, methanol, ethanol or dimerise

786

methane. Enzyme use is also being combined with chemical approaches for cell-free catalysis

787

(GreenLight Biosciences) and new metabolic pathways are also being engineered into methylotrophs

788

to convert methanol into butanol (University of Delaware). Synthetic biology may offer alternative

789

routes for regenerating reducing equivalents, using advanced enzymes or new metabolic pathways,

790

or improving yields and conversion efficiencies, but these are complex, time-consuming endeavours

791

and great patience and perseverance will be required.

792 793

There are still serious challenges to commercialising methanotroph applications, but the recent

794

funding allocation and intensive research committed to biologically generating transport fuels from

795

methanotrophs demonstrates the potential envisaged for these unique bacteria. Although they have

796

been known and researched for just over a century, their ability to use methane under ambient

ACS Paragon Plus Environment

Page 34 of 51

Page 35 of 51

Environmental Science & Technology

797

conditions, coupled to the abilities of specialist thermophilic, halophilic and acidophilic

798

methanotrophs, as well as the potential to further improve their capabilities via genetic

799

enhancement and synthetic biology, translates to decades of intriguing research ahead.

800 801

6. Acknowledgements

802

The Centre for Solid Waste Bioprocessing gratefully acknowledges funding from Remondis. W.P.C.

803

and S.H. acknowledge funding from the Australian Research Commission (DP 140104572). The

804

authors acknowledge input from S. Freguia, T. Stewart, M. Patel, J. Bors and I. Pikaar.

805 806

7. References

807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838

1. Kirschke, S.; Bousquet, P.; Ciais, P.; Saunois, M.; Canadell, J. G.; Dlugokencky, E. J.; Bergamaschi, P.; Bergmann, D.; Blake, D. R.; Bruhwiler, L.; Cameron-Smith, P.; Castaldi, S.; Chevallier, F.; Feng, L.; Fraser, A.; Heimann, M.; Hodson, E. L.; Houweling, S.; Josse, B.; Fraser, P. J.; Krummel, P. B.; Lamarque, J. F.; Langenfelds, R. L.; Le Quéré, C.; Naik, V.; O'Doherty, S.; Palmer, P. I.; Pison, I.; Plummer, D.; Poulter, B.; Prinn, R. G.; Rigby, M.; Ringeval, B.; Santini, M.; Schmidt, M.; Shindell, D. T.; Simpson, I. J.; Spahni, R.; Steele, L. P.; Strode, S. A.; Sudo, K.; Szopa, S.; Van Der Werf, G. R.; Voulgarakis, A.; Van Weele, M.; Weiss, R. F.; Williams, J. E.; Zeng, G., Three decades of global methane sources and sinks. Nature Geoscience 2013, 6 (10), 813-823. 2. (a) Bousquet, P.; Ciais, P.; Miller, J. B.; Dlugokencky, E. J.; Hauglustaine, D. A.; Prigent, C.; Van Der Werf, G. R.; Peylin, P.; Brunke, E. G.; Carouge, C.; Langenfelds, R. L.; Lathière, J.; Papa, F.; Ramonet, M.; Schmidt, M.; Steele, L. P.; Tyler, S. C.; White, J., Contribution of anthropogenic and natural sources to atmospheric methane variability. Nature 2006, 443 (7110), 439-443; (b) EPA, Methane and Nitrous Oxide Emissions From Natural Sources. EPA 430-R-10-001 2010. 3. Nwaoha, C.; Wood, D. A., A review of the utilization and monetization of Nigeria's natural gas resources: Current realities. Journal of Natural Gas Science and Engineering 2014, 18 (0), 412432. 4. PSU, CHP Electric Technologies. 2014, http://www.maceac.psu.edu/cleanenergy_chp_technologies.html (accessed 15/05/2014). 5. (a) Conrado, R. J.; Gonzalez, R., Envisioning the bioconversion of methane to liquid fuels. Science 2014, 343 (6171), 621-623; (b) Fei, Q.; Guarnieri, M. T.; Tao, L.; Laurens, L. M. L.; Dowe, N.; Pienkos, P. T., Bioconversion of natural gas to liquid fuel: Opportunities and challenges. Biotechnology Advances 2014, 32 (3), 596-614. 6. (a) Park, D.; Lee, J., Biological conversion of methane to methanol. Korean Journal of Chemical Engineering 2013, 30 (5), 977-987; (b) Hanson, R. S.; Hanson, T. E., Methanotrophic bacteria. Microbiological Reviews 1996, 60 (2), 439-+; (c) Lieberman, R. L.; Rosenzweig, A. C., Biological methane oxidation: Regulation, biochemistry, and active site structure of particulate methane monooxygenase. Critical Reviews in Biochemistry and Molecular Biology 2004, 39 (3), 147164; (d) Scheutz, C.; Kjeldsen, P.; Bogner, J. E.; De Visscher, A.; Gebert, J.; Hilger, H. A.; Huber-Humer, M.; Spokas, K., Microbial methane oxidation processes and technologies for mitigation of landfill gas emissions. Waste Management & Research 2009, 27 (5), 409-455. 7. (a) Zhao, T.; Xing, Z.; Zhang, L., Research progress and discovery process of facultative methanotrophs-A review. Weishengwu Xuebao 2013, 53 (8), 781; (b) Semrau, J. D.; DiSpirito, A. A.;

ACS Paragon Plus Environment

Environmental Science & Technology

839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889

Vuilleumier, S., Facultative methanotrophy: false leads, true results, and suggestions for future research. Fems Microbiology Letters 2011, 323 (1), 1-12; (c) Ralf, C., Microbial Ecology of Methanogens and Methanotrophs. In Advances in Agronomy, Donald, L. S., Ed. Academic Press: 2007; Vol. Volume 96, pp 1-63. 8. (a) Ward, N.; Larsen, Ø.; Sakwa, J.; al, e., Genomic Insights into Methanotrophy: The Complete Genome Sequence of Methylococcus capsulatus (Bath). PLoS Biology 2004, 2 (10), e303; (b) Vorobev, A.; Jagadevan, S.; Jain, S.; Anantharaman, K.; Dick, G. J.; Vuilleumier, S.; Semrau, J. D., Genomic and Transcriptomic Analyses of the Facultative Methanotroph Methylocystis sp. Strain SB2 Grown on Methane or Ethanol. Applied and Environmental Microbiology 2014, 80 (10), 3044-3052. 9. Pol, A.; Heijmans, K.; Harhangi, H. R.; Tedesco, D.; Jetten, M. S.; Op den Camp, H. J., Methanotrophy below pH 1 by a new Verrucomicrobia species. Nature 2007, 450 (7171), 874-8. 10. Boetius, A.; Ravenschlag, K.; Schubert, C. J.; Rickert, D.; Widdel, F.; Gieseke, A.; Amann, R.; Jorgensen, B. B.; Witte, U.; Pfannkuche, O., A marine microbial consortium apparently mediating anaerobic oxidation of methane. Nature 2000, 407 (6804), 623-626. 11. Haroon, M. F.; Hu, S.; Shi, Y.; Imelfort, M.; Keller, J.; Hugenholtz, P.; Yuan, Z.; Tyson, G. W., Anaerobic oxidation of methane coupled to nitrate reduction in a novel archaeal lineage. Nature 2013, 500 (7464), 567-570. 12. Valentine, D. L., Biogeochemistry and microbial ecology of methane oxidation in anoxic environments: a review. Antonie van Leeuwenhoek 2002, 81 (1-4), 271-282. 13. Chistoserdova, L.; Vorholt, J.; Lidstrom, M., A genomic view of methane oxidation by aerobic bacteria and anaerobic archaea. Genome Biology 2005, 6 (2), 208. 14. (a) Dalton, H., The Leeuwenhoek Lecture 2000 - The natural and unnatural history of methane-oxidizing bacteria. Philosophical Transactions of the Royal Society B-Biological Sciences 2005, 360 (1458), 1207-1222; (b) Jiang, H.; Chen, Y.; Jiang, P.; Zhang, C.; Smith, T. J.; Murrell, J. C.; Xing, X.-H., Methanotrophs: Multifunctional bacteria with promising applications in environmental bioengineering. Biochemical Engineering Journal 2010, 49 (3), 277-288; (c) Wendlandt, K.-D.; Stottmeister, U.; Helm, J.; Soltmann, B.; Jechorek, M.; Beck, M., The potential of methane-oxidizing bacteria for applications in environmental biotechnology. Engineering in Life Sciences 2010, 10 (2), 87-102. 15. Semrau, J. D.; Jagadevan, S.; Dispirito, A. A.; Khalifa, A.; Scanlan, J.; Bergman, B. H.; Freemeier, B. C.; Baral, B. S.; Bandow, N. L.; Vorobev, A.; Haft, D. H.; Vuilleumier, S.; Murrell, J. C., Methanobactin and MmoD work in concert to act as the 'copper-switch' in methanotrophs. Environ Microbiol 2013. 16. Murrell, J. C.; Gilbert, B.; McDonald, I. R., Molecular biology and regulation of methane monooxygenase. Archives of microbiology 2000, 173 (5-6), 325-332. 17. (a) Dedysh, S. N.; Liesack, W.; Khmelenina, V. N.; Suzina, N. E.; Trotsenko, Y. A.; Semrau, J. D.; Bares, A. M.; Panikov, N. S.; Tiedje, J. M., Methylocella palustris gen. nov., sp. nov., a new methaneoxidizing acidophilic bacterium from peat bogs, representing a novel subtype of serine-pathway methanotrophs. International Journal of Systematic and Evolutionary Microbiology 2000, 50 (3), 95569; (b) Dedysh, S. N.; Berestovskaya, Y. Y.; Vasylieva, L. V.; Belova, S. E.; Khmelenina, V. N.; Suzina, N. E.; Trotsenko, Y. A.; Liesack, W.; Zavarzin, G. A., Methylocella tundrae sp. nov., a novel methanotrophic bacterium from acidic tundra peatlands. Int J Syst Evol Microbiol 2004, 54 (Pt 1), 151-6; (c) Dunfield, P. F.; Khmelenina, V. N.; Suzina, N. E.; Trotsenko, Y. A.; Dedysh, S. N., Methylocella silvestris sp. nov., a novel methanotroph isolated from an acidic forest cambisol. International Journal of Systematic and Evolutionary Microbiology 2003, 53 (5), 1231-1239. 18. Nikiema, J.; Brzezinski, R.; Heitz, M., Elimination of methane generated from landfills by biofiltration: a review. Reviews in Environmental Science and Bio/Technology 2007, 6 (4), 261-284. 19. Bussmann, I.; Rahalkar, M.; Schink, B., Cultivation of methanotrophic bacteria in opposing gradients of methane and oxygen. FEMS Microbiology Ecology 2006, 56 (3), 331-344. 20. Kalyuzhnaya, M. G.; Yang, S.; Rozova, O. N.; Smalley, N. E.; Clubb, J.; Lamb, A.; Gowda, G. A. N.; Raftery, D.; Fu, Y.; Bringel, F.; Vuilleumier, S.; Beck, D. A. C.; Trotsenko, Y. A.; Khmelenina, V. N.;

ACS Paragon Plus Environment

Page 36 of 51

Page 37 of 51

890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939

Environmental Science & Technology

Lidstrom, M. E., Highly efficient methane biocatalysis revealed in a methanotrophic bacterium. Nature Communications 2013, 4. 21. (a) Dedysh, S. N.; Knief, C.; Dunfield, P. F., Methylocella Species Are Facultatively Methanotrophic. Journal of Bacteriology 2005, 187 (13), 4665-4670; (b) Theisen, A. R.; Murrell, J. C., Facultative Methanotrophs Revisited. Journal of Bacteriology 2005, 187 (13), 4303-4305. 22. Trotsenko, Y. A.; Murrell, J. C., Metabolic aspects of aerobic obligate methanotrophy. Adv Appl Microbiol 2008, 63, 183-229. 23. Auman, A. J.; Speake, C. C.; Lidstrom, M. E., nifH Sequences and Nitrogen Fixation in Type I and Type II Methanotrophs. Applied and Environmental Microbiology 2001, 67 (9), 4009-4016. 24. (a) Semrau, J. D.; DiSpirito, A. A.; Murrell, J. C., Life in the extreme: thermoacidophilic methanotrophy. Trends in Microbiology 2008, 16 (5), 190-193; (b) Trotsenko, Y. A.; Khmelenina, V. N., Biology of extremophilic and extremotolerant methanotrophs. Archives of microbiology 2002, 177 (2), 123-31. 25. (a) Wei, S.-z., Methanotrophs and their applications in environment treatment: A review. Yingyong Shengtai Xuebao 2012, 23 (8), 2309-2318; (b) Chowdhury, T. R.; Dick, R. P., Ecology of aerobic methanotrophs in controlling methane fluxes from wetlands. Applied Soil Ecology 2013, 65 (0), 8-22; (c) Brigmon, R. L., Methanotrophic Bacteria: Use in Bioremediation. In: U.S. Department of Energy, Report No. WSRC-MS-2001-00058 37831–0062. Oak Ridge, TN 2001; (d) Hamer, G., Methanotrophy: From the environment to industry and back. Chemical Engineering Journal 2010, 160 (2), 391-397. 26. (a) Auman, A. J.; Speake, C. C.; Lidstrom, M. E., nifH sequences and nitrogen fixation in type I and type II methanotrophs. Appl Environ Microbiol 2001, 67 (9), 4009-16; (b) Dedysh, S. N.; Ricke, P.; Liesack, W., NifH and NifD phylogenies: an evolutionary basis for understanding nitrogen fixation capabilities of methanotrophic bacteria. Microbiology 2004, 150 (Pt 5), 1301-13; (c) Stein, L. Y.; Klotz, M. G., Nitrifying and denitrifying pathways of methanotrophic bacteria. Biochem Soc Trans 2011, 39 (6), 1826-31. 27. (a) Semrau, J. D.; DiSpirito, A. A.; Yoon, S., Methanotrophs and copper. FEMS Microbiol Rev 2010, 34 (4), 496-531; (b) Murrel, J. C., The Aerobic Methane Oxidizing Bacteria (Methanotrophs). In: Handbook of Hydrocarbon and Lipid Microbiology edited by Timmis K. N. Springer-Verlag Berlin Heidelberg. 2010; (c) Smith, T. J.; Trotsenko, Y. A.; Murrell, J. C., Physiology and Biochemistry of the Aerobic Methane Oxidizing Bacteria in: Kenneth N. Timmis Handbook of Hydrocarbon and Lipid Microbiology Springer-Verlag Berlin Heidelberg 2010. 28. (a) Culpepper, M. A.; Rosenzweig, A. C., Architecture and active site of particulate methane monooxygenase. Crit Rev Biochem Mol Biol 2012, 47 (6), 483-92; (b) Han, B.; Su, T.; Li, X.; Xing, X., Research progresses of methanotrophs and methane monooxygenases. Sheng Wu Gong Cheng Xue Bao 2008, 24 (9), 1511-9; (c) Hakemian, A. S.; Rosenzweig, A. C., The biochemistry of methane oxidation. In Annual Review of Biochemistry, 2007; Vol. 76, pp 223-241; (d) Balasubramanian, R.; Rosenzweig, A. C., Structural and mechanistic insights into methane oxidation by particulate methane monooxygenase. Acc Chem Res 2007, 40 (7), 573-80; (e) Baik, M. H.; Newcomb, M.; Friesner, R. A.; Lippard, S. J., Mechanistic studies on the hydroxylation of methane by methane monooxygenase. Chem Rev 2003, 103 (6), 2385-419; (f) Kopp, D. A.; Lippard, S. J., Soluble methane monooxygenase: activation of dioxygen and methane. Current Opinion in Chemical Biology 2002, 6 (5), 568-576; (g) Merkx, M.; Kopp, D. A.; Sazinsky, M. H.; Blazyk, J. L.; Muller, J.; Lippard, S. J., Dioxygen Activation and Methane Hydroxylation by Soluble Methane Monooxygenase: A Tale of Two Irons and Three Proteins A list of abbreviations can be found in Section 7. Angewandte Chemie (International ed. in English) 2001, 40 (15), 2782-2807. 29. (a) Balasubramanian, R.; Rosenzweig, A. C., Copper methanobactin: a molecule whose time has come. Current Opinion in Chemical Biology 2008, 12 (2), 245-249; (b) Murrell, J. C.; McDonald, I. R.; Gilbert, B., Regulation of expression of methane monooxygenases by copper ions. Trends in Microbiology 8 (5), 221-225.

ACS Paragon Plus Environment

Environmental Science & Technology

940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990

30. Dedysh, S. N.; Dunfield, P. F., Facultative Methane Oxidizers. In Handbook of Hydrocarbon and Lipid Microbiology, Timmis, K., Ed. Springer Berlin Heidelberg: 2010; pp 1967-1976. 31. Smith, T. J.; Dalton, H., Biocatalysis by methane monooxygenase and its implications for the petroleum industry. In: Petroleum Biotechnology – Developments and Perspectives edited by Vazquez-Duhalt, R. and Quintero-Ramirez R. Elsevier Science 151: 177–192 2004. 32. Khosravi-Darani, K.; Mokhtari, Z. B.; Amai, T.; Tanaka, K., Microbial production of poly(hydroxybutyrate) from C-1 carbon sources. Applied Microbiology and Biotechnology 2013, 97 (4), 1407-1424. 33. Modin, O.; Fukushi, K.; Yamamoto, K., Denitrification with methane as external carbon source. Water Res 2007, 41 (12), 2726-38. 34. Trotsenko, I. A.; Doronina, N. V.; Khmelenina, V. N., [Biotechnological potential of methylotrophic bacteria: a review of current status and future prospects]. Prikl Biokhim Mikrobiol 2005, 41 (5), 495-503. 35. Overland, M.; Tauson, A. H.; Shearer, K.; Skrede, A., Evaluation of methane-utilising bacteria products as feed ingredients for monogastric animals. Archives of Animal Nutrition 2010, 64 (3), 171189. 36. Miller, B. M.; Litsky, W., Single Cell Protein in Microbiology. . McGraw-Hill Book Company. 1976, 408. 37. Yazdian, F.; Hajizadeh, S.; Shojaosadati, S. A.; Khalilzadeh, R.; Jahanshahi, M.; Nosrati, M., Production of Single Cell Protein from Natural Gas: Parameter Optimization and RNA Evaluation. Iran J Biotech 2005, 3 (4), 235-42. 38. Bothe, H.; Moller Jensen, K.; Mergel, A.; Larsen, J.; Jorgensen, C.; Bothe, H.; Jorgensen, L., Heterotrophic bacteria growing in association with Methylococcus capsulatus (Bath) in a single cell protein production process. Appl Microbiol Biotechnol 2002, 59 (1), 33-9. 39. (a) Anderson, A. J.; Dawes, E. A., Occurrence, metabolism, metabolic role, and industrial uses of bacterial polyhydroxyalkanoates. Microbiological Reviews 1990, 54 (4), 450-472; (b) Lee, S. Y., Bacterial polyhydroxyalkanoates. Biotechnol Bioeng 1996, 49 (1), 1-14. 40. (a) Tempest, D. W.; Wouters, J. T. M., Properties and performance of microorganisms in chemostat culture. Enzyme and Microbial Technology 1981, 3 (4), 283-290; (b) Jendrossek, D.; Knoke, I.; Habibian, R.; Steinbüchel, A.; Schlegel, H., Degradation of poly(3-hydroxybutyrate), PHB, by bacteria and purification of a novel PHB depolymerase fromComamonas sp. Journal of environmental polymer degradation 1993, 1 (1), 53-63; (c) Karthikeyan, O. P.; Karthigeyan, C. P.; Cirés, S.; Heimann, K., Review of sustainable methane mitigation and bio-polymer production. Critical Reviews in Environmental Science and Technology 2014, 00-00. 41. (a) Asenjo, J. A.; Suk, J. S., Microbial Conversion of Methane into poly-β-hydroxybutyrate (PHB): Growth and intracellular product accumulation in a type II methanotroph. Journal of Fermentation Technology 1986, 64 (4), 271-278; (b) Kim, S. W.; Kim, P.; Kim, J. H., Maximization of poly-beta-hydroxybutyrate accumulation by potassium limitation in Methylobacterium organophilum and its related metabolic analysis. Journal of Microbiology and Biotechnology 1999, 9 (2), 140-146; (c) Kim, P.; Kim, S. W.; Lee, G. M.; Lee, H. S.; Kim, J. H., Isolation and characterization of a methylotroph producing 3-hydroxybutyrate-3-hydroxyvalerate copolymer. Journal of Microbiology and Biotechnology 1995, 5 (3), 167-171; (d) Ostafin, M.; Haber, J.; Doronina, N. V.; Sokolov, A. P.; Trotsenko, Y. A., Methylobacterium extorquens strain P14, a new methylotrophic bacteria producing poly-beta-hydroxybutyrate (PHB). Acta Microbiologica Polonica 1999, 48 (1), 39-51. 42. Choi, J. I.; Lee, S. Y., Process analysis and economic evaluation for poly(3-hydroxybutyrate) production by fermentation. Bioprocess Engineering 1997, 17 (6), 335-342. 43. Lee, S. Y., Bacterial polyhydroxyalkanoates. Biotechnology and Bioengineering 1996, 49 (1), 1-14. 44. Zhang, Y.; Xin, J.; Chen, L.; Song, H.; Xia, C., Biosynthesis of poly-3-hydroxybutyrate with a high molecular weight by methanotroph from methane and methanol. Journal of Natural Gas Chemistry 2008, 17 (1), 103-109.

ACS Paragon Plus Environment

Page 38 of 51

Page 39 of 51

991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041

Environmental Science & Technology

45. Wendlandt, K. D.; Geyer, W.; Mirschel, G.; Hemidi, F. A., Possibilities for controlling a PHB accumulation process using various analytical methods. Journal of Biotechnology 2005, 117 (1), 119129. 46. Shah, N. N.; Hanna, M. L.; Taylor, R. T., Batch cultivation of Methylosinus trichosporium OB3b: V. Characterization of poly-beta-hydroxybutyrate production under methane-dependent growth conditions. Biotechnol Bioeng 1996, 49 (2), 161-71. 47. Zúñiga, C.; Morales, M.; Le Borgne, S.; Revah, S., Production of poly-β-hydroxybutyrate (PHB) by Methylobacterium organophilum isolated from a methanotrophic consortium in a two-phase partition bioreactor. Journal of Hazardous Materials 2011, 190 (1–3), 876-882. 48. Kim, S. W.; Kim, P.; Kim, J. H., PHB accumulation stimulated by ammonium ions in potassium-limited cultures of Methylobacterium organophilum. Journal of Microbiology and Biotechnology 1998, 8 (4), 301-304. 49. Lee, S. Y.; Chang, H. N., Effect of complex nitrogen source on the synthesis and accumulation of poly(3-hydroxybutyric acid) by recombinantEscherichia coli in flask and fed-batch cultures. Journal of environmental polymer degradation 1994, 2 (3), 169-176. 50. Helm, J.; Wendlandt, K. D.; Jechorek, M.; Stottmeister, U., Potassium deficiency results in accumulation of ultra-high molecular weight poly-beta-hydroxybutyrate in a methane-utilizing mixed culture. J Appl Microbiol 2008, 105 (4), 1054-61. 51. Choi, J.; Lee, S. Y., Factors affecting the economics of polyhydroxyalkanoate production by bacterial fermentation. Applied Microbiology and Biotechnology 1999, 51 (1), 13-21. 52. Lee, S. Y.; Park, S. J.; Park, J. P.; Lee, Y.; Lee, S. H., Economic aspects of biopolymer production. Volume 2. Wiley-VCH, Weinheim, Germany 2005. 53. (a) Choi, M. H.; Yoon, S. C.; Lenz, R. W., Production of poly(3-hydroxybutyric acid-co-4hydroxybutyric acid) and poly(4-hydroxybutyric acid) without subsequent degradation by Hydrogenophaga pseudoflava. Appl Environ Microbiol 1999, 65 (4), 1570-7; (b) Pieja, A. J.; Sundstrom, E. R.; Criddle, C. S., Poly-3-Hydroxybutyrate Metabolism in the Type II Methanotroph Methylocystis parvus OBBP. Applied and Environmental Microbiology 2011, 77 (17), 6012-6019. 54. Pouton, C. W.; Akhtar, S., Biosynthetic polyhydroxyalkanoates and their potential in drug delivery. Advanced Drug Delivery Reviews 1996, 18 (2), 133-162. 55. (a) Bonartsev, A. P.; Myshkina, V. L.; Nikolaeva, D. A.; Furina, E. K.; Makhina, T. A.; Livshits, V. A.; Boskhomdzhiev, A. P.; Ivanov, E. A.; Iordanskii, A. L.; Bonartseva, G. A., Biosynthesis, biodegradation, and application of poly(3-hydroxybutyrate) and its copolymers - natural polyesters produced by diazotrophic bacteria. in Communicating Current Research and Educational Topics and Trends in Applied Microbiology. Edited by: A. Méndez-Vilas 2007, 295-307; (b) Brigham, C. J.; Sinskey, A. J., Applications of polyhydroxyalkanoates in the medical industry. International Journal of Biotechnology for Wellness Industries 2012, 1, 53-60. 56. Wu, Q.; Wang, Y.; Chen, G. Q., Medical application of microbial biopolyesters polyhydroxyalkanoates. Artif Cells Blood Substit Immobil Biotechnol 2009, 37 (1), 1-12. 57. Koffas, M.; Odom, J. M.; Wang, S.; Wang, T.; Ye, R. W., Genes encoding exopolysaccharide production. Google Patents: 2003. 58. Malashenko, I. P.; Pirog, T. P.; Romanovskaia, V. A.; Sokolov, I. G.; Gringerg, T. A., Search for methanotrophic producers of exopolysaccharides. Applied Biochemistry and Microbiology 2001, 37 (6), 599–602. 59. (a) Chiemchaisri, W.; Wu, J. S.; Visvanathan, C., Methanotrophic production of extracellular polysaccharide in landfill cover soils. Water Sci Technol 2001, 43 (6), 151-8; (b) Dedysh, S. N.; Khmelenina, V. N.; Suzina, N. E.; Trotsenko, Y. A.; Semrau, J. D.; Liesack, W.; Tiedje, J. M., Methylocapsa acidiphila gen. nov., sp. nov., a novel methane-oxidizing and dinitrogen-fixing acidophilic bacterium from Sphagnum bog. Int J Syst Evol Microbiol 2002, 52 (Pt 1), 251-61. 60. Khmelenina, V. N.; Kalyuzhnaya, M. G.; Sakharovsky, V. G.; Suzina, N. E.; Trotsenko, Y. A.; Gottschalk, G., Osmoadaptation in halophilic and alkaliphilic methanotrophs. Archives of microbiology 1999, 172 (5), 321-329.

ACS Paragon Plus Environment

Environmental Science & Technology

1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092

61. Bunger, J.; Driller, H. J., Protecting, stabilization human skins; enzyme inhibitors, viricides, cosmetics. Google Patents: 2003. 62. Schwibbert, K.; Marin-Sanguino, A.; Bagyan, I.; Heidrich, G.; Lentzen, G.; Seitz, H.; Rampp, M.; Schuster, S. C.; Klenk, H. P.; Pfeiffer, F.; Oesterhelt, D.; Kunte, H. J., A blueprint of ectoine metabolism from the genome of the industrial producer Halomonas elongata DSM 2581 T. Environ Microbiol 2011, 13 (8), 1973-94. 63. Sauer, T.; Galinski, E. A., Bacterial milking: A novel bioprocess for production of compatible solutes. Biotechnology and Bioengineering 1998, 57 (3), 306-313. 64. (a) Egelseer, E. M.; Ilk, N.; Pum, D.; Messner, P.; Schäffer, C.; Schuster, B.; Sleytr, U. B.; Flickinger, M. C., S-Layers, Microbial, Biotechnological Applications. In Encyclopedia of Industrial Biotechnology, John Wiley & Sons, Inc.: 2009; (b) Sleytr, U. B.; Sara, M., Bacterial and archaeal Slayer proteins: structure-function relationships and their biotechnological applications. Trends Biotechnol 1997, 15 (1), 20-6. 65. (a) Shchukin, V. N.; Khmelenina, V. N.; Eshinimayev, B. T.; Suzina, N. E.; Trotsenko, Y. A., Primary characterization of dominant cell surface proteins of halotolerant methanotroph Methylomicrobium alcaliphilum 20Z. Microbiology 2011, 80 (5), 608-618; (b) Khmelenina, V. N.; Suzina, N. E.; Trotsenko, Y. A., Surface layers of methanotrophic bacteria. Microbiology 2013, 82 (5), 529-541. 66. Li, Q.; Du, W.; Liu, D., Perspectives of microbial oils for biodiesel production. Appl Microbiol Biotechnol 2008, 80 (5), 749-56. 67. Sharma, K. K.; Schuhmann, H.; Schenk, P. M., High Lipid Induction in Microalgae for Biodiesel Production. Energies 2012, 5 (5), 1532-1553. 68. (a) Bowman, J. P.; Skerratt, J. H.; Nichols, P. D.; Sly, L. I., Phospholipid fatty acid and lipopolysaccharide fatty acid signature lipids in methane-utilizing bacteria. FEMS Microbiology Letters 1991, 85 (1), 15-22; (b) Fang, J.; Barcelona, M. J.; Semrau, J. D., Characterization of methanotrophic bacteria on the basis of intact phospholipid profiles. FEMS Microbiol Lett 2000, 189 (1), 67-72. 69. Silverman, J.; Resnick, S. M.; Mendez, M. Making fuel e.g. diesel fuel involves converting biomass from culture primarily comprising single carbon-metabolizing non-photosynthetic microorganism into oil composition and refining the oil composition into a fuel. US2014024872-A1, US2014024872-A1 23 Jan 2014 C10G-003/00 201411, 2014. 70. Mueller, H.; Skrede, A.; Kleppe, G., Lipids from methanotrophic bacteria for cholesterol reduction. Patent: WO 2005004888 A1: 2005. 71. Müller, H.; Hellgren, L.; Olsen, E.; Skrede, A., Lipids rich in phosphatidylethanolamine from natural gas-utilizing bacteria reduce plasma cholesterol and classes of phospholipids: A comparison with soybean oil. Lipids 2004, 39 (9), 833-841. 72. Da Moen, E. N.; Jorgensen, J. M. N. D. A. S.; Jensen, K. M. N. D. A. S.; Da Johannessen, A. N., Growth medium for microorganisms comprising the biomass of methanotrophic and heterotrophic bacteria. Patent: CA2481400 C 2006. 73. Ivanova, E. G.; Fedorov, D. N.; Doronina, N. V.; Trotsenko Iu, A., Production of vitamin B12 in aerobic methylotrophic bacteria. Mikrobiologiia 2006, 75 (4), 570-2. 74. Tabata, K.; Okura, I., Hydrogen and Methanol Formation Utilizing Bioprocesses. Journal of the Japan Petroleum Institute 2008, 51 (5), 255-263. 75. Han, J. S.; Ahn, C. M.; Mahanty, B.; Kim, C. G., Partial oxidative conversion of methane to methanol through selective inhibition of methanol dehydrogenase in methanotrophic consortium from landfill cover soil. Appl Biochem Biotechnol 2013, 171 (6), 1487-99. 76. Murakami, Y.; Yamashita, N.; Tsubota, J.; Hasumi, H.; Takeguchi, M.; Ichimura, N.; Sakai, T.; Okubo, T., Industrial methan/methanol conversion using thermophilic methanotrophs. . Abstracts Of The General Meeting Of The American Society For Microbiology 2003, 103, O-092. 77. Mehta, P. K.; Ghose, T. K.; Mishra, S., Methanol biosynthesis by covalently immobilized cells of Methylosinus trichosporium: batch and continuous studies. Biotechnol Bioeng 1991, 37 (6), 551-6.

ACS Paragon Plus Environment

Page 40 of 51

Page 41 of 51

1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142

Environmental Science & Technology

78. Xin, J.-y.; Cui, J.-r.; Niu, J.-z.; Hua, S.-f.; Xia, C.-g.; Li, S.-b.; Zhu, L.-m., Production of methanol from methane by methanotrophic bacteria. Biocatalysis and Biotransformation 2004, 22 (3), 225229. 79. Corder, R. E.; Johnson, E. R.; Vega, J. L.; Clausen, E. C.; Gaddy, J. L., Biological production of methanol from methane. ACS Energy and Fuels Symposium 1988, 1988 Fall (LOS ANGELES) 33(3). 80. Duan, C.; Luo, M.; Xing, X., High-rate conversion of methane to methanol by Methylosinus trichosporium OB3b. Bioresour Technol 2011, 102 (15), 7349-53. 81. Reshetnikov, A. S.; Mustakhimov, II; Khmelenina, V. N.; Trotsenko, Y. A., Cloning, purification, and characterization of diaminobutyrate acetyltransferase from the halotolerant methanotroph Methylomicrobium alcaliphilum 20Z. Biochemistry (Mosc) 2005, 70 (8), 878-83. 82. Dominguez-Benetton, X.; S., S.; Y., S.; K., V.; D., P., Enzymatic Electrosynthesis: An Overview on the Progress in Enzyme-Electrodes for the Production of Electricity, Fuels and Chemicals. Journal of Microbial and Biochemical Technology 2013, doi: 10.4172/1948-5948.S6-007 83. Shiemke, A. K.; Cook, S. A.; Miley, T.; Singleton, P., Detergent solubilization of membranebound methane monooxygenase requires plastoquinol analogs as electron donors. Arch Biochem Biophys 1995, 321 (2), 421-8. 84. Torres Pazmino, D. E.; Winkler, M.; Glieder, A.; Fraaije, M. W., Monooxygenases as biocatalysts: Classification, mechanistic aspects and biotechnological applications. J Biotechnol 2010, 146 (1-2), 9-24. 85. Choi, D. W.; Kunz, R. C.; Boyd, E. S.; Semrau, J. D.; Antholine, W. E.; Han, J. I.; Zahn, J. A.; Boyd, J. M.; de la Mora, A. M.; DiSpirito, A. A., The membrane-associated methane monooxygenase (pMMO) and pMMO-NADH:quinone oxidoreductase complex from Methylococcus capsulatus Bath. J Bacteriol 2003, 185 (19), 5755-64. 86. (a) Park, S.; Hanna, L.; Taylor, R. T.; Droege, M. W., Batch cultivation of Methylosinus trichosporium OB3b. I: Production of soluble methane monooxygenase. Biotechnology and Bioengineering 1991, 38 (4), 423-433; (b) Yu, Y.; Ramsay, J. A.; Ramsay, B. A., Production of soluble methane monooxygenase during growth of Methylosinus trichosporium on methanol. J Biotechnol 2009, 139 (1), 78-83. 87. Chan, S. I.; Nguyen, H. H.; Chen, K. H.; Yu, S. S., Overexpression and purification of the particulate methane monooxygenase from Methylococcus capsulatus (Bath). Methods Enzymol 2011, 495, 177-93. 88. Dalton, H.; Hill, H. A. O.; Kazlauskaite, J.; Wilkins, P. C., Direct electrochemistry of enzymes. Patent WO 1997043632 A1: 1997. 89. Srikanth, S.; Maesen, M.; Dominguez-Benetton, X.; Vanbroekhoven, K.; Pant, D., Enzymatic electrosynthesis of formate through CO sequestration/reduction in a bioelectrochemical system (BES). Bioresour Technol 2014. 90. Addo, P. K.; Arechederra, R. L.; Waheed, A.; Shoemaker, J. D.; Sly, W. S.; Minteer, S. D., Methanol Production via Bioelectrocatalytic Reduction of Carbon Dioxide: Role of Carbonic Anhydrase in Improving Electrode Performance. Electrochemical and Solid-State Letters 2011, 14 (4), E9-E13. 91. Colby, J.; Stirling, D.; Dalton, H., The soluble methane monooxygenase of Methylococcus capsulatus (Bath): its ability to oxygenate n-alkanes, n-alkenes, ethers, and acyclic, aromatic and heterocyclic compounds. Biochemistry Journal 1977, 165, 395-402. 92. Smith, T. J.; Murrell, J. C., Methanotrophs: biotechnological potential and emerging applications. In: Encyclopedia of Industrial Biotechnology. M Flickinger (ed.). New York: Willey. in Press. 2009. 93. (a) Hou, C. T., Vapor phase. 1982; Vol. Patent: U.S. 4,348,476; (b) Suzuki, M.; Dalton, H.; Richards, A. O.; Stanley, S. H., Method for regenerating deactivated microorganisms. Google Patents: 1991; (c) Dalton, H.; Colby, J.; Stirling, D. I., Microbiological process for oxidizing organic compounds. Patent US 4594324 A: 1986; (d) Hou, C. T.; Patel, R.; Laskin, A. I.; Barnabe, N.; Barist, I., Epoxidation

ACS Paragon Plus Environment

Environmental Science & Technology

1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192

of short-chain alkenes by resting-cell suspensions of propane-grown bacteria. Appl Environ Microbiol 1983, 46 (1), 171-7. 94. Stanley, S. H.; Dalton, H., The Biotransformation of Propylene to Propylene Oxide by Methylococcus Capsulatus (Bath): 1. Optimization of Rates. Biocatalysis and Biotransformation 1992, 6 (3), 163-175. 95. Stanley, S. H.; Richards, A. O. L.; Suzuki, M.; Dalton, H., The Biotransformation of Propylene to Propylene Oxide by Methylococcus Capsulatus (Bath): 2. A Study of the Biocatalyst Stability. Biocatalysis and Biotransformation 1992, 6 (3), 177-190. 96. Richards, A. O. L.; Stanley, S. H.; Suzuki, M.; Dalton, H., The Biotransformation of Propylene to Propylene Oxide by Methylococcus capsulatus (Bath): 3. Reactivation of Inactivated Whole Cells to Give a High Productivity System. Biocatalysis and Biotransformation 1994, 8 (4), 253-267. 97. Smith, T. J.; Dalton, H., Chapter 6 Biocatalysis by methane monooxygenase and its implications for the petroleum industry. In Studies in Surface Science and Catalysis, Rafael, V.-D.; Rodolfo, Q.-R., Eds. Elsevier: 2004; Vol. Volume 151, pp 177-192. 98. Xin, J. Y.; Cui, J. R.; Chen, J. B.; Li, S. B.; Xia, C. G., [Continuous biosynthesis of epoxypropane in a methanotrophic attached-films reactor]. Sheng Wu Gong Cheng Xue Bao 2002, 18 (1), 89-93. 99. Su, T.; Han, B.; Yang, C.; Wu, H.; Jiang, H. L.; Li, X.; Xing, X., Production of epoxypropane from propene catalyzed by whole cells of Methylosinus trichosporium OB3b. CIESC Journal 2009, 60 (7), 1767-1772. 100. Nijhuis, T. A.; Makkee, M.; Moulijn, J. A.; Weckhuysen, B. M., The Production of Propene Oxide:  Catalytic Processes and Recent Developments. Industrial & Engineering Chemistry Research 2006, 45 (10), 3447-3459. 101. Company, D. C., Product Safety Assessment: Propylene Oxide. 2013. 102. (a) Zhang, M.; Lidstrom, M. E., Promoters and transcripts for genes involved in methanol oxidation in Methylobacterium extorquens AM1. Microbiology 2003, 149 (4), 1033-1040; (b) Van Dien, S. J.; Okubo, Y.; Hough, M. T.; Korotkova, N.; Taitano, T.; Lidstrom, M. E., Reconstruction of C3 and C4 metabolism in Methylobacterium extorquens AM1 using transposon mutagenesis. Microbiology 2003, 149 (3), 601-609; (c) Toyama, H.; Anthony, C.; Lidstrom, M. E., Construction of insertion and deletion mxa mutants of Methylobacterium extorquens AM1 by electroporation. FEMS microbiology letters 1998, 166 (1), 1-7; (d) Marx, C. J.; Lidstrom, M. E., Development of improved versatile broad-host-range vectors for use in methylotrophs and other Gram-negative bacteria. Microbiology 2001, 147 (8), 2065-2075; (e) Semrau, J.; Chistoserdov, A.; Lebron, J.; Costello, A.; Davagnino, J.; Kenna, E.; Holmes, A.; Finch, R.; Murrell, J.; Lidstrom, M., Particulate methane monooxygenase genes in methanotrophs. Journal of bacteriology 1995, 177 (11), 3071-3079. 103. Costello, A. M.; Lidstrom, M. E., Molecular characterization of functional and phylogenetic genes from natural populations of methanotrophs in lake sediments. Applied and Environmental Microbiology 1999, 65 (11), 5066-5074. 104. Sharpe, P. L., Metabolic Engineering of a Methanotroph for the Production of C40 Carotenoids for Aquaculture Applications. The FASEB Journal 2008, 22, 413.2. 105. Barredo, J.-L., Microbial Carotenoids from Bacteria and Microalgae: Methods and Protocols (Methods in Molecular Biology). Springer Protocols, Humana Press, New York, NY, USA 2012. 106. Ye, R. W.; Yao, H.; Stead, K.; Wang, T.; Tao, L.; Cheng, Q.; Sharpe, P. L.; Suh, W.; Nagel, E.; Arcilla, D.; Dragotta, D.; Miller, E. S., Construction of the astaxanthin biosynthetic pathway in a methanotrophic bacterium Methylomonas sp. strain 16a. J Ind Microbiol Biotechnol 2007, 34 (4), 289-99. 107. Tao, L.; Sedkova, N.; Yao, H.; Ye, R. W.; Sharpe, P. L.; Cheng, Q., Expression of bacterial hemoglobin genes to improve astaxanthin production in a methanotrophic bacterium Methylomonas sp. Appl Microbiol Biotechnol 2007, 74 (3), 625-33. 108. Sharpe, P. L.; Dicosimo, D.; Bosak, M. D.; Knoke, K.; Tao, L.; Cheng, Q.; Ye, R. W., Use of transposon promoter-probe vectors in the metabolic engineering of the obligate methanotroph

ACS Paragon Plus Environment

Page 42 of 51

Page 43 of 51

1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242

Environmental Science & Technology

Methylomonas sp. strain 16a for enhanced C40 carotenoid synthesis. Appl Environ Microbiol 2007, 73 (6), 1721-8. 109. Intrexon, Intrexon's Industrial Products Division Achieves Bioconversion of Methane to Farnesene. Press Release 2014. 110. Lidstrom, M. E.; Wopat, A. E.; Nunn, D. N.; Toukdarian, A. E., Manipulation of Methanotrophs. In Genetic Control of Environmental Pollutants, Omenn, G.; Hollaender, A.; Chakrabarty, A. M.; Levin, M.; Nester, E.; Orians, G.; Wilson, C., Eds. Springer US: 1984; Vol. 28, pp 319-330. 111. (a) Dever, S. A.; Swarbrick, G. E.; Stuetz, R. M., Passive drainage and biofiltration of landfill gas: Australian field trial. Waste Management 2007, 27 (2), 277-286; (b) Dever, S. A.; Swarbrick, G. E.; Stuetz, R. M., Passive drainage and biofiltration of landfill gas: Results of Australian field trial. Waste Management 2011b, 31 (5), 1029-1048; (c) Huber-Humer, M.; Gebert, J.; Hilger, H., Biotic systems to mitigate landfill methane emissions. Waste Management & Research 2008, 26 (1), 33-46. 112. Gebert, J.; Singh, B. K.; Pan, Y.; Bodrossy, L., Activity and structure of methanotrophic communities in landfill cover soils. Environmental Microbiology Reports 2009, 1 (5), 414-423. 113. Dever, S. A.; Swarbrick, G. E.; Stuetz, R. M., Passive drainage and biofiltration of landfill gas: results of Australian field trial. Waste Manag 2011, 31 (5), 1029-48. 114. Heiman, K., Turning Bad Gas Into Good Solids. Queensland Mining and Energy Bulletin 2013, Environment, General Mining, http://www.qmeb.com.au/news/general-mining/turning-bad-gasgood-solids/. 115. Al Hasin, A.; Gurman, S. J.; Murphy, L. M.; Perry, A.; Smith, T. J.; Gardiner, P. H. E., Remediation of Chromium(VI) by a Methane-Oxidizing Bacterium. Environmental Science & Technology 2010, 44 (1), 400-405. 116. Pandey, V. C.; Singh, J. S.; Singh, D. P.; Singh, R. P., Methanotrophs: promising bacteria for environmental remediation. International Journal of Environmental Science and Technology 2014, 11 (1), 241-250. 117. Schuetz, C.; Bogner, J.; Chanton, J.; Blake, D.; Morcet, M.; Kjeldsen, P., Comparative oxidation and net emissions of methane and selected non-methane organic compounds in landfill cover soils. Environ Sci Technol 2003, 37 (22), 5150-8. 118. Alvarez-Cohen, L.; McCarty, P. L., Product toxicity and cometabolic competitive-inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Applied and Environmental Microbiology 1991a, 57, 1031-1037. 119. Janssen, D.; Grobben, G.; Hoekstra, R.; Oldenhuis, R.; Witholt, B., Degradation of trans-1,2dichloroethene by mixed and pure cultures of methanotrophic bacteria. Applied Microbiology and Biotechnology 1988, 29 (4), 392-399. 120. (a) Alvarez-Cohen, L.; McCarty, P. L., Effects of toxicity, aeration, and reductant supply on trichloroethylene transformation by a mixed methanotrophic culture. Appl Environ Microbiol 1991b, 57 (1), 228-35; (b) Henry, S. M.; Grbic-Galic, D., Effect of mineral media on trichloroethylene oxidation by aquifer methanotrophs. Microb Ecol 1990, 20 (1), 151-69; (c) Henry, S. M.; Grbic-Galic, D., Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by methanotrophic cultures from a groundwater aquifer. Appl Environ Microbiol 1991, 57 (1), 236-44; (d) Koh, S. C.; Bowman, J. P.; Sayler, G. S., Soluble Methane Monooxygenase Production and Trichloroethylene Degradation by a Type I Methanotroph, Methylomonas methanica 68-1. Appl Environ Microbiol 1993, 59 (4), 960-7; (e) Smith, K. S.; Costello, A. M.; Lidstrom, M. E., Methane and trichloroethylene oxidation by an estuarine methanotroph, Methylobacter sp. strain BB5.1. Applied and Environmental Microbiology 1997, 63 (11), 4617-4620. 121. (a) Gerritse, J.; Renard, V.; Visser, J.; Gottschal, J. C., Complete degradation of tetrachloroethene by combining anaerobic dechlorinating and aerobic methanotrophic enrichment cultures. Appl Microbiol Biotechnol 1995, 43 (5), 920-8; (b) Lee, S.-W.; Keeney, D. R.; Lim, D.-H.; Dispirito, A. A.; Semrau, J. D., Mixed Pollutant Degradation by Methylosinus trichosporium OB3b

ACS Paragon Plus Environment

Environmental Science & Technology

1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293

Expressing either Soluble or Particulate Methane Monooxygenase: Can the Tortoise Beat the Hare? Applied and Environmental Microbiology 2006, 72 (12), 7503-7509. 122. (a) Chang, W.-k.; Criddle, C., Biotransformation of HCFC-22, HCFC-142b, HCFC-123, and HFC134a by methanotrophic mixed culture MM1. Biodegradation 1995, 6 (1), 1-9; (b) DeFlaun, M. F.; Ensley, B. D.; Steffan, R. J., Biological Oxidation of Hydrochlorofluorocarbons (HCFCs) by a Methanotrophic Bacterium. Nature Biotechnology 1992, 10 (12), 1576-1578. 123. Nelson, Y.; Jewell, W., Vinyl Chloride Biodegradation with Methanotrophic Attached Films. Journal of Environmental Engineering 1993, 119 (5), 890-907. 124. (a) Semprini, L.; Hopkins, G. D.; Grbic-Galic, D.; McCarthy, P. L.; Roberts, P. V., A Laboratory and field evaluation of enhanced In Situ bioremediation of trichloroethylene, cis- and transdichloroethylene, and vinyl chloride by methanotrophic bacteria. In: Bioremediation: Field Experience edited by P.E. Flathman, D.E. Jerger, and J.H. Exner. CRC Boca Raton, FL pp.383-412. 1994; (b) McCarty, P. L.; Semprini, L., Ground-water treatment for chlorinated solvents. . In: Handbook of Bioremediation edited by J.E. Matthews. . Lewis Publishers. Boca Raton, FL. pp. 87-116 1994. 125. (a) Morrissey, J. P.; Walsh, U. F.; O'Donnell, A.; Moenne-Loccoz, Y.; O'Gara, F., Exploitation of genetically modified inoculants for industrial ecology applications. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology 2002, 81 (1-4), 599-606; (b) Villacieros, M.; Whelan, C.; Mackova, M.; Molgaard, J.; Sanchez-Contreras, M.; Lloret, J.; de Carcer, D. A.; Oruezabal, R. I.; Bolanos, L.; Macek, T.; Karlson, U.; Dowling, D. N.; Martin, M.; Rivilla, R., Polychlorinated biphenyl rhizoremediation by Pseudomonas fluorescens F113 derivatives, using a Sinorhizobium meliloti nod system to drive bph gene expression. Applied and Environmental Microbiology 2005, 71 (5), 2687-2694; (c) Liu, S.; Zhang, F.; Chen, J.; Sun, G. X., Arsenic removal from contaminated soil via biovolatilization by genetically engineered bacteria under laboratory conditions. Journal of Environmental Sciences-China 2011, 23 (9), 1544-1550; (d) Azad, M. A.; Amin, L.; Sidik, N. M., Genetically engineered organisms for bioremediation of pollutants in contaminated sites. Chinese Science Bulletin 2014, 59 (8), 703-714. 126. Strong, P. J.; McDonald, B.; Gapes, D. J., Enhancing denitrification using a carbon supplement generated from the wet oxidation of waste activated sludge. Bioresource Technology 2011, 102 (9), 5533-5540. 127. (a) Mason, I., Methane as a Carbon Source in Biological Denitrification. Journal (Water Pollution Control Federation) 1977, 49 (5), 855-857; (b) Davies, T. R., Isolation of bacteria capable of utilizing methane as a hydrogen donor in the process of denitrification. Water Research 1973, 7 (4), 575-579; (c) Harremoes, P.; Henze Christensen, M., Denitrification with methane (Denitrifikation med methan). Vand 1971, 1, 7-11. 128. Knowles, R., Denitrifiers associated with methanotrophs and their potential impact on the nitrogen cycle. Ecological Engineering 2005, 24 (5), 441-446. 129. (a) Sun, F. Y.; Dong, W. Y.; Shao, M. F.; Lv, X. M.; Li, J.; Peng, L. Y.; Wang, H. J., Aerobic methane oxidation coupled to denitrification in a membrane biofilm reactor: Treatment performance and the effect of oxygen ventilation. Bioresource Technology 2013, 145, 2-9; (b) Long, Y.; Zhong, Z. M.; Yin, H.; Lin, Z. Y.; Ye, J. S.; He, B. Y., Characteristic of methane oxidation coupled to denitrification in cover soils of landfill. Transactions of the Chinese Society of Agricultural Engineering 2013, 29 (15), 207-214; (c) Zhu, B.; Sanchez, J.; van Alen, T. A.; Sanabria, J.; Jetten, M. S. M.; Ettwig, K. F.; Kartal, B., Combined anaerobic ammonium and methane oxidation for nitrogen and methane removal. Biochemical Society Transactions 2011, 39, 1822-1825; (d) Liu, J. J.; Sun, F. Q.; Wang, L.; Ju, X.; Wu, W. X.; Chen, Y. X., Molecular characterization of a microbial consortium involved in methane oxidation coupled to denitrification under micro-aerobic conditions. Microbial Biotechnology 2014, 7 (1), 64-76. 130. Costa, C.; Dijkema, C.; Friedrich, M.; García-Encina, P.; Fernández-Polanco, F.; Stams, A. J. M., Denitrification with methane as electron donor in oxygen-limited bioreactors. Applied Microbiology and Biotechnology 2000, 53 (6), 754-762.

ACS Paragon Plus Environment

Page 44 of 51

Page 45 of 51

1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344

Environmental Science & Technology

131. (a) Islas-Lima, S.; Thalasso, F.; Gómez-Hernandez, J., Evidence of anoxic methane oxidation coupled to denitrification. Water Research 2004, 38 (1), 13-16; (b) Raghoebarsing, A. A.; Pol, A.; Van De Pas-Schoonen, K. T.; Smolders, A. J. P.; Ettwig, K. F.; Rijpstra, W. I. C.; Schouten, S.; Sinninghe Damsté, J. S.; Op Den Camp, H. J. M.; Jetten, M. S. M.; Strous, M., A microbial consortium couples anaerobic methane oxidation to denitrification. Nature 2006, 440 (7086), 918-921; (c) Wu, M. L.; van Alen, T. A.; van Donselaar, E. G.; Strous, M.; Jetten, M. S. M.; van Niftrik, L., Co-localization of particulate methane monooxygenase and cd(1) nitrite reductase in the denitrifying methanotroph 'Candidatus Methylomirabilis oxyfera'. Fems Microbiology Letters 2012, 334 (1), 49-56; (d) Hu, Z. Y.; Speth, D. R.; Francoijs, K. J.; Quan, Z. X.; Jetten, M. S. M., Metagenome analysis of a complex community reveals the metabolic blueprint of anannmox bacterium "Candidatus Jettenia asiatica". Frontiers in Microbiology 2012, 3, 9; (e) Luesken, F. A.; van Alen, T. A.; van der Biezen, E.; Frijters, C.; Toonen, G.; Kampman, C.; Hendrickx, T. L. G.; Zeeman, G.; Temmink, H.; Strous, M.; den Camp, H.; Jetten, M. S. M., Diversity and enrichment of nitrite-dependent anaerobic methane oxidizing bacteria from wastewater sludge. Applied Microbiology and Biotechnology 2011, 92 (4), 845-854. 132. (a) Logan, B. E.; Hamelers, B.; Rozendal, R. A.; Schrorder, U.; Keller, J.; Freguia, S.; Aelterman, P.; Verstraete, W.; Rabaey, K., Microbial fuel cells: Methodology and technology. Environmental Science & Technology 2006, 40 (17), 5181-5192; (b) Logan, B. E., Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews Microbiology 2009, 7 (5), 375-381. 133. Girguis, P.; Reimers, C. E., Methane-powered microbial fuel cells. Google Patents: 2011. 134. (a) Montpart, N.; Ribot-Llobet, E.; Garlapati, V. K.; Rago, L.; Baeza, J. A.; Guisasola, A., Methanol opportunities for electricity and hydrogen production in bioelectrochemical systems. International Journal of Hydrogen Energy 2014, 39 (2), 770-777; (b) Liu, B.; Li, B., Single chamber microbial fuel cells (SCMFCs) treating wastewater containing methanol. International Journal of Hydrogen Energy 2014, 39 (5), 2340-2344. 135. Sun, D.; Call, D. F.; Kiely, P. D.; Wang, A.; Logan, B. E., Syntrophic interactions improve power production in formic acid fed MFCs operated with set anode potentials or fixed resistances. Biotechnol Bioeng 2012, 109 (2), 405-14. 136. Omidi, H.; Sathasivan, A., Optimal temperature for microbes in an acetate fed microbial electrolysis cell (MEC). International Biodeterioration & Biodegradation 2013, 85 (0), 688-692. 137. Vasyliv, O. M.; Bilyy, O. I.; Ferensovych, Y. P.; Hnatush, S. O., Application of acetate, lactate, and fumarate as electron donors in microbial fuel cell. In: SPIE Proceedings Vol. 8825: Reliability of Photovoltaic Cells, Modules, Components, and Systems VI edited by Dhere, NG; Wohlgemuth, JH; Lynn, KW 2013. 138. Setford, S.; Newman, J., Enzyme Biosensors. In Microbial Enzymes and Biotransformations, Barredo, J., Ed. Humana Press: 2005; Vol. 17, pp 29-60. 139. Okada, T.; Karube, I.; Suzuki, S., Microbial sensor system which uses Methylomonas sp. for the determination of methane. European Journal of Applied Microbiology and Biotechnology 1981, 12 (2), 102-106. 140. (a) Damgaard, L. R.; Larsen, H.; Revsbech, N. P., Microscale biosensors for environmental monitoring. Trends in Analytical Chemistry 1995, 14 (7), 300-303; (b) Damgaard, L. R.; Nielsen, L. P.; Revsbech, N. P., Methane microprofiles in a sewage biofilm determined with a microscale biosensor. Water Research 2001, 35 (6), 1379-1386; (c) Damgaard, L. R.; Revsbech, N. P., A Microscale Biosensor for Methane Containing Methanotrophic Bacteria and an Internal Oxygen Reservoir. Analytical Chemistry 1997, 69 (13), 2262-2267; (d) Damgaard, L. R.; Revsbech, N. P.; Reichardt, W., Use of an oxygen-insensitive microscale biosensor for methane to measure methane concentration profiles in a rice paddy. Applied and Environmental Microbiology 1998, 64 (3), 864-870. 141. Wen, G.; Zheng, J.; Zhao, C.; Shuang, S.; Dong, C.; Choi, M. M. F., A microbial biosensing system for monitoring methane. Enzyme and Microbial Technology 2008, 43 (3), 257-261. 142. Boulart, C.; Connelly, D. P.; Mowlem, M. C., Sensors and technologies for in situ dissolved methane measurements and their evaluation using Technology Readiness Levels. TrAC Trends in Analytical Chemistry 2010, 29 (2), 186-195.

ACS Paragon Plus Environment

Environmental Science & Technology

1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392

143. Chuang, J. D.; Hemond, H. F., Electrochemistry of soluble methane monooxygenase on a modified gold electrode : implications for chemical sensing in natural waters. Thesis (S.M.) Massachusetts Institute of Technology, Dept. of Civil and Environmental Engineering, 2005. http://hdl.handle.net/1721.1/31156 2005. 144. Singh, J. S., Methanotrophs: the potential biological sink to mitigate the global methane load. Current Science 2011, 100 (1), 29-30. 145. (a) Vega, J. L.; Clausen, E. C.; Gaddy, J. L., Design of bioreactors for coal synthesis gas fermentations. Resources, Conservation and Recycling 1990, 3 (2-3), 149-160; (b) Munasinghe, P. C.; Khanal, S. K., Biomass-derived syngas fermentation into biofuels: Opportunities and challenges. Bioresour Technol 2010, 101 (13), 5013-22. 146. Hickey, R. F.; Tsai, S. P.; Yoon, S. H.; Basu, R.; Tobey, R. E., Submerged membrane supported bioreactor for conversion of syngas components to liquid products. Patent US 8058058 B2: 2011. 147. Jacquel, N.; Lo, C.-W.; Wei, Y.-H.; Wu, H.-S.; Wang, S. S., Isolation and purification of bacterial poly(3-hydroxyalkanoates). Biochemical Engineering Journal 2008, 39 (1), 15-27. 148. Han, B.; Su, T.; Wu, H.; Gou, Z.; Xing, X. H.; Jiang, H.; Chen, Y.; Li, X.; Murrell, J. C., Paraffin oil as a "methane vector" for rapid and high cell density cultivation of Methylosinus trichosporium OB3b. Appl Microbiol Biotechnol 2009, 83 (4), 669-77. 149. NFCSF, Opinion on the safety of BioProtein® by the Scientific Panel on Animal Feed of the Norwegian Scientific Committee for Food Safety Revised version Adopted on the 5th of October 2006 http://www.vkm.no/dav/a0782dea9c.pdf 2006. 150. Wendlandt, K. D.; Geyer, W.; Mirschel, G.; Hemidi, F. A.-H., Possibilities for controlling a PHB accumulation process using various analytical methods. Journal of Biotechnology 2005, 117 (1), 119129. 151. Khmelenina, V. N.; Kalyuzhnaya, M. G.; Sakharovsky, V. G.; Snzina, N. E.; Trotsenko, Y. A.; Gottschalk, G., Osmoadaptation in halophilic and alkaliphilic methanotrophs. Archives of microbiology 1999, 172 (5), 321-329. 152. Taher, E.; Chandran, K., High-rate, high-yield production of methanol by ammonia-oxidizing bacteria. Environ Sci Technol 2013, 47 (7), 3167-73. 153. Eshinimaev, B. T.; Khmelenina, V. N.; Sakharovskii, V. G.; Suzina, N. E.; Trotsenko, Y. A., Physiological, biochemical, and cytological characteristics of a haloalkalitolerant methanotroph grown on methanol. Mikrobiologiya 2002, 71 (5), 596-603. 154. (a) Gretsinger, B. E.; Malashenko, Y. R.; Chernyshenko, D. V. U., patent: USSR Inventor’s Certificate no. 962594. Byull. Izobret., 1982b 1982b, 36, 36-39; (b) Gretsinger, B. E.; Malashenko, Y. R.; Karpenko, V. I.; Grinberg, T. A., Patent: USSR Inventor’s Certificate no. 973869. Byull. Izobret. 1982a, 42, 12-16. 155. Trotsenko, I. A.; Doronina, N. V.; Khmelenina, V. N., Biotechnological potential of methylotrophic bacteria: a review of current status and future prospects. Prikladnaia biokhimiia i mikrobiologiia. 2005, 41 (5), 495-503. 156. (a) Lidstrom, M. E., Metabolic Engineering of Methylotrophic Bacteria for Conversion of Methanol to Higher Value Added Products. EPA Grant Number: R826729 2001; (b) Lidstrom, M. E.; Wopat, A. E., Plasmids in methanotrophic bacteria: isolation, characterization and DNA hybridization analysis. Archives of microbiology 1984, 140 (1), 27-33. 157. Santegoeds, C. M.; Damgaard, L. R.; Hesselink, G.; Zopfi, J.; Lens, P.; Muyzer, G.; de Beer, D., Distribution of sulfate-reducing and methanogenic bacteria in anaerobic aggregates determined by microsensor and molecular analyses. Appl Environ Microbiol 1999, 65 (10), 4618-29.

ACS Paragon Plus Environment

Page 46 of 51

Page 47 of 51

Environmental Science & Technology

1393

8. Figures and Tables

1394 1395

Figure 1: Methanotrophic methane mineralisation or assimilation (via the Serine or RUMP pathways).

1396 1397

ACS Paragon Plus Environment

Environmental Science & Technology

1398 1399

Figure 2: Typical reaction scheme for the trans-esterification of a triglyceride into esters and glycerine.

1400

ACS Paragon Plus Environment

Page 48 of 51

Page 49 of 51

1401

Environmental Science & Technology

Table 1: Generalised differences between Type I and Type II methanotrophs 6, 27a, 40c, 110.

Genera Formaldehyde assimilation pathway Phospholipid fatty acids of intracytoplasmic membranes Membrane arrangement Preferred gas ratio Cyst formation Nitrogen fixation CO2 assimilation

Type I: -Proteobacteria Methylobacter, Methylocaldum, Methylococcus, Methylomicrobium, Methylomonas, Methylosphaera and Methylothermus genera

Type II: -Proteobacteria

Methylocella, Methylocapsa, Methylocystis and Methylosinus genera

Ribulose monophosphate pathway

Serine pathway

Bundles of 16-carbon fatty acids: 14:0, 16:0, 16:1ω7c and 16:1ω5t

Bundles of 18-carbon fatty acids: 18:1ω8c, 18:1ώ8t and 18:1ώ6c

Stacked membranes throughout cytoplasm

Peripheral membrane rings

Lower CH4:Higher O2 Yes Methylococcus, Methylomonas and Methylosphaera 5-15% of biomass

Higher CH4: Lower O2 No Common under low oxygen tension Up to 50 % of biomass

1402

ACS Paragon Plus Environment

Environmental Science & Technology

1403

Table 2: Potential products and processes using methane as the primary carbon source for methanotrophs. Product Relative Global annual demand Near to market value Single cell protein. Low >25 000 tonnes Yes, under defined natural gas costs and product retail price. Commercial plant in Trinidad and Tobago. Internal storage polymers Low Equivalent to 0.6 billion Processing cost constraints. May be viable when (PHB) litres of oil used for petroleum products increase sufficiently in price. making plastics Potential for specialised polymers or via lactic acid production. Lipids: biodiesel Low 1.4 billion litres Not currently feasible due to heteroatoms. R&D needed. Potential via genetic manipulation. Lipids: dietary supplements High Kg - tonnes Possibly. Promising option. Ectoine Soluble metabolic products: Methanol

Very high

1-10 tonnes

Low

90 billion litres

Formaldehyde Organic acids

Low Low

Surface layers Extra-cellular polysaccharides (EPS)

Medium- Unknown High Low >10 tonnes medium

Growth media and Vit B12

Medium

35 billion litres 9600 tonnes (as acetic acid only)

1 tonne

Unknown. Boosts nutrition value.

>1.4 billion litres

Medium

9-10 million tonnes

Unknown. Proof of concept achieved according to press release, but no yield data. Promising, but requires improved efficiencies. Cost was parity with older technology.

None

High High

Denitrification

None low Low

Biosensors Electricity generation

Medium Low

Medium High.

Bioremediation

155 22, 34, 81-90

T: 2.4 mg astaxanthin.g-1

Low medium. Low.

GMO products: Farnesene Biotransformation of propene to propylene oxide. Process Methane mitigation

biomass No data

High

Naturally occurring. Seeded in certain instances and has worked at a field scale. Naturally occurring. Seeded in certain instances and has worked at a field scale. Achieved in laboratories. Technical requirements possibly limits this to niche WWTPs or academia. No. Technical constraints. No. Unproven. Improbable.

1404

ACS Paragon Plus Environment

106

6a, 27b, 27c, 104 110, 156 109

No data −1

−1

SP: 8.3 g.g .day

92

31, 34, 91-93, 98

6b, 14b, 14c, 18, 111-112 14b, 25c, 31, 91, 117, 124-125 33, 127-131

139-143, 157 133-135

Methane as a resource: can the methanotrophs add value?

Methane is an abundant gas used in energy recovery systems, heating, and transport. Methanotrophs are bacteria capable of using methane as their sole ...
820KB Sizes 0 Downloads 10 Views