Impact of Peat Fire on the Soil and Export of Dissolved Organic

May 21, 2018 - Tropical peatlands play an important role in the global carbon cycle, and ... of peat soils in Indonesia for three years following expo...
0 downloads 0 Views 884KB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Impact of peat fire on the soil and the export of dissolved organic carbon in tropical peat soil, Central Kalimantan, Indonesia Kazuto Sazawa, Takatoshi Wakimoto, Masami Fukushima, Yustiawati Yustiawati, M. Suhaemi Syawal, Noriko Hata, Shigeru Taguchi, Shunitz Tanaka, Daisuke Tanaka, and Hideki Kuramitz ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.8b00018 • Publication Date (Web): 21 May 2018 Downloaded from http://pubs.acs.org on May 29, 2018

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 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 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.

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 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

1

Impact of peat fire on the soil and the export of

2

dissolved organic carbon in tropical peat soil,

3

Central Kalimantan, Indonesia

4

Kazuto Sazawa†, Takatoshi Wakimoto†, Masami Fukushima‡, Yustiawati Yustiawati§, M.

5

Suhaemi Syawal‖, Noriko Hata†, Shigeru Taguchi†, Shunitz Tanaka§, Daisuke Tanaka†, and

6

Hideki Kuramitz†,*

7



8

Engineering for Research, University of Toyama, Gofuku 3190, Toyama 930-8555, Japan

9



Department of Environmental Biology and Chemistry, Graduate School of Science and

Laboratory of Chemical Resource, Division of Sustainable Resources Engineering, Faculty of

10

Engineering, Hokkaido University, Sapporo 060-8628, Japan

11

§Division

12

University, Sapporo, Hokkaido 060-0810, Japan

13

‖Research

14

Cibinong, Bogor 16911, Indonesia

of Material Science, Graduate School of Environmental Earth Science, Hokkaido

Center for Limnology, Indonesian Institutes of Sciences, Jl. Raya Jakarta-Bogor Km.46

15

ACS Paragon Plus Environment

1

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 35

16

ABSTRACT: Tropical peatlands play an important role in the global carbon cycle, and therefore,

17

their stability has important implications for climate change. In this study, we evaluated the effect

18

of fire on the physical, chemical, and biological properties of peat soils in Indonesia for three years

19

following exposure to fire. The results of the thermal analysis suggest that the organic matter

20

contents of surface soils significantly decreased because of peat fires and that charred materials

21

were produced in the subsurface layer of the burned soils. The atomic ratios of the burned soils

22

and the thermally treated samples indicate that the Indonesian peat soils were dehydrated by these

23

low-severity fires. The microbial abundance and phosphatase activity in the burned soils

24

significantly decreased compared to those of the unburned soils. Leaching of the dissolved organic

25

carbon (DOC) concentration from the burned soils is lower than that from the unburned soils. The

26

obtained laboratory results indicate that the concentration of the leached DOC increased drastically

27

after heat treatments near the ignition temperature. It was seen that the denaturation of the soil

28

organic matter caused by the heat from the fire accelerates the exodus of organic carbon in

29

peatlands, which contain huge accumulations of carbon.

30

Key words: Peat fire, Soil organic matter, Dissolved organic carbon, Thermogravimetry,

31

Three-dimensional excitation-emission matrix, Soil enzyme

32 33

34

■ INTRODUCTION The total amount of worldwide wetlands consists of only 4%–6% of the global land surface1;

35

however, the top 100 cm of peat soils contains approximately 13%–26% of the world’s soil

36

carbon stock, which is estimated to be 202–377 Gt.2–5 Peatlands, typically found in wetlands of

37

boreal and tropical regions, are formed by the incomplete degradation of root litter, falling leaf

ACS Paragon Plus Environment

2

Page 3 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

38

and branch materials under low microbial activity in anaerobic, acidic and oligotrophic

39

conditions. They play an important role in the global carbon cycle; therefore, their stability has

40

important implications for climate change. The recent increase in global warming is thought to

41

accelerate the decomposition of soil organic matter (SOM).6 Some researchers have emphasized

42

that dissolved organic carbon (DOC) leakage from northern peatlands will increase with

43

increasing river discharge and elevating atmospheric CO2 levels.7,8 The amount of DOC exported

44

from northern peat soil shows a positive correlation with temperature, as well as the lignin

45

decomposition enzyme activity, which also accelerates under more aerated conditions.9–11

46

Indonesian peatlands are estimated to cover an area of 206,950 km2, accounting for

47

approximately 50% of the global tropical peatland area and storing approximately 57.3 Gt of

48

carbon.12 The peatlands of the Sumatra and Central Kalimantan (Borneo) islands, Indonesia have

49

undergone dramatic ecological and social changes over the past decades, and the stability of the

50

peatlands has been disturbed since the early 1980s. In particular, an agricultural land

51

development project in Central Kalimantan, called “The Mega Rice Project,” was promoted from

52

1995 to 1999 by the Indonesian Government, and more than 4500 km of channels were dug in

53

this area. As a result, the drainage from the channels is causing the drying and acidification of

54

the peat soil. Dried peat soil is easy to burn.13, and it can also be assumed that the drying of

55

peatland led to increase in microbial population and soil respiration. This can result in the release

56

of large amounts of CO2 into the atmosphere via wildfires and the decomposition of SOM by

57

microorganisms.14 Furthermore, delays in the onset of the rainy season increase the number of

58

fire events and accelerate damage to peatlands. For example, during the 1997 El Niño event, an

59

extensive area of Indonesian peatland burned (1.45 Mha), approximately 50% of which was on

60

Central Kalimantan. It is estimated that between 0.81 and 2.57 Gt of carbon was released into the

ACS Paragon Plus Environment

3

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

61

atmosphere from these fires across the whole of Indonesia.15 This is equivalent to 13%–40% of

62

the mean annual global carbon emissions from fossil fuels (6.4 Gt-C yr−1). A large number of

63

forest fires have been occurring continuously according to the Fire Information for Resources

64

Management System website.16 Therefore, tropical forests and peatlands in Indonesia are one of

65

the main sources of CO2 emissions arising from land use change anywhere in the world and are

66

important in any mitigation efforts to control current global climate change.

67

Page 4 of 35

Forest fires not only destroy the surface vegetation but also strongly impact the soil via

68

heating. Previously, several excellent field and laboratory studies have investigated the effect of

69

forest fires on the physical, chemical, and biological properties of soil.17–23 The deposition of

70

peat soil requires long time periods, and the current average accumulation rate for tropical peat

71

soil in Indonesia has been estimated to be between 0.04 and 2.55 mm yr−1.24 Therefore, there has

72

been concern that the soil ecosystems of peatlands in Indonesia are strongly influenced by the

73

rapid loss of peat by fire events. The magnitude and the recovery term of these alterations

74

depend on the fire intensity (temperature), the duration of the fire, and the soil texture.22, 23

75

The aim of the present study was to evaluate the physical (particle density and combustion

76

characteristics), chemical (quantity and quality of organic matter), and biological (enzyme

77

activity and microbial population) properties of unburned and burned tropical peat collected from

78

Central Kalimantan, Indonesia, over three years. In addition, the effects of peat fire on the

79

component fractions and fluorescence properties of the water-soluble organic matter (WSOM) in

80

the peat soil were observed. The WSOM is strongly linked to the storage of carbon in catchment

81

soil and is quite vulnerable to changes in the environment. However, the observation of the effect

82

of peat fires on WSOM has not been reported. To examine the impact of the heating process, the

ACS Paragon Plus Environment

4

Page 5 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

83

denaturation of SOM and WSOM in thermally treated samples was evaluated, and these data

84

were compared to field samples.

85

■ MATERIALS AND METHODS

86

Study Sites and Sampling. The map of the sampling sites of Palangka Raya on Kalimantan

87

Island, Indonesia, is shown in Figure 1. Soil samples were collected from unburned sites (UB1–

88

3) and burned sites (B1–5) on September 24 and 25 in 2010, September 17 and 19 in 2011, and

89

September 11 in 2012. The soil samples were collected from a depth of 30–50 cm (subsurface) in

90

2010 and from two soil (surface: 0–20 cm and subsurface: 30–50 cm) in 2011 and 2012. The

91

UB1 and UB2 were a relatively intact swamp forest with little drainage.25 Dominant tree species

92

included Combretocarpus rotundatus, Cratoxylum arborescens, Buchanania sessifolia, and

93

Tetramerista glabra.26 The UB3 was located in secondary forest. Dominant tree species and the

94

microtopography of the forest floor resembled those at the UB1 and UB2.25 The B1–5 sites were

95

a drained burnt swamp forest. Forest fire were reported at B1–4 in 1997, 2002, 2004 and 2009.

96

Fern (Stenochlaena, Blechnum, and Lygodium spp.) and sedge (Cyperus, Scleria and Eleocharis

97

spp.) plants were sparsely re-growing at the B1–4. The B5 was burnt in 2011 and surface plants

98

were lost. According to some previous reports, the ground water level for the UB1 and UB3 in

99

2005–2009 were from 0 to -1.0 m and from -0.5 to -1.5 m, respectively.25 Prior to use, each

100

sample was air-dried at room temperature and ground to pass a 2-mm mesh. To assay the

101

enzyme activities, 20-g samples were placed in glass vials, and 4 mL of toluene was added at the

102

sampling site; these samples were stored at 4°C and were assayed within two weeks of sampling.

ACS Paragon Plus Environment

5

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 35

103 104 105

Figure 1. The location of the sampling sites of Palangka Raya on Kalimantan, Indonesia. UB:

106

Unburned sites, B: Burned sites.

107

Physicochemical Analysis. The pH (H2O), electrical conductivity (EC), and pH (KCl) of the

108

soils were measured in slurries (1:10 air-dried soil/distilled water or 1M KCl). The

109

concentrations of cations in the slurries were determined using an Advanced Compact IC 761

110

(Metrohm Lt., Herisau, Switzerland) after filtration through a 0.45-µm pore diameter membrane

111

filter (mixed cellulose ester, ADVANTEC, Tokyo, Japan).27 The moisture content was

112

determined from the weight loss after drying at 105°C for 48 h.28 The particle density was

113

determined based on Japanese Industrial Standards (JIS A 1202).29 In addition, these samples

114

were combusted at 600°C for 2 h to determine the organic content from the weight loss.

ACS Paragon Plus Environment

6

Page 7 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

115

Thermal Analysis. The combustion characteristics of the soil were determined via a

116

Thermogravimetry-Differential Thermal Analysis (Thermo plus TG 8120, Rigaku, Tokyo,

117

Japan). Approximately 10 mg of samples was heated from 20°C to 500°C at a rate of 3°C min−1.

118

Alumina was used as the reference standard.

119

Thermal Treatments. In the laboratory, the sample collected from UB1 (2011) at the

120

subsurface layer was heated to different temperatures (90°C, 120°C, 150°C, 200°C, 250°C,

121

350°C, and 480°C) for 1, 5, 30, 60, and 120 min using a muffle furnace (KDF007Ex, Denken,

122

Tokyo, Japan). The heating rate used was 3°C min−1. After cooling down, the thermally treated

123

samples were dehydrated under reduced pressure before the analysis was performed.

124

Elemental Analysis. The analyses of the elemental compositions of the soils and thermally

125

treated samples were carried out at the Center for Instrumental Analysis at Hokkaido University.

126

The C, H, and N contents were measured via a Micro Corder JM 10 type CHN analyzer (J-

127

Science Lab. Co. Ltd., Kyoto, Japan). The sulfate ions were analyzed using DX-500 type ion

128

chromatography (Dionex, Sunnyvale, CA). The percentage of oxygen was determined by

129

subtracting the total sum percentage of C, H, N, S, and ash elements from 100.

130

Py-TMAH-GC/MS. Soil samples (1 mg) were placed in a deactivated stainless steel cup. A

131

25-μL aliquot of tetramethylammonium hydroxide (TMAH) in methanol (40 mg mL−1) and a 10-

132

μL aliquot of nonadecanoic acid in acetone (0.06 mg mL−1) as an internal standard (ISTD) were

133

then added to the cup. After removing the solvents under reduced pressure, the cup was

134

introduced into a PY-2020D type Double-Shot Pyrolyzer (Frontier Laboratories Ltd.,

135

Fukushima, Japan) connected to a Shimadzu GC-17A/QP5050 type GC/MS system.30

ACS Paragon Plus Environment

7

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

136

Page 8 of 35

Microbial DNA Extracted from Soils and Real-Time PCR Assay. Soil microbial DNA was

137

extracted using the MO BIO Ultra Clean Soil DNA isolation kit (MO BIO Lab., Solana Beach,

138

CA, USA) according to the manufacturer’s protocol. The abundance of bacteria in the peat soils

139

was estimated via real-time PCR, whose amplifications were carried out in a PCR tube (200 μL,

140

Hi-8-Tube, Takara Bio, Shiga, Japan) with a total volume of 25 μL using a thermal cycler

141

(Takara Bio, Shiga, Japan). Each 25-μL reaction contained the following: 2.0 μL of the DNA

142

template, 1.0 μL of the primers 341f and 518r (10 μM),31 12.5 μL of the two-fold SYBR® Premix

143

Ex Taq, and 8.5 μL of sterile distilled water. The PCR conditions were as follows: initial

144

denaturation at 95°C for 30 s, then 40 cycles with denaturation at 95°C for 5 s, and annealing at

145

60°C for 30 s. The effects of fire on the soil microbial populations in the Indonesian peat soil

146

were evaluated from the comparative cycle threshold (CT) values of the unburned and burned

147

soils.

148

Soil Enzyme Assays. This study measured the activities of one oxidase (phenol oxidase (PO;

149

EC 1.14.18.1) and eight hydrolases (β-glucosidase (β-Glu; EC 3.2.1.21), β-xylosidase (β-Xyl;

150

EC 3.2.1.37), β-galactosidase (β-Gal; EC 3.2.1.23), α-mannosidase (α-Man; EC 3.2.1.24), N-

151

acetyl-glucosaminidase (NAG; EC 3.2.1.30), acid phosphatase (AcP; EC 3.1.3.2), alkaline

152

phosphatase (AlP; EC 3.1.3.1), and phosphodiesterase (PD; EC 3.1.4.1)). A summary of the

153

assay conditions is listed in Table S1. All substrates were obtained from Sigma Aldrich (St

154

Louis, MO). The PO activity method used 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid)

155

diammonium salt (ABTS+) as substrate.32 The results were expressed as mM ABTS+ g dry-soil−1

156

min−1 (with the coefficient of the extinction value of ABTS+ ε420 = 18460 M−1 cm−1). The

157

enzyme activities of β-Glu, β-Xyl, β-Gal, α-Man, NAG, AcP, AlP, and PD were determined on

158

field-moist soil using modifications of the standard methods.33–35 The enzyme activities were

ACS Paragon Plus Environment

8

Page 9 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

159

determined from colorimetric measurements of p-nitrophenol (p-NP) released when the soil was

160

incubated (30°C, 1 h) in the optimum buffer solution containing the substrate. The results were

161

expressed as mM p-NP g dry-soil−1 h−1. The difference in the enzyme activities of the unburned

162

and burned soils was analyzed using Spearman nonparametric correlations; P values of 0.05 or

163

less were considered to be significant.

164

The Properties of DOC in Water-Extracted Solutions of the Field and Thermally

165

Treated Soil Samples. The soil water-extracted solutions were obtained by shaking the 1:10 air-

166

dried soil/distilled water solutions at 190 rpm for 24 h at 25°C in the dark and then by filtering

167

through a 0.45-µm pore diameter membrane filter. The total concentrations of DOC in the soil

168

water-extracted solutions were measured via a TOC analyzer (TOC-5000A, Shimadzu, Kyoto,

169

Japan).27 The DOC components were fractionated using the SupeliteTM DAX-8 resin (40 ± 60

170

mesh, Supelco, Bellefonte, PA, USA). A total of 6 mL of DAX-8 resin was transferred into

171

columns (12 mL, 1 cm × 15 cm, FLEX-column, Kontes, Vineland, NJ) in slurry. The soil water-

172

extracted solutions were adjusted to pH < 2 with 2 M HCl, and 15 mL of the samples was

173

filtered by gravity through the DAX-8 resin with a flow rate of less than 1 mL min−1. The

174

effluent contained hydrophilic (Hp: e.g., fatty acids, sugar acids, hydroxyl acids, and

175

polysaccharides) and hydrophobic (HoB: aromatic amines) base fractions (Hp + HoB).36 The

176

hydrophobic neutral (HoN: e.g., large cellulose polymers, hydrocarbons, and carbonyl

177

compounds) and hydrophobic acid (HoA: humic and fulvic acids) fractions were retained in the

178

resin. The HoA fraction was extracted via 0.1 M NaOH at 1 bed volume with a constant flow

179

rate of 0.5 mL min−1. The Hp + HoB and HoA solutions were adjusted to pH < 2 with 2 M HCl.

180

The DOC concentrations of the Hp + HoB and HoA fractions were measured via a TOC

ACS Paragon Plus Environment

9

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

181

analyzer, and the HoN fraction was calculated by subtracting the Hp + HoB and HoA fractions

182

from the total DOC.

183

Page 10 of 35

To determine the molecular weight of DOC in the soil water-extracted solution, a 20-μL

184

aliquot was injected into a Jasco PU-2080 plus Intelligent HPLC pump system equipped with a

185

UV-2075 UV/vis detector (Japan Spectroscopic Co., Tokyo, Japan).37 The absorbance value at

186

280 nm for the soil water-extracted solution was recorded to calculate E280, which correlated

187

strongly with the total aromaticity of the dissolved humic substance.38 The absorptivity at 280

188

nm was calculated as E (cm−1 g of C−1) = absorbance value/[DOC (g L−1)] × (%C/100).

189

The 3DEEM spectrum of the soil water-extracted solutions was measured using a fluorescence

190

spectrophotometer (Mode LS-55, Perkin Elmer, CA, USA). The scanning wave ranges were

191

200–600 nm for both excitation (Ex) and emission (Em).27 The relative fluorescence intensity

192

(RFI) was calibrated to be evaluated in quinine sulfate units (QSU), 1 QSU = 1 µg L−1, of

193

quinine sulfate monohydrate in the solution of 0.05 M H2SO4 at Ex/Em = 355/450 nm.

194

■ RESULTS AND DISCUSSION

195

The Effects of Fire on the Physicochemical Properties. The means of the selected

196

physicochemical properties of the soils are shown in Table S2. Compared to the UB sites, the

197

surface soil sampled immediately after being burned (B5, 2011) indicated higher pH (UB1–3 =

198

3.04–3.46; B5 in 2011 = 5.81) and EC (UB1–3 = 0.35–2.70 mS cm−1; B5 in 2011 = 4.34 mS

199

cm−1). It is a well-known fact that the cations (K+, Na+, Ca2+, and Mg2+) released from

200

combusting SOM can increase soil pH by displacing H+ and Al3+ ions adsorbed on the negative

201

charge of the soil colloids.39 The concentration of alkaline cations in the water-extracted solution

202

obtained from B5 (2011) at the surface was significantly higher than those at the UB sites at the

ACS Paragon Plus Environment

10

Page 11 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

203

subsurface (UB1–3: K+ = 0.40–4.43 mg L−1, Na+ = 0.64–6.54 mg L−1, Ca2+ = 0.09–0.73 mg L−1,

204

and Mg2+ = 0.16–0.72 mg L−1; B5 in 2011: K+ = 14.0 mg L−1, Na+ = 6.91 mg L−1, Ca2+ = 14.9

205

mg L−1, and Mg2+ = 25.1 mg L−1). Soil pH and concentration of alkaline cations have been

206

reported to increase upon heating at temperatures above 350°C.20 Conversely, one year after the

207

peat fire, the concentration of alkaline cations in B5 at the surface indicated pre-fire levels (K+ =

208

0.54 mg L−1, Na+ = 1.21 mg L−1, Ca2+ = 0.54 mg L−1, and Mg2+ = 0.96 mg L−1), which suggests

209

that the alkaline cations were leached from the burned soil during the wet season.

210

In the case of the subsurface layer, the pH and EC values did not change because of the fire. It

211

has often been reported that the changes in the soil properties after fire events depend on the

212

temperature reached at different soil depths. In fact, a previous field study has reported that the

213

temperature of Indonesian peat soils in the 0–5 cm layer increased to 400°C during forest fires,

214

while the temperature of the 10–20 cm soil layer did not reach 100°C.40 However, in this study,

215

the subsurface burned soils had greater soil particle density (UB = 0.97 ± 0.15 g cm−3, B = 1.16 ±

216

0.16 g cm−3, and P = 0.010) and lower ignition loss (UB = 80.5 ± 2.5%, B = 76.9 ± 4.8%, and P

217

= 0.025) than did the unburned soils. Rein et al. (2017) investigated the vertical downward

218

spread of smouldering fire in column of 30 cm tall moss peat under variable moisture content.

219

They found that the downward spread rate of fire continuously increases as peat moisture content

220

increases.41 Therefore, the soil aggregates of the subsurface soils in this study sites were affected

221

by heating during peat fire.

222

The Effects of Peat Fire on the Soil Organic Matter Properties. Thermogravimetry-

223

Differential Thermal Analysis (TG-DTA) curves of the soil collected from both layers of B5 and

224

the subsurfaces of UB1, UB3 and B3 in 2011 are shown in Figure 2. Both curves show the

225

weight loss and the rate of heat released from the soil sample during the pyrolysis process. The

ACS Paragon Plus Environment

11

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 35

226

results of the TG curves indicate that the amount of SOM in the surface soil immediately after a

227

fire decreases significantly, whereas the subsurface layer SOM is not changed (Figure 2(a)).

228

From the obtained DTA curve for UB1, the ignition temperature (Tv), moisture release (Peak A),

229

combustible gas release (Peak B), and the combustion of carbide (Peak C) were determined to be

230

at 185.5°C (Tv), 54.4°C, 306.1°C, and 425.4°C, respectively. Compared to UB1 and UB3, the

231

Peak B and Peak C values of the B5 surface decreased significantly (Figure 2(b)), and in case of

232

the subsurface layer, the peak B value decreased dramatically, whereas the peak C value was

233

56% higher than that for UB1. These results suggest that the surface layer of B5 was heated

234

strongly at high temperature above 400°C and production of charred materials occurred in the

235

burned subsurface soils. Furthermore, the DTA curves of B3 and B5 did not show much

236

difference. This indicates that the thermal product remained in the subsurface burned soils over

237

two years after the fire.

238 239 240

Figure 2. (a) Weight loss (TG) and (b) differential thermal analysis (DTA) curves of the

241

Indonesian peat soils collected from UB1, UB3 (subsurface), B3 (subsurface), and the site

242

sampled immediately after a peat fire (B5 at the surface and subsurface). Tv: ignition temperature

ACS Paragon Plus Environment

12

Page 13 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

243

of the volatile matter, Peak A: moisture release, Peak B: combustible gas release, and Peak C:

244

combustion of carbide.

245

The elemental composition and atomic ratio (H/C, O/C and C/N) of the UB soil, B soil and the

246

thermally treated samples shown in Table S3. The C/N ratio of thermally treated samples

247

decreases with increasing heating temperature (50 for 90°C for 60 min, 39 for 200°C for 60 min,

248

and 21 for 480°C for 60 min). Conversely, the C/N ratios of subsurface peat soil in the burned

249

sites have a tendency to increase compared with those of unburned soils (UB = 42.2 ± 7.0, B =

250

81.4 ± 17, and P < 0.001). Several soil samples from burned sites exhibited an intriguingly high

251

ratio of C/N of 100, creating serious concerns with respect to the restoration potential of the peat

252

soil productivity. In addition, during the three years of observation, not much change was noticed

253

in the C/N ratio. In general, the C/N ratios of post-fire soils and thermally treated samples

254

(>350°C) are lower than those of unburned soils.19,42 This is due to the heat-induced formation of

255

large amounts of condensed structures, including heterocyclic nitrogen forms. The findings of

256

this study assume that some of the burned soils were affected by low-severity fires.

257

The van Krevelen plot (the H/C versus O/C ratios) of the Indonesian peat soil collected from

258

the unburned and burned sites at the subsurface in the period of 2010–2012 and the thermally

259

treated samples is shown in Figure 3. The atomic ratios of H/C and O/C obtained from the

260

samples that were heated to >190°C for 60 min did not show any change; however, in the

261

samples exposed to 200°C and 250°C, the atomic ratios of H/C and O/C decreased when the

262

heating time was increased. The H/C and O/C ratios in the samples exposed to 350°C or 480°C

263

for 60 min were similar. In addition, the H/C and O/C ratios of the burned soils were lower than

264

those of the unburned soils. The solid and dashed lines in Figure 3 indicate the accelerating

265

dehydration and decarboxylation reactions. These results reveal that some of the burned soils

ACS Paragon Plus Environment

13

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 35

266

were dehydrated by low-severity fires, which were near the ignition temperature (>200°C), and

267

that carbonized soils remain for over three years after a peat fire. According to a previous study,

268

the heat released in a fire is transported faster and penetrates deeper in moist soils than in dry

269

soils.43 That is, the organic matter in high-moisture soils such as peatlands is more strongly

270

influenced by wildfires than that in other types of soils.

271 272

Figure 3. The van Krevelen plot (the H/C versus O/C ratios) of the Indonesian peat soils

273

collected from unburned and burned sites at the subsurface in 2011 and 2012 and thermally

274

treated samples. The solid and dashed lines indicate the accelerating dehydration and

275

decarboxylation reactions. The accelerating dehydration reaction indicates that the number of H

276

and O atoms decreased by a ratio of 2 to 1 from the H and O values of UB1 in 2011.

277

The total ion chromatograms for the Py-TMAH-GC/MS and the fractions of groups for

278

pyrolysate compounds collected from the field and thermally treated samples are shown in

279

Figures S1, S2, and 4. Table S4 lists the major pyrolysate compounds identified in the

ACS Paragon Plus Environment

14

Page 15 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

280

chromatograms, which can be classified into the following major groups: lignin-derived

281

compounds (cinnamyl, guaiacyl, and syringyl structures), non-lignin-derived compounds, and

282

fatty acids. The results from the burned soils at two depths show that the relative abundance

283

values of lignin-, non-lignin-, and fatty-acid-derived compounds were decreased by the fire.

284

Similar phenomena were observed in the thermally treated samples. A previous study reported

285

that the peak of the pyrolysis products was significantly decreased by wildfires.19 They

286

concluded that charred “non-pyrolysable” refractory carbonaceous materials formed during the

287

heating. It is known that short-chain fatty acids ( 0.05 level. Variance components

380

significant at P < 0.05 are indicated in bold type. Asterisks represent statistical significance, * P < 0.05, ** P < 0.01, and *** P < 0.001.

381

N.D. = Not determined.

Surface

UB B

PO

β-Glu

β-Xyl

β-Gal

α-Man

NAG

AcP

AlP

PD

2011

1.08 ± 0.5

2.02 ± 0.8

0.21 ± 0.1

0.41 ± 0.07

0.06 ± 0.03

0.65 ± 0.07

28.9 ± 3.8

17.3 ± 1.1

6.92 ± 1.6

2012

0.24 ± 0.1

0.84 ± 0.01

0.45 ± 0.2

0.19 ± 0.02

0.05 ± 0.07

0.39 ± 0.03

11.5 ± 6.1

8.71 ± 2.8

2.61 ± 0.9

2011

0.19 ± 0.1

0.22 ± 0.2

0.22 ± 0.03

0.11 ± 0.08

0.11 ± 0.04

0.18 ± 0.07

0.69 ± 0.2

0.49 ± 0.1

0.17 ± 0.08

2012

0.09 ± 0.03

0.16 ± 0.1

0.22 ± 0.01

0.15 ± 0.03

0.03 ± 0.02

0.30 ± 0.07

0.66 ± 0.2

0.23 ± 0.1

0.44 ± 0.06

***

***

0.0002***

P Subsurface

UB

B

P

0.008

**

0.020

*

0.185

0.020

*

0.815

0.0025

**

0.0002

0.0002

2010

0.10 ± 0.04

0.99 ± 0.2

0.58 ± 0.08

0.39 ± 0.2

N.D.

0.46 ± 0.1

23.2 ± 1.7

7.96 ± 0.6

1.91 ± 0.2

2011

0.42 ± 0.02

0.68 ± 0.3

0.68 ± 0.2

0.06 ± 0.05

0.18 ± 0.08

0.44 ± 0.2

15.1 ± 4.5

9.14 ± 2.6

2.74 ± 0.8

2012

0.20 ± 0.01

0.04 ± 0.03

0.78 ± 0.1

0.35 ± 0.07

0.07 ± 0.05

0.44 ± 0.09

6.20 ± 0.8

5.55 ± 2.5

2.17 ± 0.06

2010

0.09 ± 0.01

0.71 ± 0.1

0.64 ± 0.1

0.16 ± 0.2

0.17 ± 0.1

0.44 ± 0.1

4.22 ± 1.7

4.62 ± 1.1

0.39 ± 0.2

2011

0.22 ± 0.08

0.04 ± 0.04

0.86 ± 0.5

N.D.

0.33 ± 0.1

0.13 ± 0.1

1.66 ± 1.0

1.41 ± 0.4

0.97 ± 0.7

2012

0.50 ± 0.2

0.60 ± 0.2

0.98 ± 0.2

0.24 ± 0.1

1.13 ± 0.1

0.62 ± 0.09

2.39 ± 0.9

2.13 ± 1.5

3.38 ± 1.2

1.000

0.285

0.644

0.285

0.012*

0.667

190°C), the concentration of DOC decreased

405

with increasing heating temperature. Observations of the SOM properties indicate that the

ACS Paragon Plus Environment

22

Page 23 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

406

Indonesian soils were heated by low-severity fires (>200°C). Therefore, it is clear that the low

407

concentrations of DOC in the burned sites occurred because of the denaturation and leaching of

408

soluble SOM.

409

The average molecular weight (Mw) and E280 of the DOC extracted from the thermally treated

410

samples are shown in Figure 6(c). The Mw values decreased when the samples were heated over

411

90°C, and E280 significantly increased for thermal treatments of 150°C and 175°C for 60 min.

412

The fluorescence properties of the colored dissolved organic matter (CDOM) in the soil water-

413

extracted solutions of the UB1 and thermally treated samples were evaluated using the 3DEEM

414

spectrum (SI, Figure S3). The visible humic-like peak (Ex/Em = 320/445 nm) was detected in all

415

of the water samples.63 The relationship between the heating temperature and the RFI of the visible

416

humic-like/DOC ratio and excitation wavelength of the visible humic-like peak for the water-

417

extracted solutions from the thermally treated samples are shown in Figure 6(d). In the case of the

418

soil samples exposed to temperatures of 175°C, the RFI value of the visible humic-like/DOC was

419

15-fold higher than that of the UB1. In addition, a redshift in the excitation wavelength of the

420

visible humic-like peak occurred for sample heating at 150°C and 175°C for 60 min (Ex/Em = 285–

421

280/445 nm). The peak of the UV humic-like materials was detected at Ex/Em = 260/400–460 nm.63

422

A previous study reported that UV humic-like materials have fewer functional groups (less

423

conjugated electron resonance systems) than do visible humic-like materials.63 The investigations

424

conducted both in the field and in the laboratory reveal that peat fires cause transformations of

425

water-soluble SOM to lower molecular, higher aromaticity, and fewer functional group organic

426

compounds. Therefore, the exodus of CDOM (UV humic-like compounds) in the surface peatland

427

could have occurred during rainfall events after the fire.

428

ACS Paragon Plus Environment

23

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 35

429 430

Figure 6. The DOC fractions and concentrations of the water-extracted solutions from (a)

431

unburned sites (UB1–3), burned sites (B1–5) at the subsurface in 2011, and (b) thermally treated

432

samples. Symbols: ■ = Hydrophobic neutral fraction (HoN), ■ = hydrophobic acid fraction (HoA),

433

■ = hydrophilic and hydrophobic base fractions (Hp + HoB), and ● = ignition loss. (c) Average

434

molecular weight (Mw; ●) and E280 (●) of the water-extracted solutions in the thermally treated

435

samples (90°C, 120°C, 150°C, 175°C, 180°C, 190°C, 200°C, 250°C, and 480°C for 60 min). (d)

436

The relationship between the heating temperature and the RFI of the visible humic-like/DOC ratio

437

(●) and excitation wavelength of the visible humic-like peak (●) for the water-extracted solutions

438

from the thermally treated samples.

439

ACS Paragon Plus Environment

24

Page 25 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

440

ACS Earth and Space Chemistry

■ CONCLUSIONS

441

This study evaluated the effect of forest fire on the soil physical, chemical, biological properties

442

and the export of DOC from tropical peat land in Central Kalimantan, Indonesia. The organic

443

components in peat surface layer was destroyed by forest fires, and the flame-resistance

444

characteristics of SOM was produced in the subsurface layer of the burned soils. In addtion, the

445

decrease of microbial abundances and enzyme activities were observed in the subsurface layer of

446

the burned soils. The heat released in a fire is transported faster and penetrates deeper in moist

447

soils than in dry soils. Therefore, the organic matter in high-moisture soils such as peatlands is

448

strongly influenced by wildfires. The field and laboratory studies observed that the low-severity

449

fires accelerates the exodus of low molecular weight CDOM from peatlands. The structural

450

changes in DOC consequently will change the bioavailavility of metal and nutrient in the

451

hydrosphere. This study suggested that the denaturation of SOM and DOC strongly affect the

452

carbon cycles in Indonesian peat land, which have a huge accumulation of carbon storage in the

453

world.

454

455

■ ASSOCIATED CONTENT

456

Supporting Information. Supporting information for the assay conditions of the soil enzymes in

457

this study (Table S1), detailed physicochemical properties of the Indonesian peat soils (Table

458

S2), the assignments for the peaks identified in the Py-TMAH-GC/MS analyses of the peat soils

459

and the pyrogram total ion chromatograms from the Indonesian peat soil and thermally treated

ACS Paragon Plus Environment

25

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

460

samples (Table S3, Figure S1, and Figure S2), and the 3DEEM fluorescence spectrum of the

461

water-extracted samples in UB1 for the subsurface and thermally treated samples (Figure S3).

462

■ AUTHOR INFORMATION

463

Corresponding Author

464

* Phone & fax: +81 (0)76 445 6669; e-mail: [email protected].

465

■ ACKNOWLEDGMENT

466

Page 26 of 35

This work was supported by the JST/JICA Science and Technology Research Partnership for

467

Sustainable Development (SATREPS) Project entitled “Wild Fire and Carbon Management in

468

Peat-forest in Indonesia” and the Japan Society for the Promotion of Science (JSPS) via a Grant-

469

in-Aid for Scientific Research (80727016).

470

471

472

■ REFERENCES

473

(1) Aselmann, I.; Crutzen, P.J., Global distribution of natural freshwater wetlands and rice

474

paddies, their net primary productivity, seasonality, and possible methane emissions. J. Atmos.

475

Chem. 1989, 8, 307-358.

476

(2) Sjörs, H., Peat on Earth: multiple use of conservation. Ambio 1980, 9, 303-308.

ACS Paragon Plus Environment

26

Page 27 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

477

(3) Adams, J. M.; Faure, H.; Faure-Denard, L.; McGlade, J. M.; Woodward, F.I., Increases in

478

terrestrial carbon storage from the Last Glacial maximum to the present. Nature 1990, 348, 711-

479

714.

480 481 482 483 484 485 486 487

(4) Eswaran, H.; Van den Berg, E.; Reich, P., Organic carbon in soils of the world. Soil Sci. Soc. Am. J. 1993, 57, 192-194. (5) Batjes, N. H., Total carbon and nitrogen in the soils of the world. Eur. J. Soil Sci. 1996, 47, 151-163. (6) Bellamy, P. H.; Loveland, P. J.; Bradley, R. I.; Lark, R. M.; Kirk, G. J. D., Carbon losses from all soils across England and Wales 1978–2003. Nature 2005, 437, 245-248. (7) Tranvik, L. J.; Jansson, M., Climate change-Terrestrial export of organic carbon. Nature 2002, 415, 861-862.

488

(8) Freeman, C.; Fenner, N.; Ostle, N. J.; Kang, H.; Dowrick, D. J.; Reynolds, B.; Lock, M.

489

A.; Sleep, D.; Hughes, S.; Hudson, J., Export of dissolved organic carbon from peatlands under

490

elevated carbon dioxide levels. Nature 2004, 430, 195-198.

491 492 493 494

(9) Freeman, C.; Ostle, N.; Kang, H., An enzymic ‘latch’ on a global carbon store. Nature 2001, 409, 149. (10) Freeman, C.; Evans, C.; Monteith, D.; Reynolds, B.; Fenner, N., Export of organic carbon from peat soils. Nature 2001, 412, 785.

ACS Paragon Plus Environment

27

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

495

Page 28 of 35

(11) Pastor, J.; Solin, J.; Bridgham, S. D.; Updegraff, K.; Harth, C.; Weishampel, P.; Dewey,

496

B., Global warming and the export of dissolved organic carbon from boreal peatlands. Okios

497

2003, 100, 380-386.

498 499 500

(12) Page, S. E.; Rieley, J. O.; Banks, C. J., Global and regional importance of the tropical peatland carbon pool. Global Change Biol. 2011, 17, 798-818. (13) Putra, E. I.; Hayasaka, H., The effect of the precipitation pattern of the dry season on peat

501

fire occurrence in the Mega Rice Project area, Central Kalimantan, Indonesia. Tropics 2011, 19,

502

145-156.

503

(14) Hirano, T.; Segah, H.; Harada, T.; Limin, S.; June, T.; Hirata, R.; Osaki, M., Carbon

504

dioxide balance of a tropical peat swamp forest in Kalimantan, Indonesia. GCB Bioenergy 2007,

505

13, 412-425.

506

(15) Page, S. E.; Siegert, F.; Rieley, J. O.; Boehm, H. -D. V.; Jayak A.; Limink, S., The

507

amount of carbon released from peat and forest fires in Indonesia during 1997. Nature 2002, 420,

508

61-68.

509 510 511

(16)

Fire

Information

for

Resources

Management

System

Website;

https://earthdata.nasa.gov/earth-observation-data/near-real-time/firms (17) Neary, D. G.; Klopatek, C. C.; DeBano, L. F.; Fgolliott, P. F., Fire effects on

512

belowground sustainability: a review and synthesis. For. Ecol. Manage. 1999, 122, 51-71.

513

(18) DeBano, L. F., The role of fire and soil heating on water repellency in wildland

514

environments: a review. J. Hydrol. 2000, 231-232, 195-206.

ACS Paragon Plus Environment

28

Page 29 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

515 516

(19) Gonzalez-Perez, J. A.; Gonzalez-Vila, F. J.; Almendros G.; Knicker, H., The effect of fire on soil organic matter-a review. Environ. Int. 2004, 30, 855-870.

517 518 519 520 521 522 523 524 525

(20) Certini, G., Effects of fire on properties of forest soils: a review. Oecologia 2005, 143, 110. (21) Knicker, H., How does fire affect the nature and stability of soil organic nitrogen and carbon? A review. Biogeochemistry 2007, 85, 91-118. (22) Mataix-Solera, J.; Cerda, A.; Arcenegui, V.; Jordan, A.; Zavala, L. M., Fire effects on soil aggregation: A review. Earth-Sci. Rev. 2011, 109, 44-60. (23) Almendros G.; Gonzalez-Vila, F. J., Wildfires, soil carbon balance and resilient organic matter in Mediterranean ecosystems. A review. Spanish J. Soil Sci. 2012, 2, 8-33. (24) Page, S. E.; Wust, R. A. J.; Weiss, D.; Rieley, J. O.; Shotyk, W.; Limin, S. H., A record

526

of Late Pleistocene and Holocene carbon accumulation and climate change from an equatorial

527

peat bog (Kalimantan, Indonesia): implications for past, present and future carbon dynamics. J.

528

Quaternary. Sci. 2004, 19, 625-635.

529

(25) Hirano, T.; Segah, H.; Kusin, K.; Limin, S.; Takahashi, H.; Osaki, M., Effects of

530

disturbances on the carbon balance of tropical peat swamp forests. Glob. Change Biol. 2012, 18,

531

3410-3422.

532

(26) Tuah, S.J.; Jamal, Y.M.; Limin, S., Nutritional characteristics in leaves of plants native to

533

tropical peat swamps and heath forests of Central Kalimantan, Indonesia. TROPICS 2003, 12,

534

221-245.

ACS Paragon Plus Environment

29

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

535

Page 30 of 35

(27) Sazawa, K; Tachi, M.; Wakimoto, T.; Kawakami, T.; Hata, N.; Taguchi, S.; Kuramitz, H.,

536

The evaluation for alterations of DOM components from upstream to downstream flow of rivers

537

in Toyama (Japan) using three-dimensional excitation-emission matrix fluorescence

538

spectroscopy. Int. J. Environ. Res. Public Health 2011, 8, 1655-1670.

539

(28) Takakai, F.; Morishita, T.; Hashidoko, Y.; Darung, U.; Kuramochi, K.; Dohong, S.;

540

Limin, S.H.; Hatano, R., Effects of agricultural land-use change and forest fire on N2O emission

541

from tropical peatlands, Central Kalimantan, Indonesia. Soil Sci. Plant Nutr. 2006, 52, 662-674.

542

(29) Japanese Industrial Standards Committee (1999) Test method for density of soil particles

543

JIS A 1202, Japanese Standards Association (in Japanese).

544

(30) Fukushima, M.; Yamamoto, M.; Komai, T.; Yamamoto, K., Studies of structural

545

alterations of humic acids from conifer bark residue during composition by pyrolysis-gas

546

chromatography/mass spectrometry using tetramethylammonium hydroxide (TMAH-py-

547

GC/MS). J. Anal. Appl. Pyrolysis 2009, 86, 200-206.

548

(31) Muyzer, G.; de Waal, E. C.; Uitterlinden, A. G., Profiling of complex microbial

549

populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-

550

amplified genes coding for 16S rRNA. Appl. Environ. Microbiol. 1993, 59, 695-700.

551 552 553

(32) Floch, C.; Alarcon-Gutierrez, E.; Criquet, S., ABTS assay of phenol oxidase activity in soil. J. Microbiol. Methods 2007, 71, 319-324. (33) Tabatai, M. A.; Bremner J. M., Use of p-nitrophenyl phosphate for assay of soil

554

phosphatase activity. Soil Biol. Biochem. 1969, 1, 301-307.(34) Eivazi, F; Tabatabai, M. A.,

555

Phosphatases in soils. Soil Biol. Biochem. 1977, 9, 167-172.

ACS Paragon Plus Environment

30

Page 31 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

556 557

ACS Earth and Space Chemistry

(35) Eivazi, F; Tabatabai, M. A., Glucosidases and galactosidases in soils. Soil Biol. Biochem. 1988, 20, 601-606.

558

(36) Leenheer, J. A., Comprehensive approach to preparative isolation and fractionation of

559

dissolved organic carbon from natural waters and wastewaters. Environ. Sci. Technol. 1981, 15,

560

578-587.

561

(37) Kuramitz, H.; Sazawa, K.; Nanayama, Y.; Hata, N.; Taguchi, S.; Sugawara, K.;

562

Fukushima, M., Electrochemical genotoxicity assay based on a SOS/umu test using

563

hydrodynamic voltammetry in a droplet. Sensors 2012, 12, 17414-17432.

564 565 566 567 568

(38) Peuravuori, J.; Pihlaja, K., Molecular size distribution and spectroscopic properties of aquatic humic substances. Anal. Chim. Acta 1997, 337, 133-149. (39) Arocena, J. M.; Opio, C., Prescribed fire-induced changes in properties of sub-boreal forest soils. 2003, Geoderma 113, 1-16. (40) Usup, A.; Hashimoto, Y.; Takahashi, H.; Hayasaka, H., Combustion and thermal

569

characteristics of peat fire in tropical peatland in Central Kalimantan, Indonesia. Tropics 2004,

570

14, 1-19.

571 572

(41) Huang, X.; Rein, G., Downward spread of smouldering peat fire: the role of moisture, density and oxygen supply. Int. J. Wildland Fire 2017, 26, 907-918.

573

(42) Almendros, G.; Knicker, H.; Gonza´lez-Vila, F. J., Rearrangement of carbon and nitrogen

574

forms in peat after progressive isothermal heating as determined by solid-state 13C- and 15N-NMR

575

spectroscopies. Org. Geochem. 2003, 34, 1559-1568.

ACS Paragon Plus Environment

31

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 35

576

(43) Choromanska, U.; DeLuca, T. H., Microbial activity and nitrogen mineralization in forest

577

mineral soils following heating: evaluation of post-fire effects. Soil Biol. Biochem. 2002, 34, 263-

578

271.

579 580 581 582 583 584 585 586

(44) Almendros, G.; Martın, F.; Gonza´lez-Vila, F. J., Effects of fire on humic and lipid fractions in a Dystric Xerochrept in Spain. Geoderma 1988, 42, 115-127. (45) Fritze, H.; Pennanen, T.; Pietikainen, J., Recovery of soil microbial biomass and activity from prescribed burning. Can. J. For. Res. 1993, 23, 1286-1290. (46) Warcup, M.G.; Horne, D.J., Occurrence of dormant ascospores in soil. Nature 1963, 197, 1317-1318. (47) Eivasi, F.; Bayan, M. R., Effects of long-term prescribed burning on the activity of selected soil enzymes in an oak-hickory forest. Can. J. Forest Res. 1996, 26, 1799-1804.

587

(48) Hernandez, T.; Garcia, C.; Reinhardt, I., Short-term effect of wildfire on the chemical,

588

biochemical and microbiological properties of Mediterranean pine forest soils. Biol. Fertil. Soils

589

1997, 25, 109-116.

590

(49) Ajwa, H. A.; Dell, C. J.; Rice, C. W., Changes in enzyme activities and microbial biomass

591

of tallgrass prairie soil as related to burning and nitrogen fertilization. Soil Biol. Biochem. 1999,

592

31, 769-777.

593 594

(50) Boerner, R. E. J.; Brinkman, J. A., Fire frequency and soil enzyme activity in southern Ohio oak-hickory forests. Appl. Soil Eco. 2003, 23, 137-146.

ACS Paragon Plus Environment

32

Page 33 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

595

ACS Earth and Space Chemistry

(51) Gutknecht, J. L. M.; Henry, H. A. L.; Balser, T. C., Inter-annual variation in soil extra-

596

cellular enzyme activity in response to simulated global change and fire disturbance.

597

Pedobiologia 2010, 53, 283-293.

598

(52) Sinsabaugh, R. L.; Antibus, R. K.; Linkins, A. E.; Rayburn, L.; Repert, D.; Weiland, T.,

599

Wood decomposition in a first order watershed: mass loss as a function of exoenzyme activity.

600

Soil Biol. Biochem. 1992, 24, 743-749.

601

(53) Sinsabaugh, R. L.; Antibus, R. K.; Linkins, A. E.; McClaugherty, C. A.; Rayburn, L.;

602

Weiland, T., Wood decomposition: nitrogen and phosphorus dynamics in relation to

603

extracellular enzyme activity. Ecology 1993, 74, 1586-1593.

604 605 606 607 608 609 610

(54) Speir, T. W.; Cowling, J. C., Phosphatase activities of pasture plants and soils: relationship with plant productivity and soil P fertility indices. Biol. Fert. Soils 1991, 12, 189-194. (55) Dinkelaker, B.; Marschner, H., In vivo demonstration of acid phosphatase activity in the rhizosphere of soil-grown plants. Plant Soil 1992, 144, 199-205. (56) Nakas, J. P.; Gould, W. D.; Klein, D. A., Origin and expression of phosphatase activity in a semi-arid grassland soil. Soil Biol. Biochem. 1987, 19, 13-18. (57) Tarafdar, J. C.; Claassen, N., Organic phosphorus compounds as a phosphorus source for

611

higher plants through the activity of phosphatase produced by plant roots and microorganisms.

612

Biol. Fert. Soils 1988, 5, 308-312.

ACS Paragon Plus Environment

33

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 35

613

(58) Saa, A.; Trasar-Cepeda, M. C.; Gil-Sotres, F.; Carballas, T., Changes in soil phosphorus

614

and acid phosphatase activity immediately following forest fires. Soil Biol. Biochem. 1993, 25,

615

1223-1230.

616

(59) Zhang, Y.-M.; Wu, N.; Zhou, G.-Y.; Bao, W.-K., Changes in enzyme activities of spruce

617

(Picea balfouriana) forest soil as related to burning in the eastern Qinghai-Tibetan Plateau. Appl.

618

Soil Eco. 2005, 30, 215-225.

619 620 621 622

(60) Tabatabai, M. A.; Bremner, J. M., Arylsulfatase activity of soils. Soil Sci. Soc. Am. J. 1970, 34, 225-229. (61) Skujins, J. J., Enzymes in soil. In Soil Biochemistry Vol.1; McLaren, A. D., Peterson, G. H., Eds.; Dekker, New York. 1967; pp 371-414.

623

(62) Turner, B. L.; Haygarth, P. M., Phosphatase activity in temperate pasture soil: Potential

624

regulation of labile organic phosphorus turnover by phosphodiesterase activity. Sci. Total Environ.

625

2005, 344, 27-36.

626 627

(63) Coble, P.G.; Del Castillo, C. E.; Avril, B., Distribution and optical properties of CDOM in the Arabian Sea during the 1995 Southwest Monsoon. Deep-Sea Res. II 1998, 45, 2195-2223.

628

ACS Paragon Plus Environment

34

Page 35 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

629 630

ACS Earth and Space Chemistry

Table of Content (TOC)

631

ACS Paragon Plus Environment

35