07.05.2017 Views

Cobb_et_al

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

(Proposed cover art)


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

This manuscript (PNAS MS#2017-01090R) has been accepted with minor revisions by<br />

Proceedings of the Nation<strong>al</strong> Academy of Sciences of the United States of America.<br />

How tempor<strong>al</strong> patterns in rainf<strong>al</strong>l d<strong>et</strong>ermine the<br />

geomorphology and carbon fluxes of tropic<strong>al</strong><br />

peatlands<br />

Alexander R. <strong>Cobb</strong> a,1 , Alison Hoyt b , Laure Gandois c , Jangarun Eri d , René Dommain e,f , Kamariah Abu S<strong>al</strong>im g , Fuu Ming<br />

Kai a,2 , Nur S<strong>al</strong>ihah Haji Su’ut h , and Charles F. Harvey a,b<br />

a Center for Environment<strong>al</strong> Sensing and Modeling, Singapore-MIT Alliance for Research and Technology, 138602 Singapore; b Department of Civil and Environment<strong>al</strong><br />

Engineering, Massachus<strong>et</strong>ts Institute of Technology, Cambridge, Massachus<strong>et</strong>ts, 02139 USA; c EcoLab (Laboratoire Écologie fonctionnelle <strong>et</strong> Environnement), Université de<br />

Toulouse, CNRS, INPT, UPS, Avenue de l’Agrobiopôle, F-31326 Castan<strong>et</strong>-Tolosan, France; d Forestry Department, Ministry of Industry and Primary Resources, J<strong>al</strong>an Menteri<br />

Besar, Bandar Seri Begawan BB3910, Brunei Daruss<strong>al</strong>am; e Department of Anthropology, Smithsonian Institution, Nation<strong>al</strong> Museum of Natur<strong>al</strong> History, Washington, District of<br />

Columbia, 20560 USA; f Institute of Earth and Environment<strong>al</strong> Science, University of Potsdam, 14476 Potsdam, Germany; g Biology Programme, Universiti Brunei Daruss<strong>al</strong>am,<br />

Bandar Seri Begawan BE1410, Brunei Daruss<strong>al</strong>am; h Brunei Daruss<strong>al</strong>am Heart of Borneo Centre, Ministry of Industry and Primary Resources, J<strong>al</strong>an Menteri Besar, Bandar<br />

Seri Begawan BB3910, Brunei Daruss<strong>al</strong>am; 2 Present address: Nation<strong>al</strong> M<strong>et</strong>rology Centre, Agency for Science, Technology and Research, 118221 Singapore.<br />

This manuscript was compiled on April 13, 2017<br />

Tropic<strong>al</strong> peatlands now emit hundreds of megatons of carbon dioxide<br />

per year because of human disruption of the feedbacks that link<br />

peat accumulation and groundwater hydrology. However, no quantitative<br />

theory has existed for how patterns of carbon storage and<br />

release accompanying growth and subsidence of tropic<strong>al</strong> peatlands<br />

are affected by climate and disturbance. Using comprehensive data<br />

from a pristine peatland in Brunei Daruss<strong>al</strong>am, we show how rainf<strong>al</strong>l<br />

and groundwater flow d<strong>et</strong>ermine a shape param<strong>et</strong>er (the Laplacian<br />

of the peat surface elevation) that specifies, under a given rainf<strong>al</strong>l<br />

regime, the ultimate, stable morphology, and hence carbon storage,<br />

of a tropic<strong>al</strong> peatland within a n<strong>et</strong>work of rivers or can<strong>al</strong>s. We find<br />

that peatlands reach their ultimate shape first at the edges of peat<br />

domes where they are bounded by rivers, so that the rate of carbon<br />

uptake accompanying their growth is proportion<strong>al</strong> to the area of the<br />

still-growing dome interior. We use this model to study how tropic<strong>al</strong><br />

peatland carbon storage and fluxes are controlled by changes in<br />

climate, sea level, and drainage n<strong>et</strong>works. We find that fluctuations<br />

in n<strong>et</strong> precipitation on time sc<strong>al</strong>es from hours to years can reduce<br />

long-term peat accumulation. Our mathematic<strong>al</strong> and numeric<strong>al</strong> models<br />

can be used to predict long-term effects of changes in tempor<strong>al</strong><br />

rainf<strong>al</strong>l patterns and drainage n<strong>et</strong>works on tropic<strong>al</strong> peatland geomorphology<br />

and carbon storage.<br />

tropic<strong>al</strong> peatlands | peatland geomorphology | peatland hydrology |<br />

peatland carbon storage<br />

Tropic<strong>al</strong> peatlands store gigatons of carbon in peat domes,<br />

gently mounded land forms kilom<strong>et</strong>ers across and ten or<br />

more m<strong>et</strong>ers high (1). The carbon stored as peat in these<br />

domes has been sequestered by photosynthesis of peat swamp<br />

trees (2) and preserved for thousands of years by waterlogging,<br />

which suppresses decomposition. Human disturbance<br />

of tropic<strong>al</strong> peatlands by fire and drainage for agriculture is<br />

now causing re-emission of that carbon at rates of hundreds<br />

of megatons per year (2–5): emissions from Southeast Asian<br />

peatlands <strong>al</strong>one are equiv<strong>al</strong>ent to about 2% of glob<strong>al</strong> fossil fuel<br />

emissions or 20% of glob<strong>al</strong> land use and land cover change<br />

emissions (6, 7). Because peat is mostly organic carbon, a<br />

description of the growth and subsidence of tropic<strong>al</strong> peatlands<br />

<strong>al</strong>so quantifies fluxes of carbon dioxide (1, 4, 8). Evidence<br />

from a range of studies establishes that accumulation and loss<br />

of tropic<strong>al</strong> peat are controlled by water table dynamics (4, 9).<br />

When the water table is low, aerobic decomposition occurs,<br />

releasing carbon dioxide; when the water table is high, aerobic<br />

decomposition is inhibited by lack of oxygen, production of<br />

peat exceeds its decay, and peat accumulates. In this way,<br />

the rate of peat accumulation is d<strong>et</strong>ermined by the fraction of<br />

time that peat is exposed by a low water table (Fig. 1).<br />

The water table rises and f<strong>al</strong>ls in a peatland according to the<br />

b<strong>al</strong>ance b<strong>et</strong>ween rainf<strong>al</strong>l, evapotranspiration, and groundwater<br />

flow. Water flows downslope towards the edge of each peat<br />

dome, where it is bounded by rivers. This flow occurs at a<br />

rate limited by the hydraulic transmissivity of the peat—the<br />

e ciency with which it conducts later<strong>al</strong> flow—and follows the<br />

gradient in the water table. The gradient in the water table<br />

is slightly steeper near dome boundaries where the flow of<br />

water is faster. A steeper gradient near boundaries implies<br />

a domed shape in the water table, or groundwater mound,<br />

corresponding to the domed shape of the peat surface. The<br />

doming of the peat surface is very subtle: gradients are about<br />

one m<strong>et</strong>er per kilom<strong>et</strong>er (1). Non<strong>et</strong>heless, it is the dome’s<br />

gentle curvature that accounts for the carbon storage within<br />

the drainage boundary.<br />

Once the peatland surface is su ciently domed, water is<br />

DRAFT<br />

Significance Statement<br />

A datas<strong>et</strong> from one of the last protected tropic<strong>al</strong> peat swamps<br />

in Southeast Asia reve<strong>al</strong>s how fluctuations in rainf<strong>al</strong>l on yearly<br />

and shorter timesc<strong>al</strong>es affect the growth and subsidence of<br />

tropic<strong>al</strong> peatlands over thousands of years. The pattern of<br />

rainf<strong>al</strong>l and the permeability of the peat tog<strong>et</strong>her d<strong>et</strong>ermine a<br />

particular curvature of the peat surface that defines the amount<br />

of natur<strong>al</strong>ly sequestered carbon stored in the peatland over<br />

time. This principle can be used to c<strong>al</strong>culate the long-term<br />

carbon dioxide emissions driven by changes in climate and<br />

tropic<strong>al</strong> peatland drainage. The results suggest that greater<br />

season<strong>al</strong>ity projected by climate models could lead to carbon<br />

dioxide emissions, instead of sequestration, from otherwise<br />

undisturbed peat swamps.<br />

A.R.C. and J.E. established the site and inst<strong>al</strong>led the sensors; A.R.C., L.G., J.E., R.D., K.A.S.,<br />

K.F.M., N.S.H.S., and C.F.H. collected peat cores; A.H., L.G., and J.E. compl<strong>et</strong>ed the peat surface<br />

elevation survey; A.R.C., A.H., and C.F.H. an<strong>al</strong>yzed data; A.R.C. wrote the simulation code;<br />

A.R.C. and C.F.H. designed the study and wrote the paper. All authors discussed the results and<br />

commented on the manuscript.<br />

The authors declare no conflict of interest.<br />

1 To whom correspondence should be addressed. E-mail: <strong>al</strong>ex.cobb@smart.mit.edu<br />

63<br />

64<br />

65<br />

66<br />

67<br />

68<br />

69<br />

70<br />

71<br />

72<br />

73<br />

74<br />

75<br />

76<br />

77<br />

78<br />

79<br />

80<br />

81<br />

82<br />

83<br />

84<br />

85<br />

86<br />

87<br />

88<br />

89<br />

90<br />

91<br />

92<br />

93<br />

94<br />

95<br />

96<br />

97<br />

98<br />

99<br />

100<br />

101<br />

102<br />

103<br />

104<br />

105<br />

106<br />

107<br />

108<br />

109<br />

110<br />

111<br />

112<br />

113<br />

114<br />

115<br />

116<br />

117<br />

118<br />

119<br />

120<br />

121<br />

122<br />

123<br />

124<br />

www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 1–10


125<br />

126<br />

127<br />

128<br />

129<br />

130<br />

131<br />

132<br />

133<br />

134<br />

135<br />

136<br />

137<br />

138<br />

139<br />

140<br />

141<br />

142<br />

143<br />

144<br />

145<br />

146<br />

147<br />

148<br />

149<br />

150<br />

151<br />

152<br />

153<br />

154<br />

155<br />

156<br />

157<br />

158<br />

159<br />

160<br />

161<br />

162<br />

163<br />

164<br />

165<br />

166<br />

167<br />

168<br />

169<br />

170<br />

171<br />

172<br />

173<br />

174<br />

175<br />

176<br />

177<br />

178<br />

179<br />

180<br />

181<br />

182<br />

183<br />

184<br />

185<br />

186<br />

rain and<br />

evapotranspiration<br />

groundwater<br />

flow<br />

govern<br />

and<br />

topography and<br />

transmissivity<br />

control<br />

creates<br />

water<br />

table<br />

d<strong>et</strong>ermines<br />

peat<br />

accumulation<br />

Fig. 1. Ecosystem feedback leading to peat accumulation. Peat accumulation<br />

occurs because of waterlogging of plant remains, and is therefore d<strong>et</strong>ermined by the<br />

proportion of time that peat is protected from aerobic decomposition by a high water<br />

table. Over time, peat builds up into gently mounded land forms, or domes, bounded<br />

by rivers. The slopes in a peat dome, though very sm<strong>al</strong>l, govern groundwater flow<br />

towards bounding rivers at rates limited by the transmissivity of the peat.<br />

shed so rapidly that no more organic matter can be waterlogged<br />

within the confines of the drainage n<strong>et</strong>work, and peat<br />

accumulation stops (10). This maxim<strong>al</strong>ly domed shape s<strong>et</strong>s<br />

a limit on how much carbon a peat dome can sequester and<br />

preserve under a given rainf<strong>al</strong>l regime (11). If the peat dome<br />

is flatter than its stable shape for the current climate, it will<br />

sequester carbon and grow; if it is more domed than its stable<br />

shape, it will release carbon and subside as peat decomposes.<br />

(In the tropic<strong>al</strong> peat literature, “subsidence” is used for a<br />

decline in the peat surface elevation, regardless of mechanism<br />

(5).) The volume of this stable shape times the average carbon<br />

density of the peat defines a capacity for storage of carbon as<br />

peat within the drainage boundary.<br />

If we can predict the stable shapes of peat domes and how<br />

they evolve over time in a given climate, we can d<strong>et</strong>ermine<br />

how peatland carbon storage capacity and carbon fluxes are<br />

a ected by changes in rainf<strong>al</strong>l regime, drainage n<strong>et</strong>work, and<br />

sea level. However, when predicting the stable shapes of peat<br />

domes and their evolution towards these shapes, there are two<br />

complicating factors: (1) the boundaries imposed by drainage<br />

n<strong>et</strong>works have complex shapes, and (2) rainf<strong>al</strong>l is intermittent<br />

and variable. The water table rises during rainstorms, and<br />

f<strong>al</strong>ls during dry periods, even when the peat surface is stable.<br />

These fluctuations in the water table seem to be important<br />

because it is widely believed that season<strong>al</strong>ity of rainf<strong>al</strong>l a ects<br />

tropic<strong>al</strong> peat accumulation (12, 13). But how should we take<br />

these fluctuations into account to predict the slow development<br />

and stable shapes of peat domes? Understanding the glob<strong>al</strong><br />

impact of changes in rainf<strong>al</strong>l amount and variability, drainage<br />

n<strong>et</strong>works, and sea level on tropic<strong>al</strong> peatland carbon storage and<br />

fluxes requires a theory that can accommodate the complicated<br />

drainage n<strong>et</strong>works and intermittent rainf<strong>al</strong>l of the re<strong>al</strong> world.<br />

Ingram (10) made the first prediction of the limiting shape<br />

of a temperate peat dome imposed by the b<strong>al</strong>ance b<strong>et</strong>ween<br />

rainf<strong>al</strong>l and groundwater flow. Assuming constant rainf<strong>al</strong>l,<br />

he computed the steady-state shape of a peat dome with<br />

uniform permeability b<strong>et</strong>ween par<strong>al</strong>lel rivers. Clymo (14) later<br />

developed a simple dynamic model for accumulation of peat at<br />

a single point in the landscape. Clymo’s model assumed that<br />

the thickness of peat above the water table would not change,<br />

and focused on anaerobic decomposition in deeper waterlogged<br />

peat. Hilbert <strong>et</strong> <strong>al</strong>. (15) later built on Clymo’s model to<br />

<strong>al</strong>low a varying thickness of peat above the water table via a<br />

simple water b<strong>al</strong>ance whereby drainage increases linearly with<br />

peat surface elevation. Hilbert’s model inspired a series of<br />

a<br />

5° N<br />

0°<br />

5° S<br />

500 km<br />

4° N<br />

100° E 110° E 120° E<br />

1 km<br />

Sumatra<br />

DRAFT<br />

c<br />

d<br />

Borneo<br />

b<br />

b<br />

5° N<br />

N<br />

d<br />

500 m<br />

e<br />

30 m<br />

30 km<br />

c<br />

Brunei<br />

c<br />

114° E 115° E<br />

e<br />

Sarawak<br />

Fig. 2. Site of field data collection in Brunei Daruss<strong>al</strong>am. a, Distribution of<br />

peatlands in Borneo, Sumatra and Peninsular M<strong>al</strong>aysia. b, Field site in Brunei<br />

Daruss<strong>al</strong>am, on Borneo island. c, Contour map of study area from airborne LiDAR<br />

data, showing radiocarbon-dated peat cores (points) at primary site (Mendaram,<br />

south) and degraded site (Damit, north), and the boundaries of the flow tube used<br />

for hydrologic simulations (blue). d, Piezom<strong>et</strong>ers (triangles) at the Mendaram site. e,<br />

Survey points in microtopography transect (see Fig. 3b).<br />

increasingly sophisticated models for veg<strong>et</strong>ation dynamics and<br />

peat accumulation at a point. The most recent of these point<br />

models computes water table depth from monthly rainf<strong>al</strong>l<br />

using a site-specific model (16). Meanwhile, numeric<strong>al</strong> models<br />

have been used to simulate peat accumulation under constant<br />

rainf<strong>al</strong>l (17, 18). Although these subsequent works simulate the<br />

dynamics of peat production and decomposition in increasing<br />

d<strong>et</strong>ail, a strength of Ingram’s model was that it provided<br />

quantitative intuition for how peat dome morphology depends<br />

on peat hydrologic properties and average rainf<strong>al</strong>l. Could a<br />

principle like Ingram’s exist that describes peatland dynamics<br />

as well as statics, and remains applicable with re<strong>al</strong>istic drainage<br />

n<strong>et</strong>works and rainf<strong>al</strong>l regimes?<br />

We established a field site in one of the last pristine peat<br />

swamp forests in Southeast Asia, then used measurements<br />

from this site to develop a new mathematic<strong>al</strong> model for the<br />

geomorphic evolution of tropic<strong>al</strong> peatlands that is simpler, y<strong>et</strong><br />

more gener<strong>al</strong> than Ingram’s model for high-latitude peatlands.<br />

Our model makes it possible to predict e ects of changes in<br />

rainf<strong>al</strong>l regime and drainage n<strong>et</strong>works on carbon storage and<br />

fluxes in tropic<strong>al</strong> peatlands. The model predicted, perhaps<br />

surprisingly, that surface peat would be older near dome margins.<br />

We tested these predictions by radiocarbon-dating core<br />

samples and comparing the age of each sample to the simulated<br />

age at its location and depth. Fin<strong>al</strong>ly, we explored the future<br />

of tropic<strong>al</strong> peatlands under climate projections by simulating<br />

the geomorphic evolution of an ide<strong>al</strong>ized peat dome under<br />

projected changes in rainf<strong>al</strong>l patterns and drainage.<br />

187<br />

188<br />

189<br />

190<br />

191<br />

192<br />

193<br />

194<br />

195<br />

196<br />

197<br />

198<br />

199<br />

200<br />

201<br />

202<br />

203<br />

204<br />

205<br />

206<br />

207<br />

208<br />

209<br />

210<br />

211<br />

212<br />

213<br />

214<br />

215<br />

216<br />

217<br />

218<br />

219<br />

220<br />

221<br />

222<br />

223<br />

224<br />

225<br />

226<br />

227<br />

228<br />

229<br />

230<br />

231<br />

232<br />

233<br />

234<br />

235<br />

236<br />

237<br />

238<br />

239<br />

240<br />

241<br />

242<br />

243<br />

244<br />

245<br />

246<br />

247<br />

248<br />

2 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

<strong>Cobb</strong> <strong>et</strong> <strong>al</strong>.


249<br />

250<br />

251<br />

252<br />

253<br />

254<br />

255<br />

256<br />

257<br />

258<br />

259<br />

260<br />

261<br />

262<br />

263<br />

264<br />

265<br />

266<br />

267<br />

268<br />

269<br />

270<br />

271<br />

272<br />

273<br />

274<br />

275<br />

276<br />

277<br />

278<br />

279<br />

280<br />

281<br />

282<br />

283<br />

284<br />

285<br />

286<br />

287<br />

288<br />

289<br />

290<br />

291<br />

292<br />

293<br />

294<br />

295<br />

296<br />

297<br />

298<br />

299<br />

300<br />

301<br />

302<br />

303<br />

304<br />

305<br />

306<br />

307<br />

308<br />

309<br />

310<br />

a<br />

Elevation, m a.s.l.<br />

8 decay<br />

6<br />

4<br />

2<br />

litter<br />

peat<br />

b 5.50<br />

Elevation, m a.s.l.<br />

5.25<br />

5.00<br />

pool<br />

0 50 100 150<br />

Position <strong>al</strong>ong transect, m<br />

DRAFT<br />

Height above<br />

land surface, cm<br />

20 c<br />

0<br />

-20<br />

2012-09 2012-10<br />

Fig. 3. Microtopography and water table dynamics in a tropic<strong>al</strong> peatland. a, Cartoon of tropic<strong>al</strong> peat cross-section showing variables: ˜p, the peat surface; p, the “land<br />

surface,” a smooth surface fit through loc<strong>al</strong> minima in ˜p; H, water table elevation; ’, water table height relative to the land surface, ’ = H ≠ p. The peat surface ˜p is irregular<br />

on a spati<strong>al</strong> sc<strong>al</strong>e of m<strong>et</strong>ers, with higher areas (hummocks) separating loc<strong>al</strong> depressions (hollows) that are not connected into channels. b, Tot<strong>al</strong> station survey of peat elevation<br />

˜p (black circles) <strong>al</strong>ong a transect, and the land surface p (dashed black line). The minimum, median, and maximum water table elevation H from each of 12 piezom<strong>et</strong>ers <strong>al</strong>ong<br />

the transect are <strong>al</strong>so shown (dashed blue lines). The absolute elevation of the survey points comes from matching loc<strong>al</strong> minima among survey points within 20 m x 20 m squares<br />

(white diamonds) with loc<strong>al</strong> minima in LiDAR last r<strong>et</strong>urn data within the same squares (red diamonds). The land surface p is represented by the dashed horizont<strong>al</strong> black line. c,<br />

Water table dynamics <strong>al</strong>ong survey transect (d) in late 2012, relative to the land surface p. What appears to be a single blue line is superimposed data from the 12 piezom<strong>et</strong>ers<br />

shown in (d). Also shown are the average minimum, median, and maximum water table elevation above the land surface during the same time period for <strong>al</strong>l 12 piezom<strong>et</strong>ers.<br />

M<strong>et</strong>hods Summary<br />

Field measurements. We established a field site in pristine<br />

peat forest in Brunei Daruss<strong>al</strong>am (Borneo) to study a peat<br />

dome where current processes a ecting peat accumulation<br />

are essenti<strong>al</strong>ly similar to those during its long-term development<br />

(Fig. 2). At the site, we inst<strong>al</strong>led 5 piezom<strong>et</strong>ers <strong>al</strong>ong<br />

a 2.5 km trail, 12 piezom<strong>et</strong>ers <strong>al</strong>ong a 180 m transect, and 3<br />

throughf<strong>al</strong>l gauges. We compl<strong>et</strong>ed a tot<strong>al</strong> station survey of<br />

peat surface elevation <strong>al</strong>ong the transect to characterize peat<br />

surface microtopography. To characterize large-sc<strong>al</strong>e peatland<br />

morphology, we <strong>al</strong>so obtained LiDAR data for the entire study<br />

area. To study peat dome development, we collected 9 peat<br />

cores from which we obtained 37 radiocarbon dates. To test<br />

wh<strong>et</strong>her our undisturbed site behaved similarly to sites studied<br />

by other groups, we inst<strong>al</strong>led 4 soil respiration chambers and<br />

a piezom<strong>et</strong>er at a nearby logged but undrained site.<br />

Morphology vs. microtopography. Superimposed on the gross<br />

morphology of a peat dome is a fine microtopography of m<strong>et</strong>ersc<strong>al</strong>e<br />

depressions, or hollows, separated by higher areas, or<br />

hummocks (19, 20). The hummocks consist of partly decomposed<br />

logs, branches, and leaves lodged among living<br />

buttresses, stilt roots, pneumatophores, and giant rhizomes.<br />

Whereas the microtopography in high-latitude peat bogs may<br />

have regular and oriented patterns (21), surveys by Lampela<br />

<strong>et</strong> <strong>al</strong>. (20) in a tropic<strong>al</strong> peat swamp in Centr<strong>al</strong> K<strong>al</strong>imantan<br />

showed no orientation or regularity. Similarly, our microtopography<br />

survey and other observations reve<strong>al</strong>ed no regular<br />

patterns or channels in peat dome microtopography.<br />

In describing the evolution of peat dome morphology, we<br />

would like to capture the e ects of the hummock-and-hollow<br />

microtopography without explicitly simulating its d<strong>et</strong>ails. Measurements<br />

from the 12 piezom<strong>et</strong>ers <strong>al</strong>ong our microtopography<br />

transect showed that the water table is relatively smooth, even<br />

though the peat surface is highly irregular on a spati<strong>al</strong> sc<strong>al</strong>e<br />

of centim<strong>et</strong>ers to m<strong>et</strong>ers (Fig. 3). We therefore represent the<br />

peat surface by a reference surface p, smooth like the water<br />

table, that underlies the actu<strong>al</strong> peat surface ˜p. We refer to<br />

this reference surface p as the “land surface.” The peat surface<br />

˜p is a “texture” that sits on the smooth land surface p. The<br />

bottoms of hollows provide the most readily identifiable loc<strong>al</strong><br />

reference elevation (20), so we define the land surface p as a<br />

smooth surface fit through the bottoms of hollows (loc<strong>al</strong> minima<br />

in the peat surface ˜p). On the basis of this definition, we<br />

d<strong>et</strong>ermined the current land surface at our site by smoothing a<br />

raster map obtained from loc<strong>al</strong> minima in LiDAR last-r<strong>et</strong>urn<br />

points. We <strong>al</strong>so used the transect survey and piezom<strong>et</strong>er data<br />

to find the land surface p <strong>al</strong>ong the microtopography survey<br />

transect (SI Text).<br />

Groundwater flow. We model the dynamics of the water table<br />

H subject to n<strong>et</strong> precipitation q n (rainf<strong>al</strong>l intensity R minus<br />

evapotranspiration ET) using Boussinesq’s equation for<br />

essenti<strong>al</strong>ly horizont<strong>al</strong> groundwater flow<br />

ˆH<br />

S y = qn + Ò · (T ÒH) [1]<br />

ˆt<br />

where the specific yield S y is the amount of water required<br />

for a di erenti<strong>al</strong> increment in water table elevation, and transmissivity<br />

T is the volum<strong>et</strong>ric flow per perim<strong>et</strong>er driven by<br />

a particular head gradient ÒH. Boussinesq’s equation is a<br />

standard groundwater modeling equation for flow domains like<br />

peatlands that are much wider than they are thick.<br />

At high water tables, hollows become flooded from saturation<br />

of the peat below, forming sm<strong>al</strong>l pools. These pools<br />

are not connected into channels (20), and therefore do not<br />

<strong>al</strong>low open-channel flow on a large sc<strong>al</strong>e in the peatland. Instead,<br />

flow through the peatland is limited by flow through<br />

the porous matrix of the hummocks b<strong>et</strong>ween these isolated<br />

pools. We apply Boussinesq’s equation at sc<strong>al</strong>es much larger<br />

than hummocks and hollows (tens of m<strong>et</strong>ers) and refer to<br />

the flow of water through the peatland as “groundwater flow”<br />

even though some flow occurs above the loc<strong>al</strong> peat surface, in<br />

hollows, during w<strong>et</strong> periods. Boussinesq’s equation requires<br />

only that later<strong>al</strong> flow is proportion<strong>al</strong> to the head gradient,<br />

which is the case if the over<strong>al</strong>l flow is limited by laminar flow<br />

through hummocks. We never observed ephemer<strong>al</strong> channels<br />

connecting hollows within the peatland in our six years at the<br />

site. In addition, if flow were non-laminar, we would expect<br />

di erent loc<strong>al</strong> flow behavior at the same water table height in<br />

areas with di erent water table gradients, but instead water<br />

table behavior is uniform (Results and Discussion).<br />

Loc<strong>al</strong> carbon b<strong>al</strong>ance. A broad range of studies demonstrates<br />

that the thickness of peat exposed above the water table d<strong>et</strong>ermines<br />

the rate of peat accumulation or loss (4, 22). Like<br />

others (4, 22), we modeled the dynamics of peat accumulation<br />

311<br />

312<br />

313<br />

314<br />

315<br />

316<br />

317<br />

318<br />

319<br />

320<br />

321<br />

322<br />

323<br />

324<br />

325<br />

326<br />

327<br />

328<br />

329<br />

330<br />

331<br />

332<br />

333<br />

334<br />

335<br />

336<br />

337<br />

338<br />

339<br />

340<br />

341<br />

342<br />

343<br />

344<br />

345<br />

346<br />

347<br />

348<br />

349<br />

350<br />

351<br />

352<br />

353<br />

354<br />

355<br />

356<br />

357<br />

358<br />

359<br />

360<br />

361<br />

362<br />

363<br />

364<br />

365<br />

366<br />

367<br />

368<br />

369<br />

370<br />

371<br />

372<br />

<strong>Cobb</strong><br />

<strong>et</strong> <strong>al</strong>.<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 3


373<br />

374<br />

375<br />

376<br />

377<br />

378<br />

379<br />

380<br />

381<br />

382<br />

383<br />

384<br />

385<br />

386<br />

387<br />

388<br />

389<br />

390<br />

391<br />

392<br />

393<br />

394<br />

395<br />

396<br />

397<br />

398<br />

399<br />

400<br />

401<br />

402<br />

403<br />

404<br />

405<br />

406<br />

407<br />

408<br />

409<br />

410<br />

411<br />

412<br />

413<br />

414<br />

415<br />

416<br />

417<br />

418<br />

419<br />

420<br />

421<br />

422<br />

423<br />

424<br />

425<br />

426<br />

427<br />

428<br />

429<br />

430<br />

431<br />

432<br />

433<br />

434<br />

Water table height , cm<br />

a 20<br />

0<br />

−20<br />

−40<br />

−60<br />

−80<br />

Land surface p<br />

(bottoms of hollows)<br />

−6 −4 −2 0<br />

Peat accumulation (+)<br />

or loss (−) , cm / y<br />

b<br />

0 2 4<br />

Soil surface CO 2 flux,<br />

µmol m -2 s -1<br />

Fig. 4. Peat accumulation and CO 2 flux vs. water table height in tropic<strong>al</strong> peatlands.<br />

Peat accumulation (a) represents the b<strong>al</strong>ance b<strong>et</strong>ween peat production and<br />

decomposition. Aerobic decomposition is one of the two main sources of peat surface<br />

CO 2 flux (b); the other is root respiration. a, Peat accumulation or loss vs. water<br />

table height from model c<strong>al</strong>ibration (solid line) and from literature subsidence data<br />

(circles: (4); squares: (22)). The straight line was not fit to these data, but rather, arose<br />

natur<strong>al</strong>ly from c<strong>al</strong>ibration to match the modern surface of the Mendaram peat dome<br />

(Figure 7). b, Soil surface CO 2 flux vs. water table height at our site in Brunei Daruss<strong>al</strong>am<br />

(white circles) was very similar to fluxes in other tropic<strong>al</strong> peatlands (squares<br />

(23), diamonds (19), triangles (24); pentagons (9); hexagons (25)).<br />

or loss ˆp/ˆt as the di erence b<strong>et</strong>ween the rate of peat production<br />

f p when the water table is at the land surface, and<br />

the rate of peat loss by decomposition (p ≠ H)–, which is the<br />

thickness p ≠ H of peat exposed above the water table times<br />

a decomposition rate constant –<br />

ˆp<br />

= fp ≠ (p ≠ H)– [2]<br />

ˆt<br />

(Fig. 4). The peat surface is stable, neither growing nor<br />

subsiding Ȉp/ˆtÍ =0wherever the water table fluctuates in<br />

such a way that peat production is b<strong>al</strong>anced by decomposition<br />

over time<br />

f p = Èp ≠ HÍ –, [3]<br />

where angle brack<strong>et</strong>s È·Í indicate a time average.<br />

Sever<strong>al</strong> other studies have shown a leveling-o of soil CO 2<br />

e ux at very low water tables (25, 26), and it is <strong>al</strong>so likely<br />

that very high water tables ultimately limit n<strong>et</strong> carbon uptake<br />

by trees (primary production) (16). However, including these<br />

e ects did not a ect simulations because these extreme water<br />

table heights and depths were neither observed at our site<br />

nor predicted by simulations of our site. We <strong>al</strong>so did not<br />

include anaerobic decomposition below the water table because<br />

an<strong>al</strong>ysis of peat cores from tropic<strong>al</strong> sites in Asia (2), including<br />

our site (27), do not show d<strong>et</strong>ectable loss of waterlogged peat<br />

from anaerobic decomposition.<br />

Numeric<strong>al</strong> simulations. We built a numeric<strong>al</strong> model of waterlogging<br />

and peat accumulation based on Eqn. 1 and Eqn. 2<br />

to simulate peat dome geomorphogenesis and carbon fluxes.<br />

These two equations are coupled by the water table elevation<br />

H and the peat surface elevation p, both of which vary in<br />

time and space. The equations require four param<strong>et</strong>ers: (1)<br />

a specific yield function S y, (2) a transmissivity function T ,<br />

(3) a rate of peat production f p, and (4) a decomposition rate<br />

constant –. The model employs a finite volume scheme with<br />

speci<strong>al</strong> features designed to handle the severe nonlinearity of<br />

the transmissivity function T (SI Text).<br />

We d<strong>et</strong>ermined the specific yield and transmissivity functions<br />

S y,T from the response of the water table to heavy<br />

rain and dry spells (Results and Discussion). We then fit<br />

the param<strong>et</strong>ers for peat accumulation f p, – by simulating the<br />

2700-year evolution of a peat dome at our field site in Brunei,<br />

and matching the simulated modern peat surface to the peat<br />

surface measured by LiDAR. We tested our model against radiocarbon<br />

dates from peat cores extracted from the peatland,<br />

then used the model to answer gener<strong>al</strong> questions about carbon<br />

fluxes from tropic<strong>al</strong> peatlands after perturbation by climate<br />

change and drainage.<br />

Limitations of modeling approach. Our go<strong>al</strong> was to build the<br />

simplest model that can make reasonable quantitative predictions<br />

of tropic<strong>al</strong> peat dome dynamics. In most Southeast Asian<br />

peatland complexes, every area b<strong>et</strong>ween rivers is occupied by a<br />

peat dome, so it is not apparent how any peat dome could now<br />

expand to fill a larger area. However, domes tend to be larger<br />

in older peatlands, suggesting a long-term process of dome<br />

co<strong>al</strong>escence. We did not attempt to model these long-term<br />

changes in river n<strong>et</strong>works. We <strong>al</strong>so did not consider changes in<br />

hydraulic conductivity near the surface caused by compaction<br />

or changes in microtopography under agriculture.<br />

Results and Discussion<br />

Carbon storage capacity of tropic<strong>al</strong> peatlands.<br />

Loc<strong>al</strong> water b<strong>al</strong>ance is dominated by flows near the surface. Eighteen<br />

months of data on water table height in five piezom<strong>et</strong>ers<br />

<strong>al</strong>ong a 2.5 km transect (Fig. 5) show two distinctive features<br />

of water table behavior in tropic<strong>al</strong> peatlands. First, when the<br />

water table is high, it f<strong>al</strong>ls very rapidly; and second, the water<br />

table height relative to the land surface remains approximately<br />

uniform in <strong>al</strong>l piezom<strong>et</strong>ers as the water table rises and f<strong>al</strong>ls,<br />

as observed elsewhere by Hooijer (28). In what follows, we<br />

use “water table height” ’ = H ≠ p to refer to the water<br />

table height relative to the land surface, as distinct from the<br />

water table elevation H above mean sea level. Because the<br />

water table height ’ is approximately uniform, the water table<br />

behavior can be summarized by a pair of curves describing<br />

the uniform rise of the water table during heavy rain, and<br />

the uniform decline of the water table during dry interv<strong>al</strong>s<br />

b<strong>et</strong>ween rains (Fig. 5e,f). During heavy rain, the e ects of<br />

evapotranspiration and outward flow are negligible, and the<br />

rainf<strong>al</strong>l intensity vs. rate of increase in water table height<br />

gives the specific yield. B<strong>et</strong>ween rain events, the water table<br />

DRAFT<br />

declines because of evapotranspiration and the divergence of<br />

groundwater flow.<br />

Transmissivity T is a function of water table height ’ and<br />

controls the divergence of groundwater flow Ò · (T ÒH). We<br />

d<strong>et</strong>ermined the e ect of water table height on transmissivity<br />

using our water table data. The water table declines during<br />

dry interv<strong>al</strong>s because of a combination of evapotranspiration<br />

and the divergence of groundwater flow; however, the two are<br />

easily distinguished at low water tables because evapotranspiration<br />

ceases at night (Fig. 5d). Therefore, we can obtain<br />

the divergence of groundwater flow from the declining water<br />

table during dry interv<strong>al</strong>s after accounting for evapotranspiration<br />

((28, 29); further d<strong>et</strong>ails in SI Text). We find that<br />

transmissivity increases exponenti<strong>al</strong>ly at high water tables,<br />

when water rises into hollows and flows through hummocks,<br />

but decreases dramatic<strong>al</strong>ly at low water tables when water<br />

flows through fine pores in the peat matrix (Fig. 5c). Very<br />

high permeability near the peat surface is consistent with our<br />

observations of more void space higher in the peat profile, and<br />

435<br />

436<br />

437<br />

438<br />

439<br />

440<br />

441<br />

442<br />

443<br />

444<br />

445<br />

446<br />

447<br />

448<br />

449<br />

450<br />

451<br />

452<br />

453<br />

454<br />

455<br />

456<br />

457<br />

458<br />

459<br />

460<br />

461<br />

462<br />

463<br />

464<br />

465<br />

466<br />

467<br />

468<br />

469<br />

470<br />

471<br />

472<br />

473<br />

474<br />

475<br />

476<br />

477<br />

478<br />

479<br />

480<br />

481<br />

482<br />

483<br />

484<br />

485<br />

486<br />

487<br />

488<br />

489<br />

490<br />

491<br />

492<br />

493<br />

494<br />

495<br />

496<br />

4 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

<strong>Cobb</strong> <strong>et</strong> <strong>al</strong>.


497<br />

498<br />

499<br />

500<br />

501<br />

502<br />

503<br />

504<br />

505<br />

506<br />

507<br />

508<br />

509<br />

510<br />

511<br />

512<br />

513<br />

514<br />

515<br />

516<br />

517<br />

518<br />

519<br />

520<br />

521<br />

522<br />

523<br />

524<br />

525<br />

526<br />

527<br />

528<br />

529<br />

530<br />

531<br />

532<br />

533<br />

534<br />

535<br />

536<br />

537<br />

538<br />

539<br />

540<br />

541<br />

542<br />

543<br />

544<br />

545<br />

546<br />

547<br />

548<br />

549<br />

550<br />

551<br />

552<br />

553<br />

554<br />

555<br />

556<br />

557<br />

558<br />

Water table height , cm<br />

Water table height , cm<br />

20<br />

0<br />

−20<br />

20<br />

0<br />

−20<br />

a<br />

(Maximum loc<strong>al</strong> peat surface height in transect)<br />

2012-02 2012-04 2012-06 2012-08 2012-10 2012-12<br />

d<br />

Heavy rain<br />

Time<br />

No rain<br />

e<br />

Heavy rain<br />

0 10<br />

Rainf<strong>al</strong>l depth P, cm<br />

f Bog plain<br />

piezom<strong>et</strong>er<br />

Rainf<strong>al</strong>l intensity R, cm / h<br />

DRAFT<br />

10<br />

5<br />

0<br />

Water table height , cm<br />

20<br />

−20<br />

b<br />

0 0.25 0.5 10 2 10 4<br />

Specific yield S y ,<br />

cm / cm<br />

Transmissivity T,<br />

m 2 d -1<br />

No rain<br />

0 10 20 30<br />

Time since rain stopped, d<br />

Fig. 5. Site hydrology and c<strong>al</strong>ibration. a, Superimposed water table height (jagged blue lines) from five piezom<strong>et</strong>ers spanning 2.5 km and rainf<strong>al</strong>l intensity (vertic<strong>al</strong> lines)<br />

from three automated rain gauges over a 10-month interv<strong>al</strong>. The piezom<strong>et</strong>er farthest from the river (red) lies in a region with a different surface Laplacian (the “bog plain”),<br />

corresponding to an area of current peat accumulation (Fig. 6). Also shown are the minimum, median, and maximum loc<strong>al</strong> peat surface elevation (dashed horizont<strong>al</strong> lines) from<br />

a 180 m microtopography survey transect (Fig. 3). b, c, Hillslope-sc<strong>al</strong>e specific yield and transmissivity curves for field site, d<strong>et</strong>ermined from recharge and recession curves (e,<br />

f). d, Short interv<strong>al</strong> of water table data from a single piezom<strong>et</strong>er selected from (a). Ins<strong>et</strong> shows declining water tables during day (unshaded) and steady water tables at night<br />

(shaded) driven by diurn<strong>al</strong> cycles of evapotranspiration. e, f, Master recharge curve (e) and recession curve (f) assembled from interv<strong>al</strong>s of heavy rain and no rain, respectively,<br />

by <strong>al</strong>ignment of sequences with overlapping water table depth. During heavy rain, n<strong>et</strong> precipitation intensity q n = R ≠ ET is dominated by rainf<strong>al</strong>l intensity R (e); with no rain,<br />

n<strong>et</strong> precipitation consists only of evapotranspiration ET (f). Dashed black lines in (e) and (f) show water table response computed from specific yield and transmissivity (b,c),<br />

blue translucent lines are assembled from field data in (a). As in (a), the red curve is from the piezom<strong>et</strong>er in the flatter bog plain region (Fig. 6).<br />

<strong>al</strong>so with recent data from other tropic<strong>al</strong> peatlands (30). The<br />

water table curves (Fig. 5e,f) indicate that the near-surface<br />

permeability is so great that the tot<strong>al</strong> thickness of deeper peat<br />

is unimportant for groundwater flow. Therefore, transmissivity<br />

is approximately independent of peat depth, and depends only<br />

on the water table height ’, which is uniform in space (though<br />

highly variable in time).<br />

Morphology of peat surface explains uniform water table behavior.<br />

According to Boussinesq’s equation, uniform transmissivity<br />

is not, by itself, enough to explain the uniform fluctuation of<br />

the water table. Even in hydrologic systems where hydraulic<br />

properties are uniform, the water table can behave di erently<br />

at di erent locations because of topography. For example, in<br />

most hydrologic systems a rainstorm drives a di erent water<br />

table response at a topographic divide than it does near where<br />

groundwater discharges to a river.<br />

To understand the uniform water table behavior in peatlands,<br />

we refer back to Boussinesq’s equation (Eqn. 1). If both<br />

the specific yield S y and the transmissivity T depend only on<br />

the loc<strong>al</strong> water table height relative to the surface and not on<br />

position within the peatland, uniform water table movement<br />

occurs if the divergence of the peat surface gradient, or the<br />

peat surface Laplacian Ò 2 p, is uniform (Fig. S4c–e). (The<br />

“Laplacian of the peat surface” Ò 2 p, or just “Laplacian,” is<br />

the sc<strong>al</strong>ar result of applying the Laplacian operator Ò 2 to the<br />

land surface elevation p.) To see why a uniform land surface<br />

Laplacian explains uniform water table behavior, we re-write<br />

Boussinesq’s equation (Eqn. 1) in terms of the water table<br />

height relative to the land surface (’ = H ≠ p), instead of the<br />

water table elevation H:<br />

S y<br />

ˆ(p + ’)<br />

ˆt<br />

= q n + Ò · [T Ò(p + ’)] . [4]<br />

0<br />

We observe that water table height is uniform (Ò’ = 0). If<br />

transmissivity T is <strong>al</strong>so spati<strong>al</strong>ly uniform, the groundwater<br />

divergence term simplifies to the transmissivity times the<br />

peat surface Laplacian (Ò · [T Ò(p + ’)] = T Ò 2 p). The time<br />

derivative ˆp/ˆt of the land surface elevation is negligible<br />

because peat accumulation or loss is much slower than rise<br />

or f<strong>al</strong>l of the water table, so the term p can be dropped from<br />

the time derivative. We observe that the fluctuations in water<br />

table height ˆ’/ˆt are uniform, as is n<strong>et</strong> precipitation q n,so<br />

the groundwater divergence term T Ò 2 p must <strong>al</strong>so be spati<strong>al</strong>ly<br />

uniform. Thus, Boussinesq’s equation simplifies to an ordinary<br />

di erenti<strong>al</strong> equation (ODE) describing the uniform fluctuation<br />

of the water table relative to the peat surface<br />

S y<br />

d’<br />

dt = qn + T Ò2 p [5]<br />

where the peat surface Laplacian Ò 2 p is uniform.<br />

The peat surface Laplacian describes the curvature of the<br />

peat surface: it is equ<strong>al</strong> to the sum of the second derivatives of<br />

the surface elevation in two perpendicular horizont<strong>al</strong> directions<br />

(Ò 2 p = ˆ2p/ˆx 2 + ˆ2p/ˆy 2 ). Thus, an<strong>al</strong>ysis of water table<br />

dynamics predicts uniform curvature of the peat surface where<br />

water table fluctuations are uniform. This uniformity of surface<br />

elevation curvature can be tested against elevation maps.<br />

Maps of the peat surface Laplacian are highly sensitive to<br />

microtopographic noise in the surface elevation map because<br />

the Laplacian uses the second derivative of the surface elevation.<br />

However, by the Divergence Theorem, the average<br />

Laplacian within any closed contour is equ<strong>al</strong> to the integr<strong>al</strong> of<br />

the norm<strong>al</strong> gradient <strong>al</strong>ong the contour divided by the enclosed<br />

area. Therefore, we can examine the uniformity of the surface<br />

Laplacian by studying the slope of a regression b<strong>et</strong>ween the<br />

c<br />

559<br />

560<br />

561<br />

562<br />

563<br />

564<br />

565<br />

566<br />

567<br />

568<br />

569<br />

570<br />

571<br />

572<br />

573<br />

574<br />

575<br />

576<br />

577<br />

578<br />

579<br />

580<br />

581<br />

582<br />

583<br />

584<br />

585<br />

586<br />

587<br />

588<br />

589<br />

590<br />

591<br />

592<br />

593<br />

594<br />

595<br />

596<br />

597<br />

598<br />

599<br />

600<br />

601<br />

602<br />

603<br />

604<br />

605<br />

606<br />

607<br />

608<br />

609<br />

610<br />

611<br />

612<br />

613<br />

614<br />

615<br />

616<br />

617<br />

618<br />

619<br />

620<br />

<strong>Cobb</strong><br />

<strong>et</strong> <strong>al</strong>.<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 5


621<br />

622<br />

623<br />

624<br />

625<br />

626<br />

627<br />

628<br />

629<br />

630<br />

631<br />

632<br />

633<br />

634<br />

635<br />

636<br />

637<br />

638<br />

639<br />

640<br />

641<br />

642<br />

643<br />

644<br />

645<br />

646<br />

647<br />

648<br />

649<br />

650<br />

651<br />

652<br />

653<br />

654<br />

655<br />

656<br />

657<br />

658<br />

659<br />

660<br />

661<br />

662<br />

663<br />

664<br />

665<br />

666<br />

667<br />

668<br />

669<br />

670<br />

671<br />

672<br />

673<br />

674<br />

675<br />

676<br />

677<br />

678<br />

679<br />

680<br />

681<br />

682<br />

b<br />

Land surface<br />

elevation p, m<br />

c<br />

Integrated<br />

norm<strong>al</strong> gradient, m<br />

a<br />

6<br />

4<br />

2<br />

4<br />

Bog plain<br />

piezom<strong>et</strong>er<br />

Bog plain<br />

Bog plain<br />

Bog plain piezom<strong>et</strong>er<br />

Bog plain<br />

2 0<br />

Distance from river, km<br />

Enclosed area, km 2<br />

Stable<br />

Stable<br />

(uniform surface<br />

Laplacian)<br />

Stable<br />

River flooding<br />

influence<br />

River<br />

flooding<br />

influence<br />

River flooding<br />

influence<br />

Fig. 6. Estimation of peat surface Laplacian. a, Regions of different morphology<br />

and water table behavior within the flow tube used for field site simulations, and<br />

locations of piezom<strong>et</strong>ers (triangles). Furthest from the river, the land surface is<br />

relatively flat (“bog plain”), next there is a region in which the Laplacian of the land<br />

surface elevation is uniform (“stable”) and fin<strong>al</strong>ly a narrow region near the river where<br />

hydrologic processes and peat accumulation are affected by the rise and f<strong>al</strong>l of the<br />

bounding river (“river flooding influence”). b, Profile of LiDAR land surface elevation<br />

from (a), showing piezom<strong>et</strong>er locations (vertic<strong>al</strong> dashed lines). c, Norm<strong>al</strong> gradient<br />

driving efflux, integrated <strong>al</strong>ong contours, vs. enclosed area. The slope in the “stable”<br />

region gives the average land surface Laplacian of the land surface there and was<br />

used for c<strong>al</strong>ibration of hydrologic param<strong>et</strong>ers.<br />

integrated norm<strong>al</strong> gradient and the enclosed area (Fig. 6).<br />

Indeed, we find a linear relationship b<strong>et</strong>ween the integrated<br />

norm<strong>al</strong> gradient <strong>al</strong>ong each contour and the area enclosed by<br />

the contour in our LiDAR-derived peat surface elevation map,<br />

indicating a uniform surface Laplacian in the region of uniform<br />

water table behavior (Fig. 6). In contrast, outside the region<br />

of uniform Laplacian, the water table behaves di erently (‘bog<br />

plain piezom<strong>et</strong>er’ in Figs. 5,6).<br />

Uniform surface Laplacian d<strong>et</strong>ermines stable tropic<strong>al</strong> peatland morphology.<br />

The uniform peat surface Laplacian provides a remarkably<br />

simple way to compute a stable morphology for a tropic<strong>al</strong><br />

peat dome. By “stable morphology,” we mean a morphology<br />

in which the peat surface and water table continue to fluctuate<br />

with the vagaries of climate, but there is no long-term average<br />

change in the peat surface or water table elevation (they are<br />

stationary; Ȉp/ˆtÍ =0, ȈH/ˆtÍ =0). Uniform water table<br />

height is the simplest behavior that could make an entire<br />

peatland stable, because if the water table height is spati<strong>al</strong>ly<br />

uniform, the loc<strong>al</strong> rate of peat accumulation is <strong>al</strong>so uniform.<br />

In a stable peatland, there is no long-term change in the water<br />

table height, so any water added by n<strong>et</strong> precipitation must<br />

eventu<strong>al</strong>ly be removed by groundwater flow<br />

e f<br />

d’<br />

0= S y = Èq nÍ + ÈT ÍÒ 2 p Œ. [6]<br />

dt<br />

Thus the Laplacian Ò 2 p Œ of the stable peatland surface p Œ<br />

is minus the average n<strong>et</strong> precipitation divided by the average<br />

transmissivity<br />

Ò 2 p Œ = ≠ ÈqnÍ<br />

ÈT Í . [7]<br />

We can compute the stable topography of any tropic<strong>al</strong><br />

peatland by solving Poisson’s equation (Eqn. 7) for the stable<br />

peat surface morphology p Πusing the appropriate Laplacian<br />

v<strong>al</strong>ue for that climate. The average transmissivity ÈT Í is a<br />

complicated function of the tempor<strong>al</strong> pattern of rainf<strong>al</strong>l and<br />

the hydrologic-biologic<strong>al</strong> system. However, for any rainf<strong>al</strong>l<br />

regime, one can find the stable surface Laplacian Ò 2 p Œ by<br />

repeatedly simulating water table fluctuations (Eqn. 5) with a<br />

tri<strong>al</strong> Laplacian Ò 2 p, and adjusting the Laplacian v<strong>al</strong>ue until<br />

peat production b<strong>al</strong>ances decomposition (Eqn. 3) everywhere<br />

in the peatland (SI Text). In this way, one finds a shape<br />

param<strong>et</strong>er (Ò 2 p Œ) that describes stable peatland morphology<br />

under a given rainf<strong>al</strong>l regime in any drainage n<strong>et</strong>work.<br />

Climate and drainage n<strong>et</strong>work d<strong>et</strong>ermine tropic<strong>al</strong> peatland carbon<br />

storage capacity. By specifying the stable peatland topography,<br />

the uniform Laplacian principle gives the peat carbon storage<br />

capacity inside any drainage boundary and in any given<br />

climate. The volume under the surface satisfying Poisson’s<br />

equation times the mean carbon density of the peat gives the<br />

carbon storage capacity of the peatland. For example, the<br />

peat dome at our primary site currently has a mean peat<br />

depth of 3.88 m (max 4.92), and stores about 1535 t C ha ≠1 ;<br />

however, if the climate remains similar to the climate during<br />

its 2300-year development, we predict that in about 2500 y it<br />

will reach a stable shape with a mean peat depth of 4.54 m<br />

(max 7.10 m) and store 1800 t C ha ≠1 (Fig. S3; simulations of<br />

dynamics are described in the next section).<br />

The uniformity of the stable peat surface Laplacian is an<br />

approximation that requires that (1) peat accumulation rate<br />

ˆp/ˆt is a non-decreasing function of water table height; (2)<br />

flow of water is proportion<strong>al</strong> to water table gradient (Boussinesq’s<br />

equation); and (3) transmissivity is independent of location<br />

because flow through deep peat is negligible compared to<br />

DRAFT<br />

near-surface flow. In re<strong>al</strong>ity, groundwater flow through deeper<br />

peat will result in a deviation of the stable peat dome surface<br />

from the uniform Laplacian shape in very large peat domes.<br />

Specific<strong>al</strong>ly, groundwater flow through deep, low-permeability<br />

peat will tend to flatten the dome center, because of slow<br />

infiltration of water into the deep peat, and steepen the dome<br />

margin, because of exfiltration of water back into the high permeability<br />

near-surface peat near the boundary. Deep groundwater<br />

flow should be manifested as a downward (dome center)<br />

or upward (dome margin) trend in the water table during<br />

nights without rain when the water table is low; no such trend<br />

is apparent in our piezom<strong>et</strong>er data (Fig. 5d), suggesting that<br />

deep groundwater flow is sm<strong>al</strong>l. A sm<strong>al</strong>l deep groundwater flow<br />

term is further supported by radiocarbon dating of porewater<br />

DOC at our site (31), which suggests a maximum downward<br />

velocity of water of about one m<strong>et</strong>er per year, or at most a<br />

1.4 mm water table decline during a single twelve-hour night,<br />

one-sixteenth of the 22 mm water table decline from evapotranspiration<br />

during the day (Fig. 5). (Evapotranspirative<br />

683<br />

684<br />

685<br />

686<br />

687<br />

688<br />

689<br />

690<br />

691<br />

692<br />

693<br />

694<br />

695<br />

696<br />

697<br />

698<br />

699<br />

700<br />

701<br />

702<br />

703<br />

704<br />

705<br />

706<br />

707<br />

708<br />

709<br />

710<br />

711<br />

712<br />

713<br />

714<br />

715<br />

716<br />

717<br />

718<br />

719<br />

720<br />

721<br />

722<br />

723<br />

724<br />

725<br />

726<br />

727<br />

728<br />

729<br />

730<br />

731<br />

732<br />

733<br />

734<br />

735<br />

736<br />

737<br />

738<br />

739<br />

740<br />

741<br />

742<br />

743<br />

744<br />

6 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

<strong>Cobb</strong> <strong>et</strong> <strong>al</strong>.


745<br />

746<br />

747<br />

748<br />

749<br />

750<br />

751<br />

752<br />

753<br />

754<br />

755<br />

756<br />

757<br />

758<br />

759<br />

760<br />

761<br />

762<br />

763<br />

764<br />

765<br />

766<br />

767<br />

768<br />

769<br />

770<br />

771<br />

772<br />

773<br />

774<br />

775<br />

776<br />

777<br />

778<br />

779<br />

780<br />

781<br />

782<br />

783<br />

784<br />

785<br />

786<br />

787<br />

788<br />

789<br />

790<br />

791<br />

792<br />

793<br />

794<br />

795<br />

796<br />

797<br />

798<br />

799<br />

800<br />

801<br />

802<br />

803<br />

804<br />

805<br />

806<br />

Land surface elevation p, m<br />

a<br />

b<br />

Simulated peat age,<br />

c<strong>al</strong> y BP × 1000<br />

6<br />

3<br />

0 Clay<br />

2<br />

1<br />

4<br />

0<br />

0<br />

1 2<br />

Peat sample age, c<strong>al</strong> y BP × 1000<br />

Modern peat surface (LiDAR)<br />

2 0<br />

Distance from river, km<br />

Age at 25–67 cm depth,<br />

c<strong>al</strong> y BP × 1000<br />

c<br />

1.0<br />

0.5<br />

0<br />

Simulated peat surface vs time<br />

Primary site<br />

Deforested site<br />

Radiocarbon-dated<br />

core sample<br />

River<br />

3 2 1 0<br />

Distance from river, km<br />

Fig. 7. Morphogenesis of Mendaram peat dome. a, Shape of peat dome over time,<br />

including modeled peat surface (contours), modern peat surface from LiDAR (dashed<br />

black line), and c<strong>al</strong>ibrated radiocarbon dates from peat core samples (colored points).<br />

The deepest peat layers prior to 2250 c<strong>al</strong> y BP represent a uniformly deposited<br />

mangrove peat on a gently sloping clay plain (27). b, Simulated age of peat vs.<br />

c<strong>al</strong>ibrated radiocarbon ages from samples in the Mendaram peat dome. c, Age of<br />

sh<strong>al</strong>low peat samples (25 cm–65 cm depth) vs. distance from river at primary site<br />

(solid markers) and a nearby deforested site (open markers). Note the old peat near<br />

the surface close to the river as predicted by the model.<br />

flux is about one-tenth of the rate of decline of the water table<br />

from evapotranspiration because about one tenth of the deep<br />

peat cross-section is available for water flow; see specific yield<br />

curve, Fig. 5b.)<br />

A shape param<strong>et</strong>er related to our stable peatland Laplacian<br />

(Eqn. 7) appeared in Ingram’s model for temperate peatland<br />

morphology (10) assuming constant precipitation, uniform<br />

hydraulic conductivity, and simple river geom<strong>et</strong>ry (Ingram’s<br />

param<strong>et</strong>er is n<strong>et</strong> precipitation divided by hydraulic conductivity,<br />

instead of average transmissivity). Our result is more<br />

gener<strong>al</strong>, because it handles varying rainf<strong>al</strong>l and arbitrary landscapes,<br />

but is <strong>al</strong>so mathematic<strong>al</strong>ly simpler, because of our<br />

finding that transmissivity in tropic<strong>al</strong> peatlands is approximately<br />

independent of peat depth.<br />

Dynamics of tropic<strong>al</strong> peatland topography and carbon fluxes.<br />

Peat accumulation param<strong>et</strong>ers regulate dome dynamics. Our an<strong>al</strong>ysis<br />

shows how the rate of peat production f p and decomposition<br />

rate constant – a ect both the stable morphology and the<br />

dynamics of tropic<strong>al</strong> peat domes. These param<strong>et</strong>ers of the<br />

peat accumulation function (Eqn. 2) have an indirect but<br />

strong e ect on the stable peat surface Laplacian and hence<br />

peatland carbon storage capacity via the mean transmissivity<br />

ÈT Í (Eqn. 7) because the mean water table depth must be<br />

equ<strong>al</strong> to the ratio of the peat production rate to the decomposition<br />

rate constant (f p/–; Eqn. 3). A higher decomposition<br />

rate constant implies a higher mean water table in stable<br />

peat domes, meaning a higher transmissivity, a sm<strong>al</strong>ler stable<br />

surface Laplacian, and less carbon storage. If both peat<br />

production f p and the decomposition rate constant – increase<br />

tog<strong>et</strong>her, carbon storage capacity does not change, but peat<br />

dome dynamics are faster.<br />

Fit param<strong>et</strong>ers match literature data. Peat accumulation param<strong>et</strong>ers<br />

fit to the topography of a peat dome at our Brunei field<br />

site agree with published data from other sites, and <strong>al</strong>so with<br />

our other field data (see next section). We obtained peat<br />

accumulation param<strong>et</strong>ers f p, – by simulating the evolution of<br />

the dome (Fig. 7) and minimizing the least-squared di erence<br />

b<strong>et</strong>ween the simulated peat surface and the modern peat surface<br />

measured by LiDAR. We then compared our c<strong>al</strong>ibrated<br />

peat accumulation function to literature data on subsidence<br />

in drained, veg<strong>et</strong>ated peat swamps (4, 22). Our linear peat<br />

accumulation function was not c<strong>al</strong>ibrated to these subsidence<br />

data from the literature—only to the modern peat surface—<br />

but non<strong>et</strong>heless matched the subsidence data <strong>al</strong>most exactly<br />

(Fig. 4a; f p =1.46 mm y ≠1 , – =1.80 d ≠1 ). Our soil CO 2<br />

chamber measurements were <strong>al</strong>so very similar to those from<br />

other sites, suggesting that the e ect of water table on fluxes<br />

is similar at our site and in other tropic<strong>al</strong> peatlands (Fig. 4b).<br />

The uniform Laplacian principle predicts a centr<strong>al</strong> bog plain, and old<br />

peat near the surface at bog margins. We find that a tropic<strong>al</strong> peat<br />

dome reaches its stable shape first at its boundaries, because<br />

the stable dome surface is lowest there (Fig. 7,8,S4). Meanwhile,<br />

the interior of the peat dome continues growing at an<br />

approximately uniform rate, forming a relatively flat (sm<strong>al</strong>ler<br />

magnitude Laplacian) centr<strong>al</strong> “bog plain.” The veg<strong>et</strong>ation of<br />

tropic<strong>al</strong> bog plains may not be distinct (1), unlike the unforested<br />

bog plains of high-latitude peatlands (21); instead,<br />

we define the bog plain of a tropic<strong>al</strong> peat dome as the centr<strong>al</strong><br />

region that has not y<strong>et</strong> reached its stable Laplacian. While<br />

the dome center continues to accumulate peat and sequester<br />

carbon, the margin has reached its stable shape and stopped<br />

growing, so peat near the surface is older there.<br />

Older peat near dome margins had not been predicted<br />

before, so we collected 21 addition<strong>al</strong> radiocarbon dates from<br />

bas<strong>al</strong> and near-surface peat samples to test this prediction.<br />

These radiocarbon dates confirmed that near-surface peat was<br />

older near dome margins than at the same depths towards<br />

the interior of the same domes (Fig. 7c). We <strong>al</strong>so compared<br />

radiocarbon dates in deeper peat to simulated ages at the<br />

same locations and depths, excluding bas<strong>al</strong> samples from the<br />

mangrove peat prior to the establishment of the peat swamp<br />

forest (Fig. 7, SI Text; (1, 27)). Radiocarbon dates and<br />

simulated ages at the same locations and depths matched<br />

well (Fig. 7b). We did not expect radiocarbon dates from<br />

cores to match simulated peat ages exactly because (1) the<br />

drainage n<strong>et</strong>work may have shifted during the 2300 years of<br />

dome growth; (2) tree root growth may inject young carbon<br />

into peat below the surface; and (3) tree f<strong>al</strong>ls in peat swamp<br />

forests remove older peat to form tip-up pools which then fill<br />

with younger peat. In an earlier study, we estimated that<br />

replacement of older peat by younger peat in tip-up pools<br />

would bias radiocarbon dates of deep peat to about 500 years<br />

later than when materi<strong>al</strong> was first deposited in that stratum<br />

(Fig. 11 in (27)), consistent with the o s<strong>et</strong> b<strong>et</strong>ween measured<br />

radiocarbon dates and ages simulated by our model (Fig. 7b).<br />

DRAFT<br />

Carbon sequestration rate is proportion<strong>al</strong> to bog plain area. The centrip<strong>et</strong><strong>al</strong><br />

pattern of dome development makes the rate of carbon<br />

sequestration roughly proportion<strong>al</strong> to the area of the centr<strong>al</strong><br />

807<br />

808<br />

809<br />

810<br />

811<br />

812<br />

813<br />

814<br />

815<br />

816<br />

817<br />

818<br />

819<br />

820<br />

821<br />

822<br />

823<br />

824<br />

825<br />

826<br />

827<br />

828<br />

829<br />

830<br />

831<br />

832<br />

833<br />

834<br />

835<br />

836<br />

837<br />

838<br />

839<br />

840<br />

841<br />

842<br />

843<br />

844<br />

845<br />

846<br />

847<br />

848<br />

849<br />

850<br />

851<br />

852<br />

853<br />

854<br />

855<br />

856<br />

857<br />

858<br />

859<br />

860<br />

861<br />

862<br />

863<br />

864<br />

865<br />

866<br />

867<br />

868<br />

<strong>Cobb</strong><br />

<strong>et</strong> <strong>al</strong>.<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 7


869<br />

870<br />

871<br />

872<br />

873<br />

874<br />

875<br />

876<br />

877<br />

878<br />

879<br />

880<br />

881<br />

882<br />

883<br />

884<br />

885<br />

886<br />

887<br />

888<br />

889<br />

890<br />

891<br />

892<br />

893<br />

894<br />

895<br />

896<br />

897<br />

898<br />

899<br />

900<br />

901<br />

902<br />

903<br />

904<br />

905<br />

906<br />

907<br />

908<br />

909<br />

910<br />

911<br />

912<br />

913<br />

914<br />

915<br />

916<br />

917<br />

918<br />

919<br />

920<br />

921<br />

922<br />

923<br />

924<br />

925<br />

926<br />

927<br />

928<br />

929<br />

930<br />

approx.<br />

5 m<br />

Effect of peat depth<br />

on transmissivity<br />

is negligible<br />

Prediction of ultimate<br />

CO 2 sequestration<br />

from boundary and climate<br />

growing<br />

rain - ET<br />

stable<br />

approx. 10 km<br />

Centrip<strong>et</strong><strong>al</strong><br />

convergence<br />

to stable shape<br />

CO 2 sequestration rate<br />

proportion<strong>al</strong> to area<br />

of centr<strong>al</strong> bog plain<br />

Fig. 8. Model of tropic<strong>al</strong> peat dome development. The surface p of a tropic<strong>al</strong> peat<br />

dome evolves towards a shape compl<strong>et</strong>ely described by a uniform surface Laplacian<br />

Ò 2 p Œ given by the ratio of average n<strong>et</strong> precipitation Èq nÍ to average hydraulic<br />

transmissivity ÈT Í. The surface Laplacian Ò 2 p Œ defines the stable shape and<br />

carbon storage capacity of a peat dome inside any drainage boundary. When the<br />

dome surface has a uniform Laplacian, the water table height fluctuates uniformly,<br />

and peat production is b<strong>al</strong>anced by decomposition everywhere in the dome. When<br />

a peat dome is growing, it sequesters carbon at a rate proportion<strong>al</strong> to the area of a<br />

flatter (sm<strong>al</strong>ler-magnitude surface Laplacian) area in the middle, the centr<strong>al</strong> bog plain.<br />

Gray boxes, established results; black boxes, findings presented here.<br />

bog plain (Fig. 8). Under a given climate, the rate of sequestration<br />

decreases as the dome approaches its stable shape and the<br />

centr<strong>al</strong> region of peat accumulation—the bog plain—shrinks in<br />

area. For example, our simulations imply that the current rate<br />

of CO 2 sequestration at our site (0.80 t ha ≠1 y ≠1 ,100-year<br />

average) is less than a quarter of its initi<strong>al</strong> rate about 2300<br />

years ago (3.81 t ha ≠1 y ≠1 ), and CO 2 sequestration is more<br />

than five times faster at the dome interior (1.89 t ha ≠1 y ≠1 ,<br />

6.37 km from river) than at its edge (0.36 t ha ≠1 y ≠1 ,1km<br />

from river; Fig. S3). The mechanism of tropic<strong>al</strong> peat dome<br />

development that we describe therefore creates landscape-sc<strong>al</strong>e<br />

patterns in loc<strong>al</strong> carbon fluxes and radiocarbon date profiles.<br />

Loc<strong>al</strong> measurements of carbon fluxes or radiocarbon dates cannot<br />

be upsc<strong>al</strong>ed to region<strong>al</strong> fluxes without considering dome<br />

morphology because the flatter interior of each peat dome<br />

sequesters carbon while the margins do not (Fig. 8). Old peat<br />

near the peatland surface (2), <strong>al</strong>though in some cases caused<br />

by loc<strong>al</strong> climate change or disturbance, <strong>al</strong>so can be expected<br />

at the margin of any peat dome.<br />

Future effects of changes in drainage n<strong>et</strong>works and climate.<br />

Our an<strong>al</strong>ysis provides a simple way of predicting long-term<br />

change in peat dome morphology and carbon storage in response<br />

to changes in drainage n<strong>et</strong>work, climate or sea level<br />

because the stable peat surface Laplacian compl<strong>et</strong>ely specifies<br />

the stable peat topography with given drainage boundary<br />

conditions. If the drainage n<strong>et</strong>work changes, we can solve<br />

Poisson’s equation in the new drainage boundary to compute<br />

the gain or loss of peat, and the n<strong>et</strong> carbon emissions, as the<br />

peat surface approaches its new stable topography. If the<br />

climate changes, we can compute a new stable Laplacian v<strong>al</strong>ue<br />

for the new climatic conditions, and d<strong>et</strong>ermine how much a<br />

currently stable peatland will grow or subside.<br />

Subdivision of a peatland by drainage can<strong>al</strong>s reduces carbon storage.<br />

The average surface elevation of a stable peat dome is<br />

proportion<strong>al</strong> to the area of the dome because of the uniform<br />

Laplacian principle. If we sc<strong>al</strong>e the area of a peat dome by<br />

some factor k by multiplying both x and y coordinates by Ô k,<br />

the surface elevation p must increase by the same factor k<br />

to keep the same Laplacian. Therefore, the carbon storage<br />

capacity of a peat dome sc<strong>al</strong>es with its area. For example, a<br />

peat dome that is cut into h<strong>al</strong>ves of approximately the same<br />

shape as the origin<strong>al</strong> dome will have one-h<strong>al</strong>f the carbon storage<br />

capacity (h<strong>al</strong>f the mean stable peat depth) of the origin<strong>al</strong><br />

dome. This provides a straightforward way to estimate the<br />

long-term impacts of artifici<strong>al</strong> drainage n<strong>et</strong>works that are now<br />

a ecting over 50% of the peatlands of Southeast Asia (32)<br />

and from which a robust quantification of carbon emissions is<br />

urgently needed (6).<br />

The dynamic response of a peat dome to changes in rainf<strong>al</strong>l<br />

and sea level <strong>al</strong>so depends on its area because of the centrip<strong>et</strong><strong>al</strong><br />

pattern of dome development (Fig. 8). Because of their higher<br />

stable mean depth, larger-area domes reach their stable shape<br />

more slowly than sm<strong>al</strong>ler-area domes.<br />

Relative effects of climate change on carbon storage capacity are independent<br />

of drainage n<strong>et</strong>work. Although peatland drainage n<strong>et</strong>works<br />

play a centr<strong>al</strong> role in d<strong>et</strong>ermining absolute carbon storage<br />

and dynamics, we can c<strong>al</strong>culate the proportion<strong>al</strong> e ect of<br />

climate change on long-term carbon storage of a tropic<strong>al</strong> peatland<br />

independent of the drainage n<strong>et</strong>work. Poisson’s equation<br />

(Eqn. 7) must be solved in each drainage boundary to obtain<br />

the topography of the stable peat surface. However, we can<br />

then predict the e ects of changes in climate independent of<br />

the drainage n<strong>et</strong>work because of the linearity of the Laplacian<br />

operator. By the definition of linearity for a mathematic<strong>al</strong><br />

operator, a peat surface Laplacian that is larger by some factor<br />

k corresponds to a peat surface that is vertic<strong>al</strong>ly str<strong>et</strong>ched<br />

by the same factor (kÒ 2 p = Ò 2 kp), and which therefore has<br />

a mean peat depth that is larger by the same factor. Thus,<br />

carbon storage capacity per area ¯p Œ is proportion<strong>al</strong> to the<br />

stable peat surface Laplacian ¯p Œ ÃÒ 2 p Œ.<br />

DRAFT<br />

Dynamic simulations converge to new stable morphologies after<br />

changes in conditions. Our simulations of peat dome dynamics<br />

demonstrate the convergence of initi<strong>al</strong>ly stable domes to<br />

new, stable, uniform-Laplacian morphologies after perturbations<br />

(Fig. 9). The simulations show the e ect of increased<br />

tot<strong>al</strong> rainf<strong>al</strong>l (Fig. 9a,e), which is a recognized climate feedback<br />

for tropic<strong>al</strong> peatlands (12), and <strong>al</strong>so show that artifici<strong>al</strong><br />

drainage for agriculture (Fig. 9d) can dominate <strong>al</strong>l natur<strong>al</strong><br />

feedbacks if not curtailed (4, 16). In addition, our simulations<br />

demonstrate a third feedback: the increase in rainf<strong>al</strong>l variability<br />

from warming climates (33) can cause peat loss if not<br />

compensated by an increase in tot<strong>al</strong> rainf<strong>al</strong>l (Fig. 9c,f). For<br />

these simulations, we generated new rainf<strong>al</strong>l time series as similar<br />

to current rainf<strong>al</strong>l as possible but with larger annu<strong>al</strong> and<br />

El Niño–Southern Oscillation (ENSO) fluctuations (Fig S1a,b;<br />

SI Text). Either greater season<strong>al</strong>ity or a stronger ENSO decreased<br />

peatland carbon storage capacity, but an increase in<br />

season<strong>al</strong>ity had a larger maximum e ect, partly because the<br />

magnitude of the ENSO fluctuation is sm<strong>al</strong>ler. In contrast, sea<br />

level rise could drive peat accumulation in the long term by<br />

elevating the tid<strong>al</strong> rivers draining most peat domes (Fig. 9b,e).<br />

In gener<strong>al</strong>, losses can be much more rapid than accumulation<br />

931<br />

932<br />

933<br />

934<br />

935<br />

936<br />

937<br />

938<br />

939<br />

940<br />

941<br />

942<br />

943<br />

944<br />

945<br />

946<br />

947<br />

948<br />

949<br />

950<br />

951<br />

952<br />

953<br />

954<br />

955<br />

956<br />

957<br />

958<br />

959<br />

960<br />

961<br />

962<br />

963<br />

964<br />

965<br />

966<br />

967<br />

968<br />

969<br />

970<br />

971<br />

972<br />

973<br />

974<br />

975<br />

976<br />

977<br />

978<br />

979<br />

980<br />

981<br />

982<br />

983<br />

984<br />

985<br />

986<br />

987<br />

988<br />

989<br />

990<br />

991<br />

992<br />

8 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

<strong>Cobb</strong> <strong>et</strong> <strong>al</strong>.


993<br />

994<br />

995<br />

996<br />

997<br />

998<br />

999<br />

1000<br />

1001<br />

1002<br />

1003<br />

1004<br />

1005<br />

1006<br />

1007<br />

1008<br />

1009<br />

1010<br />

1011<br />

1012<br />

1013<br />

1014<br />

1015<br />

1016<br />

1017<br />

1018<br />

1019<br />

1020<br />

1021<br />

1022<br />

1023<br />

1024<br />

1025<br />

1026<br />

1027<br />

1028<br />

1029<br />

1030<br />

1031<br />

1032<br />

1033<br />

1034<br />

1035<br />

1036<br />

1037<br />

1038<br />

1039<br />

1040<br />

1041<br />

1042<br />

1043<br />

1044<br />

1045<br />

1046<br />

1047<br />

1048<br />

1049<br />

1050<br />

1051<br />

1052<br />

1053<br />

1054<br />

a<br />

c<br />

e<br />

Average peat<br />

depth , m<br />

f<br />

0<br />

CO 2 flux,<br />

t ha -1 y -1<br />

1.0<br />

0.5<br />

0.0<br />

2<br />

0<br />

-2<br />

Response of peat topography to perturbations<br />

sea level rise<br />

more rain<br />

b<br />

initi<strong>al</strong> stable surface<br />

increased season<strong>al</strong>ity<br />

drainage<br />

300<br />

sea level rise<br />

drainage<br />

increased season<strong>al</strong>ity<br />

sea level rise<br />

Time, y<br />

more rain<br />

increased season<strong>al</strong>ity<br />

0 200 400<br />

Time, y<br />

d<br />

drainage<br />

600 900<br />

Fig. 9. Dynamic effects of climate change on carbon storage in tropic<strong>al</strong> peatlands.<br />

a–d, Simulated peat surface elevation vs. time of an initi<strong>al</strong>ly stable peat dome<br />

after different perturbations. The dashed line indicates the stable morphology for the<br />

peat dome b<strong>et</strong>ween two par<strong>al</strong>lel rivers, colored lines give the peat dome morphology<br />

at subsequent time steps. a, Annu<strong>al</strong> rainf<strong>al</strong>l increase from 2237 to 2430 mm / y<br />

causes peat accumulation until the peat dome reaches a new stable morphology.<br />

b, Sea level rise of 0.5 m leads to an upward shift in peat surface elevation as tid<strong>al</strong><br />

rivers bounding the peat dome rise. c, Increase in season<strong>al</strong> fluctuation in rainf<strong>al</strong>l<br />

from 902 to 1095 mm / y causes loss of peat. d, Sustained drainage to a depth<br />

of 50 cm drives rapid peat loss from aerobic decomposition. e, Spati<strong>al</strong>ly-averaged<br />

peat depth vs. time for simulations with more rain (a, long dashes), sea level rise (b,<br />

dot-dash), increased season<strong>al</strong>ity (c, dot-dot-dash) drainage (d, short dashes), or with<br />

no change in conditions (solid) or increased ENSO sign<strong>al</strong> (long dot-dash). f, Average<br />

CO 2 emission (negative) or sequestration (positive) vs. time for simulations as in (e).<br />

Because peat is mostly organic carbon, peat accumulation or loss causes uptake or<br />

release of carbon, respectively. The initi<strong>al</strong> CO 2 emission for the drainage scenario is<br />

off the chart at ≠24 t ha ≠1 y ≠1 .<br />

(Fig. 9e), because subsidence of drained peatlands can be far<br />

faster than typic<strong>al</strong> accumulation rates (4). For example, the<br />

estimated area-averaged current CO 2 sequestration rate at our<br />

site is 0.80 t ha ≠1 y ≠1 , whereas Hooijer <strong>et</strong> <strong>al</strong>. estimated CO 2<br />

emissions of at least 73 t ha ≠1 y ≠1 from tropic<strong>al</strong> peatlands<br />

under plantation agriculture (5).<br />

Intermittency of rainf<strong>al</strong>l reduces tropic<strong>al</strong> peatland carbon storage.<br />

We find that fluctuations in n<strong>et</strong> precipitation on time sc<strong>al</strong>es<br />

from hours to years can reduce long-term peat accumulation.<br />

We further explored the e ects of variability in rainf<strong>al</strong>l seen<br />

in our dynamic simulations (Fig. 9) by computing the e ect<br />

of interstorm arriv<strong>al</strong> time, annu<strong>al</strong> and ENSO fluctuations on<br />

peatland carbon storage capacity (Fig. 10). The simulations<br />

demonstrate that long-term peat accumulation is controlled<br />

by variation in rainf<strong>al</strong>l, not only by mean rainf<strong>al</strong>l, because<br />

fluctuations in the water table cause exponenti<strong>al</strong> changes in<br />

groundwater flow. The high outward flow during peak water<br />

tables is not compensated by low flow rates after the water<br />

table declines. For example, a steady drizzle at the same<br />

average intensity as the intermittent rainf<strong>al</strong>l actu<strong>al</strong>ly observed<br />

at our site would sustain more than 10 times more long-term<br />

carbon storage (19.5 kt ha ≠1 vs. 1.80 kt ha ≠1 ; Fig. S1d,e).<br />

The intermittency of tropic<strong>al</strong> convective storms significantly<br />

a ects long-term carbon storage: carbon storage capacity can<br />

a<br />

320<br />

Mean annu<strong>al</strong> rainf<strong>al</strong>l, cm<br />

240<br />

160<br />

Norm<strong>al</strong>ized carbon<br />

storage capacity<br />

b 4<br />

Annu<strong>al</strong> amplitude, mm d -1<br />

0<br />

0.4 0.8 1.2 0.0 0.5 1.0<br />

Interstorm arriv<strong>al</strong> time, d ENSO amplitude, mm d -1<br />

DRAFT<br />

2<br />

Norm<strong>al</strong>ized carbon<br />

storage capacity<br />

increased<br />

season<strong>al</strong>ity<br />

Fig. 10. Effects of climate change on carbon storage capacity of tropic<strong>al</strong> peatlands.<br />

a, Simulated carbon storage capacity (contours) versus time-averaged rainf<strong>al</strong>l<br />

and interv<strong>al</strong> b<strong>et</strong>ween storms in a simple rainf<strong>al</strong>l model (Poisson process for storm<br />

incidents, exponenti<strong>al</strong>ly distributed rain depth per storm). The b<strong>al</strong>ance b<strong>et</strong>ween<br />

rainf<strong>al</strong>l and groundwater flow s<strong>et</strong>s a limit on the curvature of the peat surface, and<br />

therefore limits the amount of carbon that can be stored as peat in a peatland. This<br />

carbon storage capacity is proportion<strong>al</strong> to the Laplacian of the stable peat surface<br />

elevation (Results and Discussion), so the relative effect of changes in rainf<strong>al</strong>l patterns<br />

on carbon storage capacity can be c<strong>al</strong>culated independent of the drainage n<strong>et</strong>work.<br />

Higher rainf<strong>al</strong>l increases carbon storage capacity, whereas increased time b<strong>et</strong>ween<br />

storms reduces it. b, Carbon storage capacity (contours) as in (a), but driven by<br />

rainf<strong>al</strong>l at our site (diamond), or with a weakened or strengthened annu<strong>al</strong> or El Niño-<br />

Southern Oscillation fluctuation in rainf<strong>al</strong>l. The vertic<strong>al</strong> shift to lower carbon storage<br />

with increased annu<strong>al</strong> variation in rainf<strong>al</strong>l (up arrow) corresponds to the simulated<br />

effect of increased season<strong>al</strong>ity in Fig. 9.<br />

decrease by a third depending on wh<strong>et</strong>her convective storms<br />

arrive every fourteen hours on average, as at our site, or every<br />

twenty four hours, with the same mean rainf<strong>al</strong>l (Fig. 10a).<br />

Our simulations with smoothed rainf<strong>al</strong>l intensity and evapotranspiration<br />

show that models must consider the e ects<br />

of subdiurn<strong>al</strong> fluctuations in rainf<strong>al</strong>l to correctly predict the<br />

long-term evolution and carbon storage of tropic<strong>al</strong> peatlands.<br />

The exact d<strong>et</strong>ails of the fluctuations in rainf<strong>al</strong>l are not important,<br />

in the sense that many distinct rainf<strong>al</strong>l time series<br />

can give the same stable surface Laplacian, and the same<br />

carbon storage capacity. However, carbon storage capacity<br />

can be severely overestimated by simulations that entirely<br />

ignore the e ects of fluctuations in rainf<strong>al</strong>l. We explored the<br />

e ects of neglecting fluctuations in rainf<strong>al</strong>l by computing the<br />

stable surface Laplacian after averaging n<strong>et</strong> precipitation on<br />

hourly and longer interv<strong>al</strong>s. Treating rainf<strong>al</strong>l intensity and<br />

evapotranspiration as constant each hour, instead of every 20<br />

minutes, increased the simulated stable surface Laplacian by<br />

a few percent, but averaging over a day led to an overestimate<br />

by 20%, a week 100%, a month 400%, and a year more than<br />

1000% (Fig. S1d,e).<br />

Conclusions<br />

The mathematic<strong>al</strong> and numeric<strong>al</strong> models presented here predict<br />

the long-term e ects of changes in rainf<strong>al</strong>l regimes and<br />

drainage n<strong>et</strong>works on the morphology of tropic<strong>al</strong> peat domes.<br />

Because tropic<strong>al</strong> peat domes are mostly organic carbon, these<br />

predictions of peat dome morphogenesis <strong>al</strong>so quantify peat<br />

dome carbon storage capacity and carbon fluxes. Our approach<br />

shows that tropic<strong>al</strong> peatlands approach a limiting shape in<br />

which the Laplacian of the land surface is uniform. This stable<br />

peatland surface Laplacian can be computed from any rainf<strong>al</strong>l<br />

time series, and compl<strong>et</strong>ely summarizes the e ects of the rain-<br />

1055<br />

1056<br />

1057<br />

1058<br />

1059<br />

1060<br />

1061<br />

1062<br />

1063<br />

1064<br />

1065<br />

1066<br />

1067<br />

1068<br />

1069<br />

1070<br />

1071<br />

1072<br />

1073<br />

1074<br />

1075<br />

1076<br />

1077<br />

1078<br />

1079<br />

1080<br />

1081<br />

1082<br />

1083<br />

1084<br />

1085<br />

1086<br />

1087<br />

1088<br />

1089<br />

1090<br />

1091<br />

1092<br />

1093<br />

1094<br />

1095<br />

1096<br />

1097<br />

1098<br />

1099<br />

1100<br />

1101<br />

1102<br />

1103<br />

1104<br />

1105<br />

1106<br />

1107<br />

1108<br />

1109<br />

1110<br />

1111<br />

1112<br />

1113<br />

1114<br />

1115<br />

1116<br />

<strong>Cobb</strong><br />

<strong>et</strong> <strong>al</strong>.<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 9


1117<br />

1118<br />

1119<br />

1120<br />

1121<br />

1122<br />

1123<br />

1124<br />

1125<br />

1126<br />

1127<br />

1128<br />

1129<br />

1130<br />

1131<br />

1132<br />

1133<br />

1134<br />

1135<br />

1136<br />

1137<br />

1138<br />

1139<br />

1140<br />

1141<br />

1142<br />

1143<br />

1144<br />

1145<br />

1146<br />

1147<br />

1148<br />

1149<br />

1150<br />

1151<br />

1152<br />

1153<br />

1154<br />

1155<br />

1156<br />

1157<br />

1158<br />

1159<br />

1160<br />

1161<br />

1162<br />

1163<br />

1164<br />

1165<br />

1166<br />

1167<br />

1168<br />

1169<br />

1170<br />

1171<br />

1172<br />

1173<br />

1174<br />

1175<br />

1176<br />

1177<br />

1178<br />

f<strong>al</strong>l pattern on the stable morphology and storage capacity of<br />

carbon within the peatland drainage boundary.<br />

The uniform Laplacian principle is supported by a range of<br />

observations: (1) the peat surface Laplacian is approximately<br />

uniform in a region near the dome edge (Fig. 6c); (2) water<br />

table behavior is uniform where the surface Laplacian is uniform,<br />

and is di erent in the dome interior (Fig. 5a); (3) water<br />

table behavior is the same in areas with di ering gradients<br />

within the uniform-Laplacian region (Fig. 5a); (4) transmissivity<br />

increases exponenti<strong>al</strong>ly at high water tables, so that loc<strong>al</strong><br />

water b<strong>al</strong>ance is dominated by flow near the surface (Fig. 5c);<br />

and (5) peat accumulation param<strong>et</strong>ers match literature data,<br />

even though those data were not used for c<strong>al</strong>ibration (Fig. 4a).<br />

Our an<strong>al</strong>ysis underscores the importance of considering<br />

geomorphology when measuring and modeling carbon fluxes<br />

in tropic<strong>al</strong> peatlands. On a growing peat dome, the perim<strong>et</strong>er<br />

of the dome reaches a steady elevation first while centr<strong>al</strong><br />

areas continue to accumulate carbon (Fig. 8). This pattern<br />

of dome morphogenesis implies that the locations of groundtruth<br />

carbon flux measurements within tropic<strong>al</strong> peat domes<br />

are important considerations for earth system models (34).<br />

For example, measurements of carbon flux in the center of a<br />

growing dome overestimate the average flux for the whole dome,<br />

because peat accumulation is fastest in the center (Figs. 8,S3).<br />

The distribution of peat dome areas within a peatland complex<br />

1. Anderson JAR (1964) The structure and development of the peat swamps of Sarawak and<br />

Brunei. Journ<strong>al</strong> of Tropic<strong>al</strong> Geography 18:7–16.<br />

2. Dommain R, Couwenberg J, Joosten H (2011) Development and carbon sequestration of<br />

tropic<strong>al</strong> peat domes in south-east Asia: links to post-glaci<strong>al</strong> sea-level changes and Holocene<br />

climate variability. Quaternary Science Reviews 30:999–1010.<br />

3. Page SE, <strong>et</strong> <strong>al</strong>. (2002) The amount of carbon released from peat and forest fires in Indonesia<br />

during 1997. Nature 420:61–65.<br />

4. Couwenberg J, Dommain R, Joosten H (2010) Greenhouse gas fluxes from tropic<strong>al</strong> peatlands<br />

in Southeast Asia. Glob<strong>al</strong> Change Biology 16:1715–1732.<br />

5. Hooijer A, <strong>et</strong> <strong>al</strong>. (2012) Subsidence and carbon loss in drained tropic<strong>al</strong> peatlands. Biogeosciences<br />

9:1053–1071.<br />

6. C<strong>al</strong>le L, <strong>et</strong> <strong>al</strong>. (2016) Region<strong>al</strong> carbon fluxes from land use and land cover change in Asia,<br />

1980–2009. Environment<strong>al</strong> Research L<strong>et</strong>ters 074011:1–12.<br />

7. Page S, Hooijer A (2016) In the line of fire: the peatlands of Southeast Asia. Philosophic<strong>al</strong><br />

Transactions of the Roy<strong>al</strong> Soci<strong>et</strong>y B 371:20150176.<br />

8. Winston RB (1994) Models of the geomorphology, hydrology, and development of domed<br />

peat bodies. Geologic<strong>al</strong> Soci<strong>et</strong>y of America Bull<strong>et</strong>in 106:1594–1604.<br />

9. Hirano T, Jauhiainen J, Inoue T, Takahashi H (2009) Controls on the carbon b<strong>al</strong>ance of tropic<strong>al</strong><br />

peatlands. Ecosystems 12(6):873–887.<br />

10. Ingram HAP (1982) Size and shape in raised mire ecosystems - a geophysic<strong>al</strong> model. Nature<br />

297:300–303.<br />

11. Glaser PH, Hansen BCS, Siegel DI, Reeve AS, Morin PJ (2004) Rates, pathways and drivers<br />

for peatland development in the Hudson Bay Lowlands, northern Ontario, Canada. Journ<strong>al</strong><br />

of Ecology 92:1036–1053.<br />

12. Lottes AL, Ziegler AM (1994) World peat occurrence and the season<strong>al</strong>ity of climate and veg<strong>et</strong>ation.<br />

P<strong>al</strong>aeogeography, P<strong>al</strong>aeoclimatology, P<strong>al</strong>aeoecology 106:23–37.<br />

13. Gast<strong>al</strong>do RA (2010) Peat or no peat: Why do the Rajang and Mahakam Deltas differ? Internation<strong>al</strong><br />

Journ<strong>al</strong> of Co<strong>al</strong> Geology 83:162–172.<br />

14. Clymo RS (1984) The limits to peat bog growth. Philosophic<strong>al</strong> Transactions of the Roy<strong>al</strong><br />

Soci<strong>et</strong>y of London Series B-Biologic<strong>al</strong> Sciences 303:605–654.<br />

15. Hilbert D, Roul<strong>et</strong> N, Moore T (2000) Modelling and an<strong>al</strong>ysis of peatlands as dynamic<strong>al</strong> systems.<br />

Journ<strong>al</strong> of Ecology 88:230–242.<br />

16. Kurnianto S, <strong>et</strong> <strong>al</strong>. (2015) Carbon accumulation of tropic<strong>al</strong> peatlands over millennia: a modeling<br />

approach. Glob<strong>al</strong> Change Biology 21(1):431–444.<br />

17. Morris PJ, Baird AJ, Belyea LR (2012) The DigiBog peatland development model 2: ecohydrologic<strong>al</strong><br />

simulations in 2D. Ecohydrology 5:256–268.<br />

18. Baird AJ, Morris PJ, Belyea LR (2012) The DigiBog peatland development model 1: ration<strong>al</strong>e,<br />

conceptu<strong>al</strong> model, and hydrologic<strong>al</strong> basis. Ecohydrology 5:242–255.<br />

19. Jauhiainen J, Takahashi H, Heikkinen JEP, Martikainen PJ, Vasanders H (2005) Carbon<br />

fluxes from a tropic<strong>al</strong> peat swamp forest floor. Glob<strong>al</strong> Change Biology 11:1788–1797.<br />

20. Lampela M, <strong>et</strong> <strong>al</strong>. (2016) Ground surface microtopography and veg<strong>et</strong>ation patterns in a tropic<strong>al</strong><br />

peat swamp forest. Catena 139:127–136.<br />

21. Glaser PH, Janssens JA (1986) Raised bogs in eastern North America: transitions in landforms<br />

and gross stratigraphy. Canadian Journ<strong>al</strong> of Botany 64:395–415.<br />

22. Carlson KM, Goodman LK, May-Tobin CC (2015) Modeling relationships b<strong>et</strong>ween water table<br />

depth and peat soil carbon loss in Southeast Asian plantations. Environment<strong>al</strong> Research<br />

L<strong>et</strong>ters 10:074006.<br />

is <strong>al</strong>so important, because sm<strong>al</strong>ler domes reach their stable<br />

shapes faster after a change in conditions. Improved earth<br />

system models could use the uniform Laplacian principle to<br />

e ciently account for the e ects of changing rainf<strong>al</strong>l, sea level,<br />

and drainage on tropic<strong>al</strong> peat carbon storage given a re<strong>al</strong>istic<br />

distribution of peat dome sizes. The approach outlined here<br />

<strong>al</strong>so provides a framework for including the e ects of other<br />

long-term processes that remain understudied, such as shifts<br />

in river n<strong>et</strong>works, changes in tree community composition and<br />

s<strong>al</strong>t water intrusion from rising sea levels.<br />

ACKNOWLEDGMENTS. We thank Mahmud Yussof of Brunei<br />

Daruss<strong>al</strong>am Heart of Borneo Centre and the Brunei Daruss<strong>al</strong>am<br />

Ministry of Industry and Primary Resources for their support of<br />

this project; Hajah Jamilah J<strong>al</strong>il and Jo re Ali Ahmad of the<br />

Brunei Daruss<strong>al</strong>am Forestry Department for facilitation of field<br />

work and release of sta ;AmyChuaforlogistic<strong>al</strong>support;and<br />

Bernard Jun Long Ng, Rahayu Sukmaria binti Haji Sukri, Watu bin<br />

Awok, Azlan Pandai, Rosaidi Mureh, Muhammad Wafiuddin Zain<strong>al</strong><br />

Ari nandSylvainFerrantforfieldassistance.We<strong>al</strong>sothankPaul<br />

Glaser and two anonymous reviewers for their d<strong>et</strong>ailed comments<br />

on the manuscript. This research was supported by the Nation<strong>al</strong><br />

Research Foundation Singapore through the Singapore-MIT Alliance<br />

for Research and Technology’s Center for Environment<strong>al</strong> Sensing<br />

and Modeling interdisciplinary research program and by the USA<br />

Nation<strong>al</strong> Science Foundation under Grant Nos. 1114155 and 1114161<br />

to C.F.H.<br />

DRAFT<br />

23. Melling L, Hatano R, Goh KJ (2005) Soil CO 2 flux from three ecosystems in tropic<strong>al</strong> peatland<br />

of Sarawak, M<strong>al</strong>aysia. Tellus 57B:1–11.<br />

24. Ali M, Taylor D, Inubushi K (2006) Effects of environment<strong>al</strong> variations on CO 2 efflux from a<br />

tropic<strong>al</strong> peatland in eastern Sumatra. W<strong>et</strong>lands 26(2):612–618.<br />

25. Jauhiainen J, Hooijer A, Page SE (2012) Carbon dioxide emissions from an Acacia plantation<br />

on peatland in Sumatra, Indonesia. Biogeosciences 9:617–630.<br />

26. Hirano T, <strong>et</strong> <strong>al</strong>. (2012) Effects of disturbances on the carbon b<strong>al</strong>ance of tropic<strong>al</strong> peat swamp<br />

forests. Glob<strong>al</strong> Change Biology 18:3410–3422.<br />

27. Dommain R, <strong>et</strong> <strong>al</strong>. (2015) Forest dynamics and tip-up pools drive pulses of high carbon accumulation<br />

rates in a tropic<strong>al</strong> peat dome in Borneo (Southeast Asia). Journ<strong>al</strong> of Geophysic<strong>al</strong><br />

Research: Biogeosciences 120(4):617–640.<br />

28. Hooijer A (2005) Hydrology of tropic<strong>al</strong> w<strong>et</strong>land forests: recent research results from Sarawak<br />

peatswamps in Forests, Water and People in the Humid Tropics, eds. Bonell M, Bruijnzeel LA.<br />

(Cambridge University Press, Cambridge, UK), pp. 447–461.<br />

29. Dolan TJ, Hermann AJ, Bayley SE, Zoltek, Jr. J (1984) Evapotranspiration of a Florida, U.S.A.,<br />

freshwater w<strong>et</strong>land. Journ<strong>al</strong> of Hydrology 74:355–371.<br />

30. Baird AJ, <strong>et</strong> <strong>al</strong>. (2017) High permeability explains the vulnerability of the carbon store in<br />

drained tropic<strong>al</strong> peatlands. Geophysic<strong>al</strong> Research L<strong>et</strong>ters 44.<br />

31. Gandois L, <strong>et</strong> <strong>al</strong>. (2014) Origin, composition, and transformation of dissolved organic matter<br />

in tropic<strong>al</strong> peatlands. Geochimica <strong>et</strong> Cosmochimica Acta 137:35–47.<br />

32. Mi<strong>et</strong>tinen J, Shi C, Liew SC (2016) Land cover distribution in the peatlands of Peninsular<br />

M<strong>al</strong>aysia, Sumatra and Borneo in 2015 with changes since 1990. Glob<strong>al</strong> Ecology and Conservation<br />

6:67–78.<br />

33. Huang P, Xie SP, Hu K, Huang G, Huang R (2013) Patterns of the season<strong>al</strong> response of<br />

tropic<strong>al</strong> rainf<strong>al</strong>l to glob<strong>al</strong> warming. Nature Geoscience 6:357–361.<br />

34. Quéré CL, <strong>et</strong> <strong>al</strong>. (2009) Trends in the sources and sinks of carbon dioxide. Nature Geoscience<br />

2:831–836.<br />

35. Moore ID, Grayson RB (1991) Terrain-based catchment partitioning and runoff prediction<br />

using vector elevation data. Water Resources Research 27(6):1177–1191.<br />

36. Stuiver M, Reimer PJ (1993) Extended 14 C data base and revised CALIB 3.0 14 C age<br />

c<strong>al</strong>ibration program. Radiocarbon 35(1):215–230.<br />

37. Reimer PJ, <strong>et</strong> <strong>al</strong>. (2013) IntC<strong>al</strong>13 and MARINE13 radiocarbon age c<strong>al</strong>ibration curves 0–50000<br />

years c<strong>al</strong> BP. Radiocarbon 55(4):1869–1887.<br />

38. Ferziger JH, Perić M (1999) Computation<strong>al</strong> M<strong>et</strong>hods for Fluid Dynamics. (Springer-Verlag,<br />

Berlin, Germany), 2nd edition.<br />

39. Patankar SV (1980) Numeric<strong>al</strong> Heat Transfer and Fluid Flow. (Taylor & Francis, Boca Raton,<br />

FL, USA).<br />

40. Moerman JW, <strong>et</strong> <strong>al</strong>. (2013) Diurn<strong>al</strong> to interannu<strong>al</strong> rainf<strong>al</strong>l ” 18 O variations in northern Borneo<br />

driven by region<strong>al</strong> hydrology. Earth and Plan<strong>et</strong>ary Science L<strong>et</strong>ters 369–370:108–119.<br />

41. Partin JW, <strong>Cobb</strong> KM, Adkins JF, Clark B, Fernandez DP (2007) Millenni<strong>al</strong>-sc<strong>al</strong>e trends in west<br />

Pacific warm pool hydrology since the Last Glaci<strong>al</strong> Maximum. Nature 449:25–30.<br />

42. Moore JC, Grinsted A, Zwinger T, Jevrejeva S (2013) Semiempiric<strong>al</strong> and process-based<br />

glob<strong>al</strong> sea level projections. Reviews of Geophysics 51:484–522.<br />

43. Roundtable on Sustainable P<strong>al</strong>m Oil (2013) Principles and criteria for the production of sustainable<br />

p<strong>al</strong>m oil (http://www.rspo.org/file/PnC_RSPO_Rev1.pdf).<br />

1179<br />

1180<br />

1181<br />

1182<br />

1183<br />

1184<br />

1185<br />

1186<br />

1187<br />

1188<br />

1189<br />

1190<br />

1191<br />

1192<br />

1193<br />

1194<br />

1195<br />

1196<br />

1197<br />

1198<br />

1199<br />

1200<br />

1201<br />

1202<br />

1203<br />

1204<br />

1205<br />

1206<br />

1207<br />

1208<br />

1209<br />

1210<br />

1211<br />

1212<br />

1213<br />

1214<br />

1215<br />

1216<br />

1217<br />

1218<br />

1219<br />

1220<br />

1221<br />

1222<br />

1223<br />

1224<br />

1225<br />

1226<br />

1227<br />

1228<br />

1229<br />

1230<br />

1231<br />

1232<br />

1233<br />

1234<br />

1235<br />

1236<br />

1237<br />

1238<br />

1239<br />

1240<br />

10 | www.pnas.org/cgi/doi/10.1073/pnas.XXXXXXXXXX<br />

<strong>Cobb</strong> <strong>et</strong> <strong>al</strong>.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

SI Text<br />

S1. Extended Materi<strong>al</strong>s and M<strong>et</strong>hods<br />

Field data collection. Field data were collected in the Ulu<br />

Mendaram Conservation Area of the Belait District in Brunei<br />

Daruss<strong>al</strong>am (4.359863 N, 114.352252 E; Fig. 2). The site<br />

is described in more d<strong>et</strong>ail in Dommain <strong>et</strong> <strong>al</strong> (2015) (27).<br />

Average rainf<strong>al</strong>l at the nearby Seria and Ku<strong>al</strong>a Belait weather<br />

stations is 2880 mm / y (1947 through 2004). The area has<br />

never been logged in recorded history, <strong>al</strong>though commonly<br />

domesticated tree species planted <strong>al</strong>ong the river suggest that<br />

there were s<strong>et</strong>tlements there some decades ago. The peat dome<br />

shows a typic<strong>al</strong> north-west Borneo peat swamp catena (1),<br />

with mixed swamp forest at the river’s edge, followed by <strong>al</strong>an<br />

batu community (scattered, t<strong>al</strong>l Shorea <strong>al</strong>bida in upper canopy,<br />

large gaps mostly dominated by Pandanus andersonii), and<br />

then <strong>al</strong>an bunga community furthest from the river (closed<br />

upper canopy of Shorea <strong>al</strong>bida 50 m t<strong>al</strong>l). The forest floor is<br />

a dense tangle of buttresses, Pandanus rhizomes and woody<br />

debris, with no clearly defined channels for the flow of water<br />

(Fig. 3). Data are archived at the Dryad Digit<strong>al</strong> Repository,<br />

http://datadryad.org/.<br />

Soil respiration. Soil respiration measurements were made with<br />

an automated dynamic chamber system (LI-8100 with LI-<br />

8150 multiplexer, LI-COR Biosciences, Lincoln, NE, USA),<br />

in which carbon dioxide concentrations were measured after<br />

periodic closure of four chambers connected to an infrared gas<br />

an<strong>al</strong>yzer via a gas multiplexer. Each chambers was closed for<br />

6 minutes and a flux computed every hour using instrument<br />

software (LI-8100 software version 2.0.0). Chambers were<br />

supported 30 cm above the peat by four two-inch PVC pipes,<br />

1 m long, to avoid submergence of the chambers. Although<br />

temperature a ected individu<strong>al</strong> flux measurements around<br />

the diurn<strong>al</strong> cycle, the e ect of temperature on daily mean<br />

fluxes was negligible. Water table height was simultaneously<br />

monitored using a logging pressure transducer as described<br />

below (Piezom<strong>et</strong>ers).<br />

Microtopography transect. We used a tot<strong>al</strong> station (TC-GTS-<br />

105N (60536) 5" Tot<strong>al</strong>station, B.L. Makepeace, Inc.) to survey<br />

peat microtopography and 12 piezom<strong>et</strong>ers <strong>al</strong>ong a 180 m transect<br />

at the field site in August 2012 (Fig. 2d,e). We refer<br />

to the piezom<strong>et</strong>ers in this transect as “transect piezom<strong>et</strong>ers”<br />

to distinguish them from the 5 “trail piezom<strong>et</strong>ers” <strong>al</strong>ong the<br />

2.5 km trail. We located the transect survey points in geographic<br />

space by rotating the points to match the compass<br />

orientation b<strong>et</strong>ween the first two piezom<strong>et</strong>ers, then shifting<br />

the rotated points so that the position of the first piezom<strong>et</strong>er<br />

matched its GPS coordinates (Fig. 2e).<br />

To obtain the elevation of the survey points above mean<br />

sea level, we matched the survey data to the LiDAR data.<br />

The LiDAR DEM was defined by finding loc<strong>al</strong> minima in<br />

last-r<strong>et</strong>urn elevations on a 20 m x 20 m grid. Therefore, we<br />

found loc<strong>al</strong> minima in surveyed peat surface points on the<br />

same grid, and then <strong>al</strong>igned the (arbitrary) vertic<strong>al</strong> coordinate<br />

of the minim<strong>al</strong> survey points and the DEM elevation in the<br />

corresponding grid cell (Fig. 3b).<br />

We then defined a reference land surface p in the transect<br />

as follows. First, we fit a line to survey points by orthogon<strong>al</strong><br />

distance regression, then projected <strong>al</strong>l points on to that line.<br />

To find the water table elevation in each piezom<strong>et</strong>er, we subtracted<br />

the water table depth (Solinst Model 101 Water Level<br />

M<strong>et</strong>er, Solinst Canada, Georg<strong>et</strong>own, Ontario, Canada) from<br />

the elevation of the piezom<strong>et</strong>er casing. We then found the<br />

c<strong>al</strong>ibrated water table elevation vs. time in the piezom<strong>et</strong>er by<br />

matching the measured water table elevation to the piezom<strong>et</strong>er<br />

output at the time of measurement (Fig. 3b,c). We defined<br />

the land surface p in the microtopography transect as a linear<br />

interpolant of the average water table in each piezom<strong>et</strong>er,<br />

shifted up to <strong>al</strong>ign with the loc<strong>al</strong> minima in the surveyed peat<br />

surface points (Fig. 3b). By this definition, the land surface p<br />

is smooth because its shape comes from the water table, and<br />

it passes through the bottom of hollows because it is shifted<br />

to the loc<strong>al</strong> minima in the peat surface.<br />

We next used the transect hydrologic and microtopographic<br />

data to define the land surface at each trail piezom<strong>et</strong>er. Our<br />

definition of the land surface p as a smooth surface fit through<br />

the bottom of hollows makes it easy to estimate the loc<strong>al</strong> land<br />

surface elevation by looking at a nearby hollow. However,<br />

because this reference land surface is smoothed, we cannot<br />

easily find the exact height of trail piezom<strong>et</strong>ers relative to<br />

the land surface. The trail piezom<strong>et</strong>er nearest the transect<br />

piezom<strong>et</strong>ers (80 m to 130 m away) had <strong>al</strong>most identic<strong>al</strong> water<br />

table behavior to the transect piezom<strong>et</strong>ers during the time<br />

interv<strong>al</strong> of the transect piezom<strong>et</strong>er data (Fig. 3c,5a). Therefore<br />

to find the land surface elevation for each trail piezom<strong>et</strong>er,<br />

we first found the land surface for this nearest piezom<strong>et</strong>er by<br />

matching the data on the overlapping time interv<strong>al</strong>. Because<br />

the behavior of the water table was so similar in <strong>al</strong>l trail<br />

piezom<strong>et</strong>ers except the bog plain piezom<strong>et</strong>er, for plotting in<br />

Fig. 5 we simply matched the water table height ’ = H ≠ p<br />

for these piezom<strong>et</strong>ers to that of the piezom<strong>et</strong>er nearest the<br />

transect by subtracting the di erence in mean. Although this<br />

may introduce some error for the piezom<strong>et</strong>er outside the area<br />

of uniform water table behavior (the “bog plain piezom<strong>et</strong>er”),<br />

we did not use that piezom<strong>et</strong>er to d<strong>et</strong>ermine the specific yield<br />

and transmissivity functions for simulations.<br />

LiDAR data and topography. LiDAR data were obtained from the<br />

Brunei Survey Department from a 2009 aeri<strong>al</strong> mission. We<br />

constructed a digit<strong>al</strong> terrain model (DTM) from the LiDAR<br />

data by taking the minimum last-r<strong>et</strong>urn elevation within 20 m<br />

◊ 20 m grid cells, and then smoothing the resulting surface,<br />

corresponding to our definition of a reference land surface p as<br />

a smooth surface fit through hollows (loc<strong>al</strong> minima in ˜p). We<br />

then created a contour vector map from the DTM (GRASS<br />

GIS 6.4.3, http://grass.osgeo.org/) and outlined a flow tube<br />

containing our piezom<strong>et</strong>er and cores by constructing boundary<br />

flow lines from contours using the TAPES-C <strong>al</strong>gorithm (35),<br />

tracing back to an estimated location of the groundwater<br />

divide on the M<strong>al</strong>aysian side of the border (Fig. 2c). Although<br />

the groundwater divide is on the M<strong>al</strong>aysian side of the border<br />

where LiDAR data are unavailable, the tapered shape of the<br />

flowtube ensures that the simulations are insensitive to the<br />

precise location of the divide (Fig. 6).<br />

We sampled the magnitude of the gradient <strong>al</strong>ong<br />

each contour using a spati<strong>al</strong> database (SpatiaLite 4.2.0,<br />

http://www.gaia-gis.it/gaia-sins/). By the Divergence Theorem,<br />

the average Laplacian of surface elevation within an<br />

enclosed region is equ<strong>al</strong> to the norm<strong>al</strong> surface gradient integrated<br />

around its boundary divided by the enclosed area. We<br />

estimated the Laplacian in the region around the piezom<strong>et</strong>er<br />

DRAFT<br />

63<br />

64<br />

65<br />

66<br />

67<br />

68<br />

69<br />

70<br />

71<br />

72<br />

73<br />

74<br />

75<br />

76<br />

77<br />

78<br />

79<br />

80<br />

81<br />

82<br />

83<br />

84<br />

85<br />

86<br />

87<br />

88<br />

89<br />

90<br />

91<br />

92<br />

93<br />

94<br />

95<br />

96<br />

97<br />

98<br />

99<br />

100<br />

101<br />

102<br />

103<br />

104<br />

105<br />

106<br />

107<br />

108<br />

109<br />

110<br />

111<br />

112<br />

113<br />

114<br />

115<br />

116<br />

117<br />

118<br />

119<br />

120<br />

121<br />

122<br />

123<br />

124<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 1


125<br />

126<br />

127<br />

128<br />

129<br />

130<br />

131<br />

132<br />

133<br />

134<br />

135<br />

136<br />

137<br />

138<br />

139<br />

140<br />

141<br />

142<br />

143<br />

144<br />

145<br />

146<br />

147<br />

148<br />

149<br />

150<br />

151<br />

152<br />

153<br />

154<br />

155<br />

156<br />

157<br />

158<br />

159<br />

160<br />

161<br />

162<br />

163<br />

164<br />

165<br />

166<br />

167<br />

168<br />

169<br />

170<br />

171<br />

172<br />

173<br />

174<br />

175<br />

176<br />

177<br />

178<br />

179<br />

180<br />

181<br />

182<br />

183<br />

184<br />

185<br />

186<br />

by linear regression of the integrated norm<strong>al</strong> gradient against<br />

the enclosed area in the four piezom<strong>et</strong>ers nearest the river,<br />

where the slope was approximately uniform (Fig. 6).<br />

Peat cores. M<strong>et</strong>hods for the two peat cores and 15 samples taken<br />

for radiocarbon dating with a modified Livingstone corer are<br />

available in Dommain <strong>et</strong> <strong>al</strong> (2015) (27), and for the seven peat<br />

cores and 22 radiocarbon samples taken with a Russian peat<br />

sampler in Gandois <strong>et</strong> <strong>al</strong> (2014) (31). Median c<strong>al</strong>endar ages<br />

were estimated from radiocarbon ages using C<strong>al</strong>ib 7.1.0 with<br />

the IntC<strong>al</strong>13 c<strong>al</strong>ibration curve (36, 37). Because cores were<br />

taken in hollows, the top of each core corresponds roughly to<br />

the reference land surface p, which is fit through the bottoms<br />

of hollows.<br />

Piezom<strong>et</strong>ers. Water table height was recorded every 20 minutes<br />

at each of 5 piezom<strong>et</strong>ers <strong>al</strong>ong a 2.5 km transect using logging<br />

pressure transducers (Solinst Levelogger Edge 3001, Solinst<br />

Canada, Georg<strong>et</strong>own, Ontario, Canada; Fig. 2c). Transducer<br />

measurements were corrected for barom<strong>et</strong>ric fluctuations using<br />

a barom<strong>et</strong>er (Solinst Barologger Gold 3001). Each transducer<br />

was suspended on a steel cable inside a 2” PVC pipe 1.5 m<br />

in length inst<strong>al</strong>led to 1.4 m depth and screened at 1.3–1.45 m<br />

below the top of the piezom<strong>et</strong>er casing. Piezom<strong>et</strong>er locations<br />

were d<strong>et</strong>ermined with a GPS (GPSmap 60csx, Garmin Ltd.,<br />

Olathe, Kansas, USA).<br />

An addition<strong>al</strong> 12 piezom<strong>et</strong>ers were inst<strong>al</strong>led <strong>al</strong>ong an 180<br />

m transect 1 km from the river (Fig. 2d,e; Microtopography<br />

transect). These piezom<strong>et</strong>ers were screened from 1 m below the<br />

peat to 5 cm above the peat and were attached to 2” PVC pipes<br />

anchored in the underlying clay. Data from these piezom<strong>et</strong>ers<br />

were recorded from 2012-08-17 through 2012-11-03.<br />

Throughf<strong>al</strong>l measurement. Three siphoning tipping-buck<strong>et</strong> rain<br />

gauges (Texas Electronics TR-525S, D<strong>al</strong>las, Texas, USA) were<br />

inst<strong>al</strong>led <strong>al</strong>ong the microtopography transect to record throughf<strong>al</strong>l<br />

in the forest understory. Each gauge was mounted on an<br />

acrylic plate supported 50 cm above the peat surface on a 2”<br />

PVC tube pushed into the peat. Large fronds of understory<br />

plants (mostly Pandanus andersonii) were cleared from above<br />

gauges to reduce the spati<strong>al</strong> variability in measurements. Tip<br />

events were recorded by a switch closure event logger (UA-<br />

003-64, Ons<strong>et</strong> Computer Corp, Bourne, MA, USA) enclosed<br />

in a sm<strong>al</strong>l PVC case attached to each gauge.<br />

C<strong>al</strong>culation of stable peat dome topography. We found<br />

the stable peat surface Laplacian for a given rainf<strong>al</strong>l<br />

regime by a one-dimension<strong>al</strong> search as follows. The<br />

water table in a stable peatland can be simulated efficiently<br />

using an integrating ODE solver (SUNDIALS,<br />

http://computation.llnl.gov/projects/sundi<strong>al</strong>s). These simulations<br />

are fast because the water table behaves the same<br />

everywhere, so there is only one variable to simulate (Eqn. 5).<br />

Initi<strong>al</strong>ly, we do not know the Laplacian of the stable peat<br />

surface Ò 2 p Œ for the ODE. However, by taking a guess at the<br />

surface Laplacian and simulating the water table using the<br />

ODE, we can d<strong>et</strong>ermine wh<strong>et</strong>her we have the correct Laplacian<br />

v<strong>al</strong>ue. If our guess of the surface Laplacian is too large in<br />

magnitude, the average water table height will be too low,<br />

and decomposition will exceed peat production; if our guess<br />

of the surface Laplacian is too sm<strong>al</strong>l, peat production will<br />

exceed decomposition. Thus, we find the stable peat surface<br />

Laplacian by simulating the water table (Eqn. 5) and checking<br />

for b<strong>al</strong>ance b<strong>et</strong>ween peat production and decomposition over<br />

time (Eqn. 3), successively refining our guess until we have the<br />

stable surface Laplacian for the given rainf<strong>al</strong>l regime (Brent’s<br />

m<strong>et</strong>hod).<br />

Dynamic model of peat dome development. We solved loc<strong>al</strong><br />

peat accumulation and groundwater flow equations numeric<strong>al</strong>ly<br />

to simulate development of peat domes. We simulated the<br />

dynamics of groundwater flow (Eqn. 1) and the dynamics<br />

of peat accumulation (Eqn. 2) using a finite volume scheme<br />

in one horizont<strong>al</strong> dimension representing a flow tube on the<br />

peat surface (Fig. S2, S3, Expanded description of peat dome<br />

simulation). Simulating tropic<strong>al</strong> peat swamp hydrology is<br />

di cult because both transmissivity and specific yield depend<br />

strongly on the water table elevation relative to the surface.<br />

We avoided the numeric<strong>al</strong> issues that arise from the severe<br />

nonlinearity of the flow problem by expressing Boussinesq’s<br />

equation in conservation form (38) via a change of variables<br />

from the water table elevation to the loc<strong>al</strong> water storage<br />

(volume per area). The switch to conservation form made it<br />

feasible to simulate storm responses over thousands of years<br />

with time steps of minutes.<br />

We first split the water table elevation H into the surface<br />

elevation p and the water table elevation relative to the surface<br />

’ = H ≠ p. Evolution of the peat surface is extremely slow<br />

relative to movement of the water table, so we treated the<br />

peat surface as quasi-steady while simulating changes in the<br />

water table, and re-wrote Boussinesq’s equation (Eqn. 1) in<br />

terms of the water elevation above the surface ’<br />

ˆ’<br />

S y = qn + Ò · (T Ò’)+Ò · (T Òp) . [8]<br />

ˆt<br />

A di erenti<strong>al</strong> change in water storage W at a point in the<br />

landscape is related to a change in water table via the specific<br />

yield, dW = S y dH, so we can express the water table elevation<br />

relative to the surface in terms of a water storage above the<br />

land surface, or “sh<strong>al</strong>low storage” W s<br />

’ =<br />

⁄ Ws<br />

DRAFT<br />

W Õ =0<br />

S ≠1<br />

y (W Õ )dW Õ [9]<br />

where W Õ is a dummy variable of integration. Because the<br />

specific yield function S y(W ) is not a function of time, we can<br />

then re-write the groundwater flow equation in conservation<br />

form<br />

ˆW s<br />

= q n + Ò · (DÒW s)+Ò · (T Òp) [10]<br />

ˆt<br />

where the aquifer di usivity D is defined convention<strong>al</strong>ly as<br />

transmissivity divided by specific yield D = T/S y. We then<br />

discr<strong>et</strong>ized time and space following standard m<strong>et</strong>hods and<br />

solved the discr<strong>et</strong>ized form of the water conservation equation<br />

(Eqn. 10) using a Crank-Nicolson scheme with a no-flow boundary<br />

at the groundwater divide and specified surface and water<br />

table elevation dynamics at the other boundary (S3, Expanded<br />

description of peat dome simulation). For each hydrologic time<br />

step, multiple intern<strong>al</strong> steps might be taken for stability and<br />

accuracy based on a standard stability heuristic (39). After<br />

each hydrologic time step, the surface was incremented by<br />

straightforward use of equation 1, and the change in water<br />

storage below the surface was c<strong>al</strong>culated and deducted from<br />

the sh<strong>al</strong>low storage W s for conservation of water volume.<br />

Model c<strong>al</strong>ibration.<br />

187<br />

188<br />

189<br />

190<br />

191<br />

192<br />

193<br />

194<br />

195<br />

196<br />

197<br />

198<br />

199<br />

200<br />

201<br />

202<br />

203<br />

204<br />

205<br />

206<br />

207<br />

208<br />

209<br />

210<br />

211<br />

212<br />

213<br />

214<br />

215<br />

216<br />

217<br />

218<br />

219<br />

220<br />

221<br />

222<br />

223<br />

224<br />

225<br />

226<br />

227<br />

228<br />

229<br />

230<br />

231<br />

232<br />

233<br />

234<br />

235<br />

236<br />

237<br />

238<br />

239<br />

240<br />

241<br />

242<br />

243<br />

244<br />

245<br />

246<br />

247<br />

248<br />

2 |


249<br />

250<br />

251<br />

252<br />

253<br />

254<br />

255<br />

256<br />

257<br />

258<br />

259<br />

260<br />

261<br />

262<br />

263<br />

264<br />

265<br />

266<br />

267<br />

268<br />

269<br />

270<br />

271<br />

272<br />

273<br />

274<br />

275<br />

276<br />

277<br />

278<br />

279<br />

280<br />

281<br />

282<br />

283<br />

284<br />

285<br />

286<br />

287<br />

288<br />

289<br />

290<br />

291<br />

292<br />

293<br />

294<br />

295<br />

296<br />

297<br />

298<br />

299<br />

300<br />

301<br />

302<br />

303<br />

304<br />

305<br />

306<br />

307<br />

308<br />

309<br />

310<br />

Master recharge and recession curves. The master recharge and<br />

recession curves were computed based on the uniform water<br />

table behavior among the four piezom<strong>et</strong>ers nearest the river.<br />

A related approach, without the assembly of master curves<br />

but using the division of water table time series into periods<br />

of heavy rain, days without precipitation, and nights without<br />

precipitation, has been used previously in other studies<br />

of w<strong>et</strong>land hydrology (28, 29). Under very heavy rain the<br />

groundwater divergence term Ò · (T ÒH) in the groundwater<br />

flow equation (Eqn. 1) becomes negligible<br />

S y<br />

ˆH<br />

ˆt<br />

¥ qr, [11]<br />

and because specific yield S y is a function of water table<br />

relative to the surface ’ only, an integr<strong>al</strong> form of the simplified<br />

di erenti<strong>al</strong> equation<br />

⁄<br />

P =<br />

S y d’ + constant [12]<br />

corresponds to the master recharge curve, giving the cumulative<br />

rainf<strong>al</strong>l depth P driving a change in water table height.<br />

During dry interv<strong>al</strong>s, “n<strong>et</strong> precipitation” consists entirely of<br />

evapotranspiration, and taking evapotranspiration as constant<br />

and equ<strong>al</strong> to its average through the diurn<strong>al</strong> cycle q <strong>et</strong> (nonnegative)<br />

yields a simplified di erenti<strong>al</strong> equation without rain<br />

S y<br />

d’<br />

dt = ≠q<strong>et</strong> + T Ò2 p [13]<br />

which, in its integr<strong>al</strong> form,<br />

⁄<br />

S y(’)<br />

t =<br />

d’ + constant [14]<br />

≠q <strong>et</strong> + T Ò 2 p<br />

is equiv<strong>al</strong>ent to the master recession curve (Fig. 5f) turned on<br />

its side. The Laplacian Ò 2 p was obtained from topographic<br />

data as described above (“LiDAR data and topography”).<br />

We assembled the master recharge and recession curves<br />

by using throughf<strong>al</strong>l data (above) to identify interv<strong>al</strong>s of the<br />

time series either with no rain, or with heavy rain (rainf<strong>al</strong>l<br />

intensity greater than 4 mm / h). Within each interv<strong>al</strong>, we resampled<br />

a spline of the head versus time data to yield time on<br />

a uniform grid of water table height. We then found an o s<strong>et</strong><br />

in time or rain depth for each interv<strong>al</strong> to minimize the leastsquares<br />

di erence b<strong>et</strong>ween overlapping interv<strong>al</strong>s, creating the<br />

master recharge and recession curves (S4, Master recession and<br />

recharge curve assembly). At low water tables, the water table<br />

was nearly constant at night and decreased linearly during the<br />

day (Fig. 5f), so we estimated evapotranspiration as 2.246 mm<br />

/ d using the slope of the daytime decrease in water table and<br />

specific yield obtained from the master recharge curve.<br />

We estimated param<strong>et</strong>ric functions for the specific yield<br />

(cubic spline) and transmissivity (piecewise-linear logarithm of<br />

hydraulic conductivity) by simultaneously fitting the recharge<br />

and recession curves to model results (Levenberg-Marquardt,<br />

http://www.pesthomepage.org). We simulated recharge (water<br />

table height vs. cumulative rainf<strong>al</strong>l depth ’(P )) or recession<br />

(water table height vs. time since rain ’(t)) bysolving<br />

the relevant di erenti<strong>al</strong> equation (Eqn. 12 for recharge,<br />

Eqn. 14 for recession) using an integrating solver (SUNDI-<br />

ALS, http://computation.llnl.gov/projects/sundi<strong>al</strong>s) with the<br />

current param<strong>et</strong>er estimates. Although the data provide no<br />

information on how conductivity is distributed in the peat<br />

below the lowest excursion of the water table, the dynamics<br />

of the water table are <strong>al</strong>ways dominated by other factors,<br />

either the conductivity of peat higher in the soil profile or<br />

evapotranspiration.<br />

C<strong>al</strong>ibration for peat surface evolution param<strong>et</strong>ers. After estimation<br />

of hydrologic param<strong>et</strong>ers T,S y, we estimated the peat accumulation<br />

param<strong>et</strong>ers f p, – by fitting the simulated shape<br />

of our peat dome to the LiDAR surface using a nonlinear<br />

least-squares optimization routine (Levenberg-Marquardt,<br />

http://www.pesthomepage.org). The coast<strong>al</strong> peat swamp<br />

forests of Brunei Daruss<strong>al</strong>am are underlain by a broad, very<br />

gently sloped mangrove clay (1, 27), so as the initi<strong>al</strong> condition<br />

we used a flat, sloped surface corresponding to a gradient in<br />

the elevation of bas<strong>al</strong> peat in our peat cores (Fig. 7a).<br />

We drove our simulation of the peat dome using a precipitation<br />

time series derived from our own throughf<strong>al</strong>l data,<br />

m<strong>et</strong>eorologic<strong>al</strong> data, and nearby speleothem ” 18 O records using<br />

an approach similar to that of Kurnianto <strong>et</strong> <strong>al</strong>. (16). Our<br />

go<strong>al</strong> was not to reconstruct the actu<strong>al</strong> rainf<strong>al</strong>l at our site<br />

over the last 3000 years, but to reasonably approximate the<br />

fluctuation in rainf<strong>al</strong>l at sub-annu<strong>al</strong>, decad<strong>al</strong>, and millenni<strong>al</strong><br />

time sc<strong>al</strong>es. For sub-annu<strong>al</strong> rainf<strong>al</strong>l, we cycled through field<br />

throughf<strong>al</strong>l intensity data from 2012 on a 20-minute grid from<br />

our site. We then sc<strong>al</strong>ed throughf<strong>al</strong>l intensities by a factor<br />

specific to each simulation year and month as follows. First,<br />

we followed the approach of Moerman <strong>et</strong> <strong>al</strong>. (40) and c<strong>al</strong>culated<br />

a regression of two-month rainf<strong>al</strong>l means at a nearby<br />

m<strong>et</strong>eorologic<strong>al</strong> station (Brunei Internation<strong>al</strong> Airport (BWN),<br />

90 km north-east of our site) against ” 18 O at Gunung Mulu airport,<br />

61 km south-east of our site, from August 2006 through<br />

April 2011 (P =2.53006 mm / d ≠ ” 18 O ◊ 1.00825 mm / d,<br />

R 2 =0.208513). We then used this regression to compute<br />

mean rainf<strong>al</strong>l intensities for each decad<strong>al</strong> to centenni<strong>al</strong> interv<strong>al</strong><br />

captured by the speleothems at Mulu, which are believed to<br />

characterize region<strong>al</strong> patterns of rainf<strong>al</strong>l (40, 41). We then<br />

repeated 44 years of monthly precipitation tot<strong>al</strong>s from the<br />

m<strong>et</strong>eorologic<strong>al</strong> station (1966–2009, mean annu<strong>al</strong> rainf<strong>al</strong>l 2.90<br />

m) for 3000 years, adjusting monthly means to match the linearly<br />

interpolated precipitation averages from the speleothem<br />

data regression. Fin<strong>al</strong>ly, we sc<strong>al</strong>ed 20-minute intensities in<br />

each month of the 3000 year time series to match the month<br />

precipitation tot<strong>al</strong> d<strong>et</strong>ermined by the m<strong>et</strong>eorologic<strong>al</strong> data<br />

DRAFT<br />

and speleothem record, sc<strong>al</strong>ed by a constant for <strong>al</strong>l months<br />

and years so that precipitation in the measured year (2012)<br />

matched throughf<strong>al</strong>l at our site. Evapotranspiration was modeled<br />

as zero at night and constant by day, equ<strong>al</strong> to e ective<br />

evapotranspiration estimated from piezom<strong>et</strong>er data as described<br />

above. Because the d<strong>et</strong>ails of changes in river stage<br />

over the last 3000 years are unknown, we used dates in the<br />

core closest to the river to drive the gradu<strong>al</strong>ly shifting surface<br />

elevation and assumed that average river stage remained in<br />

the same position relative to the surface at the boundary.<br />

Sensitivity to environment<strong>al</strong> and anthropogenic change. We<br />

explored the e ects of fluctuations in rainf<strong>al</strong>l by using a very<br />

simple model for rainf<strong>al</strong>l (Poisson process, exponenti<strong>al</strong> depth<br />

distribution with two param<strong>et</strong>ers: average storm depth and<br />

average inter-storm arriv<strong>al</strong> time. Because rain storms are<br />

instantaneous in this simple model, we computed the increase<br />

in head from each storm by numeric<strong>al</strong> integration of S y(’),<br />

then found the recession of head using an ODE solver as when<br />

311<br />

312<br />

313<br />

314<br />

315<br />

316<br />

317<br />

318<br />

319<br />

320<br />

321<br />

322<br />

323<br />

324<br />

325<br />

326<br />

327<br />

328<br />

329<br />

330<br />

331<br />

332<br />

333<br />

334<br />

335<br />

336<br />

337<br />

338<br />

339<br />

340<br />

341<br />

342<br />

343<br />

344<br />

345<br />

346<br />

347<br />

348<br />

349<br />

350<br />

351<br />

352<br />

353<br />

354<br />

355<br />

356<br />

357<br />

358<br />

359<br />

360<br />

361<br />

362<br />

363<br />

364<br />

365<br />

366<br />

367<br />

368<br />

369<br />

370<br />

371<br />

372<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 3


373<br />

374<br />

375<br />

376<br />

377<br />

378<br />

379<br />

380<br />

381<br />

382<br />

383<br />

384<br />

385<br />

386<br />

387<br />

388<br />

389<br />

390<br />

391<br />

392<br />

393<br />

394<br />

395<br />

396<br />

397<br />

398<br />

399<br />

400<br />

401<br />

402<br />

403<br />

404<br />

405<br />

406<br />

407<br />

408<br />

409<br />

410<br />

411<br />

412<br />

413<br />

414<br />

415<br />

416<br />

417<br />

418<br />

419<br />

420<br />

421<br />

422<br />

423<br />

424<br />

425<br />

426<br />

427<br />

428<br />

429<br />

430<br />

431<br />

432<br />

433<br />

434<br />

c<strong>al</strong>ibrating to the master recession curves (Model c<strong>al</strong>ibration,<br />

above). We then found the stable Laplacian for each n<strong>et</strong><br />

precipitation time series by root-finding as described above<br />

(C<strong>al</strong>culation of stable peat topography). As a point of reference<br />

for intermittency of rainf<strong>al</strong>l, we c<strong>al</strong>culated the average<br />

inter-storm arriv<strong>al</strong> time at our site as the average duration of<br />

contiguous sequences of 20-minute interv<strong>al</strong>s without rain.<br />

In a second s<strong>et</strong> of simulations, we changed the amplitudes<br />

of annu<strong>al</strong> and El Niño Southern Oscillation (ENSO) fluctuations<br />

in simulated rainf<strong>al</strong>l time series. We first created a<br />

reference rainf<strong>al</strong>l time series by superimposing the 44-year<br />

variation in monthly mean rainf<strong>al</strong>l from m<strong>et</strong>eorologic<strong>al</strong> data<br />

on our throughf<strong>al</strong>l rainf<strong>al</strong>l intensities (as in C<strong>al</strong>ibration for<br />

peat surface evolution param<strong>et</strong>ers, above, but without superimposing<br />

the speleothem sign<strong>al</strong>). We define the amplitudes<br />

of the annu<strong>al</strong> and ENSO fluctuations as the amplitudes of<br />

sinusoids with arbitrary phases fit to daily rainf<strong>al</strong>l intensities<br />

by least-squares. For each desired combination of annu<strong>al</strong> and<br />

ENSO amplitudes, we then found the rain time series with<br />

the same mean rainf<strong>al</strong>l and desired amplitudes at annu<strong>al</strong> and<br />

ENSO periods that was least-squares closest to this reference<br />

rainf<strong>al</strong>l time series, while requiring that rainf<strong>al</strong>l intensity was<br />

positive or zero and the pattern of storms was the same, using<br />

a quadratic programming solver (S6, S<strong>et</strong>ting annu<strong>al</strong> and<br />

ENSO amplitudes of rainf<strong>al</strong>l).<br />

We simulated the e ects of environment<strong>al</strong> and anthropogenic<br />

change in the future by imposing di erent forcing<br />

and boundary conditions on simulations starting with an ide<strong>al</strong>ized<br />

peat ridge b<strong>et</strong>ween two par<strong>al</strong>lel rivers. We generated<br />

the ridge with the simulated precipitation time series used<br />

for model c<strong>al</strong>ibration, and chose the breadth of the ridge so<br />

that it would be approximately stable at the end of the time<br />

series (2308 years). The base case for the projections was<br />

the reference rainf<strong>al</strong>l time series described above. Current climate<br />

projections indicate increased average rainf<strong>al</strong>l, increased<br />

season<strong>al</strong>ity of rainf<strong>al</strong>l, and probably both in the tropics (33).<br />

Therefore, we performed simulations with a stronger season<strong>al</strong><br />

sign<strong>al</strong>, and sc<strong>al</strong>ed the rainf<strong>al</strong>l time series to explore the effects<br />

of higher average rainf<strong>al</strong>l. In the first perturbation, we<br />

increased the season<strong>al</strong> fluctuations to 3.0 mm / d to match<br />

projections for average fluctuations at Brunei’s latitude (4 N)<br />

during 2081–2100 (33) (increase of 0.52866 mm / d). In a<br />

second projection, we increased mean rainf<strong>al</strong>l by the same<br />

amount. The tid<strong>al</strong> rivers draining coast<strong>al</strong> tropic<strong>al</strong> peatlands<br />

will be directly a ected by sea level changes, so in a third<br />

perturbation, we increased the boundary water level by 50 cm<br />

over an interv<strong>al</strong> of 100 y to mimic the e ect of sea level rise<br />

(42). (Isostatic subsidence or uplift, such as the postglaci<strong>al</strong><br />

rebound important in many northern peatlands (11), could<br />

be accommodated by the model in the same way.) Fin<strong>al</strong>ly, we<br />

simulated the e ects on dome shape of drainage to a constant<br />

50 cm below the peat surface, the recommended best practice<br />

of the Roundtable for Sustainable P<strong>al</strong>m Oil (43), to model the<br />

e ects of conversion to agriculture on dome shape and carbon<br />

e ux.<br />

S2. Hydrologic budg<strong>et</strong> for our site<br />

We can use water table and throughf<strong>al</strong>l measurements to c<strong>al</strong>culate<br />

an estimated water budg<strong>et</strong> for our site (28). Evapotranspiration<br />

and divergence of groundwater flow can be distinguished<br />

in the decline of the water table b<strong>et</strong>ween rain events because<br />

at low water tables, when divergence of groundwater flow is<br />

sm<strong>al</strong>l, evapotranspiration creates a distinct diurn<strong>al</strong> pattern of<br />

steady water tables at night and declining water tables during<br />

the day (Fig. 5d). Therefore, after estimating specific yield<br />

from the ascending curve (Fig. 5e), we can estimate evapotranspiration<br />

from the daytime decline in the water table when<br />

the water table is low (Fig. 5f). Based on this estimate of<br />

evapotranspiration and our throughf<strong>al</strong>l gauge data, there was<br />

307 cm of throughf<strong>al</strong>l to the peat at our site in 2012, of which<br />

82 cm was lost as evapotranspiration from the peat. Over<strong>al</strong>l<br />

precipitation f<strong>al</strong>ling on the peat forest, and evapotranspiration<br />

from the forest, are higher because some rain is intercepted by<br />

the canopy and evaporates before it reaches the peat surface.<br />

The increase in stored water in the peatland over the year,<br />

or recharge, was about 4 mm: the water table was 1.3 cm<br />

higher at the end of the data interv<strong>al</strong> (2013-01-20) than at the<br />

beginning (2012-02-06), and the specific yield in that water<br />

table height range (5.2 cm to 6.5 cm) averages 0.3 (Fig. 5).<br />

Because the recharge term was only 4 mm, nearly <strong>al</strong>l of the n<strong>et</strong><br />

precipitation (225 cm) was lost via divergence of groundwater<br />

flow.<br />

S3. Expanded description of peat dome simulation<br />

Implementation of numeric<strong>al</strong> model. The water table in a tropic<strong>al</strong><br />

peatland can rise by centim<strong>et</strong>ers in a matter of hours in<br />

response to intense convective rainstorms, and fluctuations<br />

on these short time-sc<strong>al</strong>es a ect the changes in the landscape<br />

over millennia (Fig. S1d,e). Therefore simulations required an<br />

extremely large number of steps. Two practic<strong>al</strong> consequences<br />

of this are that simulations are too large to fit in physic<strong>al</strong> memory<br />

on a current mid-range hardware, and there is a risk of<br />

numeric<strong>al</strong> losses in computing cumulants to verify water conservation<br />

by the model. To overcome these two issues, we stored<br />

forcing data and simulations in Hierarchic<strong>al</strong> Data Format<br />

(http://www.hdfgroup.org/HDF5), periodic<strong>al</strong>ly pref<strong>et</strong>ching<br />

data from disk to construct interpolants of n<strong>et</strong> precipitation<br />

intensity, and computed cumulants using exact arithm<strong>et</strong>ic (H<strong>et</strong>tinger’s<br />

lsum, http://code.activestate.com/recipes/393090/).<br />

The over<strong>al</strong>l simulation driver was implemented in Python<br />

(version 2.7.5; http://www.python.org); rate-limiting c<strong>al</strong>culations<br />

were implemented in Cython (version 0.21.1;<br />

http://cython.org).<br />

DRAFT<br />

Flow line coordinate system. Simulations of groundwater flow<br />

and peat accumulation used a flow line coordinate system.<br />

Flow lines, contours, and elevation on the peat dome form<br />

an approximately orthogon<strong>al</strong> coordinate system: streamlines<br />

and contours are <strong>al</strong>ways orthogon<strong>al</strong>, and elevation is nearly<br />

orthogon<strong>al</strong> to the plane of the surface because gradients are<br />

very sm<strong>al</strong>l. In gener<strong>al</strong>, in orthogon<strong>al</strong> curvilinear coordinates<br />

with three coordinates (r 1,r 2,r 3) and unit vectors e 1, e 2, e 3,<br />

there is <strong>al</strong>so an arc length param<strong>et</strong>er associated with each<br />

coordinate, h 1,h 2,h 3. In Cartesian coordinates, <strong>al</strong>l three arc<br />

length param<strong>et</strong>ers are identic<strong>al</strong>ly 1, but in gener<strong>al</strong> they can<br />

be functions of the coordinates. The gradient operator in<br />

orthogon<strong>al</strong> curvilinear coordinates is<br />

ÒV = 1 h 1<br />

ˆV<br />

ˆr 1<br />

e 1 + 1 h 2<br />

ˆV<br />

ˆr 2<br />

e 2 + 1 h 3<br />

ˆV<br />

ˆr 3<br />

e 3 [15]<br />

435<br />

436<br />

437<br />

438<br />

439<br />

440<br />

441<br />

442<br />

443<br />

444<br />

445<br />

446<br />

447<br />

448<br />

449<br />

450<br />

451<br />

452<br />

453<br />

454<br />

455<br />

456<br />

457<br />

458<br />

459<br />

460<br />

461<br />

462<br />

463<br />

464<br />

465<br />

466<br />

467<br />

468<br />

469<br />

470<br />

471<br />

472<br />

473<br />

474<br />

475<br />

476<br />

477<br />

478<br />

479<br />

480<br />

481<br />

482<br />

483<br />

484<br />

485<br />

486<br />

487<br />

488<br />

489<br />

490<br />

491<br />

492<br />

493<br />

494<br />

495<br />

496<br />

4 |


497<br />

498<br />

499<br />

500<br />

501<br />

502<br />

503<br />

504<br />

505<br />

506<br />

507<br />

508<br />

509<br />

510<br />

511<br />

512<br />

513<br />

514<br />

515<br />

516<br />

517<br />

518<br />

519<br />

520<br />

521<br />

522<br />

523<br />

524<br />

525<br />

526<br />

527<br />

528<br />

529<br />

530<br />

531<br />

532<br />

533<br />

534<br />

535<br />

536<br />

537<br />

538<br />

539<br />

540<br />

541<br />

542<br />

543<br />

544<br />

545<br />

546<br />

547<br />

548<br />

549<br />

550<br />

551<br />

552<br />

553<br />

554<br />

555<br />

556<br />

557<br />

558<br />

and the divergence operator is<br />

Ò · F =<br />

Ë<br />

1 ˆ<br />

(h 2h 3F 1)<br />

h 1h 2h 3 ˆr 1<br />

+ ˆ (h 3h 1F 2)<br />

ˆr 2<br />

+ ˆ È<br />

(h 1h 2F 3)<br />

ˆr 3<br />

[16]<br />

We worked in a 2-dimension<strong>al</strong> coordinate system of streamlines<br />

and contours, consistent with the use of the essenti<strong>al</strong>ly<br />

horizont<strong>al</strong> flow approximation for hydrologic c<strong>al</strong>culations, and<br />

considered elevation z approximately constant. The gradient<br />

operator is defined with reference to a streamline arc length<br />

param<strong>et</strong>er h s relating a di erenti<strong>al</strong> change in that coordinate<br />

to geom<strong>et</strong>ric distance<br />

ÒH = 1 ˆH<br />

es. [17]<br />

h s ˆs<br />

The divergence operator depends on both the streamline arc<br />

length param<strong>et</strong>er and an an<strong>al</strong>ogous contour arc length param<strong>et</strong>er<br />

h „ (s) describing the distance associated with a di erenti<strong>al</strong><br />

change in a coordinate „ <strong>al</strong>ong a contour,<br />

Ò · T ÒH = 1<br />

h sh „<br />

ˆ<br />

ˆs<br />

1<br />

h„<br />

T ˆH<br />

h s ˆs<br />

2<br />

. [18]<br />

In the speci<strong>al</strong> cases of Cartesian coordinates the arc lengths<br />

param<strong>et</strong>ers are both identic<strong>al</strong>ly 1, whereas for cylindric<strong>al</strong> polar<br />

coordinates the contour arc length param<strong>et</strong>er is equ<strong>al</strong> to the<br />

radi<strong>al</strong> position h „ = r.<br />

Along the reference streamline, the streamline arc length<br />

param<strong>et</strong>er is 1 by definition, and we approximated the contour<br />

arc length param<strong>et</strong>er h „ as the ratio of the arc length ¸(s)<br />

<strong>al</strong>ong that contour to an adjacent streamline relative to the<br />

distance ¸(s ú) to that neighboring streamline at the boundary<br />

h „ (s) = ¸(s)<br />

¸(s ú) . [19]<br />

Then in flow tube coordinates, the groundwater flow equation,<br />

rewritten in terms of sh<strong>al</strong>low storage W s (Eqn. 9) and surface<br />

elevation p, becomes<br />

ˆW s<br />

ˆt<br />

= q n + 1¸<br />

1<br />

ˆ<br />

ˆs<br />

¸T ˆH<br />

ˆs<br />

2<br />

+ 1¸<br />

1<br />

ˆ<br />

¸D ˆp 2<br />

. [20]<br />

ˆs ˆs<br />

Discr<strong>et</strong>ization of groundwater flow equations. We discr<strong>et</strong>ized the<br />

groundwater flow problem in flow tube coordinates (Eqn. 20)<br />

using a standard one-dimension<strong>al</strong> finite-volume scheme (39),<br />

with a Neumann boundary condition at the origin (zero gradient<br />

at groundwater divide) and a Dirichl<strong>et</strong> condition at the<br />

drainage boundary. We use n as an index for time, written<br />

as a superscript, and j as an index for space, written as a<br />

subscript. By integrating through the cell j and dividing by<br />

its area A j, we discr<strong>et</strong>ize the volume-conservation equation<br />

(Eqn. 20)<br />

ˆW s<br />

ˆt<br />

= q n + 1 Ë È<br />

¸ D ˆWs s2j<br />

+ 1 Ë<br />

¸T ˆp È s2j<br />

. [21]<br />

A j ˆs s 2j≠1<br />

A j ˆs s 2j≠1<br />

E ectively, we have decomposed the flux in and out of the<br />

cell into a component associated with the gradient in the loc<strong>al</strong><br />

water table elevation relative to the surface, and another component<br />

for the gradient in the surface itself. As a mnemonic,<br />

we c<strong>al</strong>l these components the sh<strong>al</strong>low flux Q s and the deep<br />

flux Q o, <strong>al</strong>though this has nothing to do with where in the<br />

soil profile the water transport occurs.<br />

We approximate the gradients with finite di erences, and<br />

find fluxes through the cell faces. Geom<strong>et</strong>ry is handled in<br />

much the same way in both cases. The sh<strong>al</strong>low fluxes are<br />

approximated as<br />

Q s;1,j ¥ (Ê 2,j≠1D j≠1 + Ê 1,jD j)(W s;j ≠ W s;j≠1) [22]<br />

Q s;2,j ¥ (Ê 2,jD j + Ê 1,jD j+1)(W s;j+1 ≠ W s;j) [23]<br />

and the deep fluxes as<br />

Q o;1,j ¥ (Ê 2,j≠1T j≠1 + Ê 1,jT j)(p j ≠ p j≠1) [24]<br />

Q o;2,j ¥ (Ê 2,jT j + Ê 1,jT j+1)(p j+1 ≠ p j) [25]<br />

where Ê is a dimensionless aspect ratio, accounting for both<br />

the ratio of cell flow line segment length to the length of the<br />

face b<strong>et</strong>ween cells ¸ and the position of the face <strong>al</strong>ong the<br />

distance b<strong>et</strong>ween nodes<br />

Ê 1,j =<br />

DRAFT<br />

¸2,j≠1(s1,j ≠ s2,j≠1)<br />

(s 1,j ≠ s 1,j≠1) 2 [26]<br />

and<br />

¸2,j(s2,j ≠ s1,j)<br />

Ê 2,j =<br />

(s 1,j+1 ≠ s . [27]<br />

1,j) 2<br />

This approach ensures that the downstream flow Q 2,j out of<br />

acellj equ<strong>al</strong>s the upstream flow Q 1,j+1 into its downstream<br />

neighbor j +1for volume conservation.<br />

After integrating with respect to time, the discr<strong>et</strong>ized problem<br />

for interior points j œ {2, 3,...,J ≠ 1} has equation<br />

W n+1<br />

s;j ≠ W n s;j = P n + a s;2,j (W s;j+1 ≠ W s;j)<br />

≠ a s;1,j (W s;j ≠ W s;j≠1)<br />

+ a o;2,j (p j+1 ≠ p j)<br />

≠ a o;2,j (p j ≠ p j≠1)<br />

in which coe cients a s = – s t for sh<strong>al</strong>low fluxes<br />

and a o = – o<br />

[28]<br />

– s;1,j = — 1,jD j≠1 + — 2,jD j<br />

– s;2,j = — 3,jD j + — 4,jD j+1<br />

[29]<br />

t for deep fluxes<br />

– o;1,j = — 1,jT j≠1 + — 2,jT j<br />

– o;2,j = — 3,jT j + — 4,jT j+1<br />

[30]<br />

depend on computed v<strong>al</strong>ues of transmissivity and specific yield<br />

and on a geom<strong>et</strong>ric param<strong>et</strong>er — that summarizes <strong>al</strong>l spati<strong>al</strong><br />

information in the problem<br />

— 1j = Ê 2,j≠1 A ≠1<br />

j<br />

— 2j = Ê 1,j A ≠1<br />

j<br />

— 3j = Ê 2,j A ≠1<br />

j<br />

— 4j = Ê 1,j+1 A ≠1<br />

j<br />

.<br />

[31]<br />

The Neumann condition (zero flux) at the origin and the<br />

Dirichl<strong>et</strong> condition (specified sh<strong>al</strong>low storage and surface elevation)<br />

at the right boundary are implemented in the standard<br />

way (39), so that the matrices describing the sh<strong>al</strong>low flux A n s<br />

are constructed as<br />

S<br />

W<br />

U<br />

– s n 2,1 ≠– s n 2,1<br />

≠– s n 1,2 – s n 1,2 + – s n 2,2 ≠– s n 2,2<br />

. .. . .. . ..<br />

X<br />

≠– s n 1,J≠1 – s n 1,J≠1 + – s n 2,J≠1 ≠– s n V<br />

2,J≠1<br />

0<br />

[32]<br />

T<br />

559<br />

560<br />

561<br />

562<br />

563<br />

564<br />

565<br />

566<br />

567<br />

568<br />

569<br />

570<br />

571<br />

572<br />

573<br />

574<br />

575<br />

576<br />

577<br />

578<br />

579<br />

580<br />

581<br />

582<br />

583<br />

584<br />

585<br />

586<br />

587<br />

588<br />

589<br />

590<br />

591<br />

592<br />

593<br />

594<br />

595<br />

596<br />

597<br />

598<br />

599<br />

600<br />

601<br />

602<br />

603<br />

604<br />

605<br />

606<br />

607<br />

608<br />

609<br />

610<br />

611<br />

612<br />

613<br />

614<br />

615<br />

616<br />

617<br />

618<br />

619<br />

620<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 5


621<br />

622<br />

623<br />

624<br />

625<br />

626<br />

627<br />

628<br />

629<br />

630<br />

631<br />

632<br />

633<br />

634<br />

635<br />

636<br />

637<br />

638<br />

639<br />

640<br />

641<br />

642<br />

643<br />

644<br />

645<br />

646<br />

647<br />

648<br />

649<br />

650<br />

651<br />

652<br />

653<br />

654<br />

655<br />

656<br />

657<br />

658<br />

659<br />

660<br />

661<br />

662<br />

663<br />

664<br />

665<br />

666<br />

667<br />

668<br />

669<br />

670<br />

671<br />

672<br />

673<br />

674<br />

675<br />

676<br />

677<br />

678<br />

679<br />

680<br />

681<br />

682<br />

and an<strong>al</strong>ogously the matrices describing the deep flux A n o are<br />

S<br />

T<br />

– o n 2,1 ≠– o n 2,1<br />

≠– o n 1,2 – o n 1,2 + – o n 2,2 ≠– o n 2,2<br />

.<br />

W<br />

.. . .. . ..<br />

X<br />

U<br />

≠– o n 1,J≠1 – o n 1,J≠1 + – o n 2,J≠1 ≠– o n V<br />

2,J≠1<br />

0<br />

[33]<br />

Avector÷ represents changes in storage with the time step<br />

forced either by n<strong>et</strong> surface-atmosphere flux q n or by boundary<br />

conditions<br />

;<br />

÷ n+ P n , j œ {1, 2,...J ≠ 1}<br />

j =<br />

W n+1<br />

J<br />

≠ WJ n [34]<br />

, j = J.<br />

where<br />

Then the system of equations to be solved is<br />

B n s W n+1<br />

s + B n o p n = C n s W n s + C n o p n + ÷ n [35]<br />

B n s = I + f n tA n+1<br />

s<br />

B n o = I + f n tA n+1<br />

o<br />

C n s = I +(1≠ f n ) tA n+1<br />

s<br />

C n s = I +(1≠ f n ) tA n+1<br />

o<br />

[36]<br />

and f represents a weighting for explicit and implicit steps;<br />

a Crank-Nicolson scheme (f = 1/2) was used for <strong>al</strong>l c<strong>al</strong>culations.<br />

Solution of the system was first attempted using<br />

Picard iteration; if that approach failed, the system<br />

was solved using Powell’s m<strong>et</strong>hod (GNU Scientific Library,<br />

https://www.gnu.org/software/gsl). If Powell’s m<strong>et</strong>hod <strong>al</strong>so<br />

failed, the step size was h<strong>al</strong>ved and Picard’s m<strong>et</strong>hod tried<br />

again. The convergence criterion for both Picard and Powell<br />

solvers was reduction of the L 2 norm of the residu<strong>al</strong> error by<br />

5 orders of magnitude (38).<br />

We chose the time step conservatively for stability by requiring<br />

that the diagon<strong>al</strong>s of the explicit coe cient matrix for<br />

sh<strong>al</strong>low fluxes C s and for deep fluxes C o be kept positive (39),<br />

that is,<br />

! #<br />

t Æ min min (–s 1,j + – s 2,j) ≠1 ,<br />

jœ2,3,...,J≠1<br />

(– o 1,j + – ≠1$" [37]<br />

o 2,j) .<br />

Strictly speaking, this is a heuristic for the stability of explicit<br />

steps, and can be exceeded when using the Crank-Nicolson<br />

approach adopted here. However, because the nonlinear solvers<br />

som<strong>et</strong>imes failed to converge with large step sizes because of<br />

the very strong nonlinearity of the problem, in practice it was<br />

most e cient to choose time steps exceeding this heuristic<br />

maximum by no more than a sm<strong>al</strong>l factor (e.g., 2).<br />

S4. Master recession and recharge curve assembly<br />

We assembled the master recession curve and master recharge<br />

curve by least-squares <strong>al</strong>ignment of short water table time<br />

series representing interv<strong>al</strong>s of either intense rain or no rain.<br />

We describe first the procedure for the master recession curve,<br />

which is slightly simpler. The head versus time sequence is<br />

split into J inter-storm series sorted by initi<strong>al</strong> head so that<br />

the lowest series index has the lowest initi<strong>al</strong> head. For each<br />

series, the head versus time data were interpolated with a<br />

cubic spline and resampled to give time versus head data with<br />

head v<strong>al</strong>ues integer multiples of a uniform step size H. In<br />

gener<strong>al</strong>, a series may hit the same head at multiple times<br />

because of measurement noise and near-constant heads low in<br />

the peat at night. The equations are nearly the same, except<br />

for weighting, if those times are replaced by their mean, so<br />

that each series j has no more than one time t ij for each head<br />

i: for each head v<strong>al</strong>ue in each short series, <strong>al</strong>l distinct times<br />

t at a particular head were replaced with their average ¯t to<br />

give distinct v<strong>al</strong>ues (k, j, ¯t) where k is the head as an integer<br />

multiple of the head step size and j is the series index.<br />

S<strong>et</strong>s of time-versus-head series can be assembled into a<br />

recession curve if one can navigate from the lowest to the<br />

highest head via regions of overlap b<strong>et</strong>ween series. In most<br />

cases, there were some short sequences at the very highest and<br />

very lowest heads where a subs<strong>et</strong> of series did not overlap with<br />

the rest. These sm<strong>al</strong>l subs<strong>et</strong>s of series were eliminated, leaving<br />

a single large connected component of time-versus head series.<br />

We <strong>al</strong>so removed heads at which there was only one series<br />

because these contribute no information when finding time<br />

o s<strong>et</strong>s.<br />

We then solved for a time o s<strong>et</strong> t j for each series by<br />

minimizing the mean squared di erence in <strong>al</strong>l times at the<br />

same heads, as follows. Ide<strong>al</strong>ly, after the o s<strong>et</strong> for a series t j<br />

has been applied, the time t ij of that series equ<strong>al</strong>s the mean<br />

time at that head of <strong>al</strong>l J(i) series with a v<strong>al</strong>ue at that head i:<br />

A<br />

t ij + t j = 1<br />

J(i)<br />

B<br />

ÿ<br />

t j + t ij [38]<br />

J(i)<br />

j=1<br />

so at each discr<strong>et</strong>e water level i, for each series j we have a<br />

single equation with an unknown time o s<strong>et</strong> t j<br />

A<br />

B<br />

J(i)<br />

1 ÿ<br />

t j ≠ t j = t ij ≠ 1<br />

J(i)<br />

ÿ<br />

t ij [39]<br />

J(i)<br />

J(i)<br />

#<br />

1<br />

J(i)<br />

j=1<br />

1<br />

...<br />

J(i)<br />

1<br />

≠ 1 ... 1<br />

J(i) J(i)<br />

DRAFT<br />

$<br />

W<br />

U<br />

j=1<br />

which can be re-written in matrix form with typic<strong>al</strong> row<br />

S<br />

t 1<br />

T<br />

t 2 X<br />

=<br />

C<br />

t ij ≠ 1<br />

J(i)<br />

.<br />

t j<br />

.<br />

t J≠1<br />

A ÿJ(i)<br />

BD<br />

t ij .<br />

j=1<br />

X<br />

V [40]<br />

The prior elimination of disconnected series ensures that there<br />

is an equation at each head, and the remov<strong>al</strong> of heads with only<br />

one series ensures that there are no trivi<strong>al</strong> equations. Note<br />

that we have excluded a reference series J from the unknown<br />

vector; its o s<strong>et</strong> is fixed to 0 to make the problem non-singular.<br />

After least-squares solution of this system of equations (LU<br />

decomposition), the solution vector contains the time o s<strong>et</strong>s<br />

t for the corresponding series.<br />

We constructed master recharge curves using an approach<br />

similar to that used for master recession curves, based on<br />

the idea that when rain is intense, discharge is negligible by<br />

comparison<br />

dH<br />

S y ¥ qr [41]<br />

dt<br />

with q r the rainf<strong>al</strong>l. We c<strong>al</strong>culated the cumulative tot<strong>al</strong> rain<br />

depth as a function of time for each interv<strong>al</strong> of heavy rain, then<br />

683<br />

684<br />

685<br />

686<br />

687<br />

688<br />

689<br />

690<br />

691<br />

692<br />

693<br />

694<br />

695<br />

696<br />

697<br />

698<br />

699<br />

700<br />

701<br />

702<br />

703<br />

704<br />

705<br />

706<br />

707<br />

708<br />

709<br />

710<br />

711<br />

712<br />

713<br />

714<br />

715<br />

716<br />

717<br />

718<br />

719<br />

720<br />

721<br />

722<br />

723<br />

724<br />

725<br />

726<br />

727<br />

728<br />

729<br />

730<br />

731<br />

732<br />

733<br />

734<br />

735<br />

736<br />

737<br />

738<br />

739<br />

740<br />

741<br />

742<br />

743<br />

744<br />

6 |


745<br />

746<br />

747<br />

748<br />

749<br />

750<br />

751<br />

752<br />

753<br />

754<br />

755<br />

756<br />

757<br />

758<br />

759<br />

760<br />

761<br />

762<br />

763<br />

764<br />

765<br />

766<br />

767<br />

768<br />

769<br />

770<br />

771<br />

772<br />

773<br />

774<br />

775<br />

776<br />

777<br />

778<br />

779<br />

780<br />

781<br />

782<br />

783<br />

784<br />

785<br />

786<br />

787<br />

788<br />

789<br />

790<br />

791<br />

792<br />

793<br />

794<br />

795<br />

796<br />

797<br />

798<br />

799<br />

800<br />

801<br />

802<br />

803<br />

804<br />

805<br />

806<br />

found the rain depth o s<strong>et</strong> for each series using the master<br />

recession curve least-squares approach. The resulting master<br />

recharge curve represents head as a function of cumulative<br />

rain depth, and is equiv<strong>al</strong>ent to the integr<strong>al</strong> of specific yield<br />

s<br />

Sy dH across <strong>al</strong>l heads H.<br />

Rain storms did not arrive simultaneously at throughf<strong>al</strong>l<br />

gauges and <strong>al</strong>l piezom<strong>et</strong>ers, so it was necessary to match<br />

storms as recorded at throughf<strong>al</strong>l gauges to the time of the<br />

storm’s arriv<strong>al</strong> at each piezom<strong>et</strong>er. We treated each contiguous<br />

interv<strong>al</strong> of rapidly increasing head as a storm, then searched<br />

the throughf<strong>al</strong>l time series for the corresponding record of<br />

that storm, considering that it might be slightly o s<strong>et</strong> in time.<br />

Usu<strong>al</strong>ly there was only one heavy rain that overlapped with the<br />

increasing head interv<strong>al</strong>; in rare cases, there were two. In those<br />

cases we chose the storm with the higher tot<strong>al</strong> depth, assuming<br />

that it would be more likely to cause the rapid increase in<br />

head. If the time o s<strong>et</strong> for ons<strong>et</strong> of the storm di ered greatly<br />

(by more than 1 h), the candidate storm was discarded.<br />

Evapotranspiration estimated from the master recession<br />

curve represents loss from the connected porewater and surface<br />

water, not the tot<strong>al</strong> evapotranspiration from the forest. Tot<strong>al</strong><br />

evapotranspiration <strong>al</strong>so includes water intercepted by live<br />

foliage and water perched on the abundant leaf litter suspended<br />

above the peat surface, and is probably considerably higher<br />

(28).<br />

S5. Effects of rainf<strong>al</strong>l aggregation time<br />

We explored the importance of short-term fluctuation in rainf<strong>al</strong>l<br />

on long-term simulations of peat accumulation by experimenting<br />

with di erent time grids. In our simulations,<br />

we represented rainf<strong>al</strong>l intensity and evapotranspiration as<br />

piecewise-constant on 20-minute interv<strong>al</strong>s. Most models represent<br />

rainf<strong>al</strong>l on a coarser time grid, making rainf<strong>al</strong>l constant<br />

on time sc<strong>al</strong>es of months, years or centuries, but we chose<br />

a much finer grid to represent rainf<strong>al</strong>l intensity because our<br />

results showed that short-term fluctuations in rainf<strong>al</strong>l can have<br />

a significant e ect on the simulated long-term evolution of peat<br />

domes. We started with our 20-minute time grid, on which<br />

rainf<strong>al</strong>l is modeled as piecewise-constant on each consecutive<br />

time interv<strong>al</strong>:<br />

R(t) =R i,tœ [t i,t i+1), iœ {1, 2,...,n≠ 1}. [42]<br />

We then computed new representations of rainf<strong>al</strong>l intensity<br />

and evapotranspiration such that the rainf<strong>al</strong>l intensity on<br />

a piecewise-constant interv<strong>al</strong> in the output is equ<strong>al</strong> to the<br />

average intensity on the same interv<strong>al</strong> in the input:<br />

Rj Õ 1<br />

=<br />

t Õ j+1 ≠ tÕ j<br />

⁄ t<br />

Õ<br />

j+1<br />

t=t Õ j<br />

R(t)dt, j œ {1, 2,...n Õ ≠ 1} [43]<br />

where the over<strong>al</strong>l time span is the same (t Õ 1 = t 1,t Õ nÕ = tn) but<br />

there are fewer tot<strong>al</strong> time steps (n Õ


869<br />

870<br />

871<br />

872<br />

873<br />

874<br />

875<br />

876<br />

877<br />

878<br />

879<br />

880<br />

881<br />

882<br />

883<br />

884<br />

885<br />

886<br />

887<br />

888<br />

889<br />

890<br />

891<br />

892<br />

893<br />

894<br />

895<br />

896<br />

897<br />

898<br />

899<br />

900<br />

901<br />

902<br />

903<br />

904<br />

905<br />

906<br />

907<br />

908<br />

909<br />

910<br />

911<br />

912<br />

913<br />

914<br />

915<br />

916<br />

917<br />

918<br />

919<br />

920<br />

921<br />

922<br />

923<br />

924<br />

925<br />

926<br />

927<br />

928<br />

929<br />

930<br />

To ensure that the new rainf<strong>al</strong>l intensity vector ˆr has the same<br />

mean µ and the specified harmonic content a, b we therefore<br />

require<br />

S T<br />

C +ˆr = U â<br />

ˆb V [49]<br />

ˆ“<br />

where C + is the Moore-Penrose pseudoinverse of C. Subtracting<br />

the equation defining the harmonic content of the origin<strong>al</strong><br />

rainf<strong>al</strong>l intensity vector C + r = [a, b, “] T , the constraint is<br />

equiv<strong>al</strong>ently expressed in terms of the di erence b<strong>et</strong>ween the<br />

new and old rainf<strong>al</strong>l time series x = ˆr ≠ r<br />

C +ˆx =<br />

S<br />

U â ≠ a<br />

ˆb ≠ b<br />

ˆ“ ≠ “<br />

T<br />

V . [50]<br />

Now that we have written the constraints as linear equ<strong>al</strong>ity<br />

constraints and box constraints, the over<strong>al</strong>l problem of finding<br />

the new rainf<strong>al</strong>l intensity vector ˆr can be written as the<br />

quadratic program<br />

minimize<br />

subject to<br />

1<br />

2 xT x<br />

Gx ∞ h<br />

Ax = b<br />

[51]<br />

where h is an m-vector consisting of the non-zero entries of r,<br />

G is a m ◊ n matrix defined by<br />

;<br />

≠1, ri ”= 0,i= j<br />

g ij =<br />

[52]<br />

0, otherwise<br />

and A is a block matrix enforcing the linear equ<strong>al</strong>ity constraints:<br />

that zero entries remain zero, that the mean is the<br />

same, and that amplitudes at the chosen frequencies are as<br />

specified in the new time series:<br />

5 6<br />

F<br />

C +<br />

and<br />

A =<br />

S<br />

W<br />

b = U<br />

0<br />

â ≠ a<br />

ˆb ≠ b<br />

ˆ“ ≠ “<br />

T<br />

[53]<br />

X<br />

V [54]<br />

with F an (n ≠ m) ◊ n matrix that keeps the zero entries equ<strong>al</strong><br />

to zero<br />

;<br />

1, ri =0,i= j<br />

f ij =<br />

. [55]<br />

0, otherwise<br />

The quadratic program (Eqn. 51) was then solved using a<br />

sparse QP solver (CVXOPT, http://cvxopt.org).<br />

DRAFT<br />

931<br />

932<br />

933<br />

934<br />

935<br />

936<br />

937<br />

938<br />

939<br />

940<br />

941<br />

942<br />

943<br />

944<br />

945<br />

946<br />

947<br />

948<br />

949<br />

950<br />

951<br />

952<br />

953<br />

954<br />

955<br />

956<br />

957<br />

958<br />

959<br />

960<br />

961<br />

962<br />

963<br />

964<br />

965<br />

966<br />

967<br />

968<br />

969<br />

970<br />

971<br />

972<br />

973<br />

974<br />

975<br />

976<br />

977<br />

978<br />

979<br />

980<br />

981<br />

982<br />

983<br />

984<br />

985<br />

986<br />

987<br />

988<br />

989<br />

990<br />

991<br />

992<br />

8 |


993<br />

994<br />

995<br />

996<br />

997<br />

998<br />

999<br />

1000<br />

1001<br />

1002<br />

1003<br />

1004<br />

1005<br />

1006<br />

1007<br />

1008<br />

1009<br />

1010<br />

1011<br />

1012<br />

1013<br />

1014<br />

1015<br />

1016<br />

1017<br />

1018<br />

1019<br />

1020<br />

1021<br />

1022<br />

1023<br />

1024<br />

1025<br />

1026<br />

1027<br />

1028<br />

1029<br />

1030<br />

1031<br />

1032<br />

1033<br />

1034<br />

1035<br />

1036<br />

1037<br />

1038<br />

1039<br />

1040<br />

1041<br />

1042<br />

1043<br />

1044<br />

1045<br />

1046<br />

1047<br />

1048<br />

1049<br />

1050<br />

1051<br />

1052<br />

1053<br />

1054<br />

SI Figure Legends<br />

Figure S1 | Rainf<strong>al</strong>l intensity data used to drive simulations.<br />

a, Annu<strong>al</strong> mean rainf<strong>al</strong>l intensity from Brunei Airport (BWN)<br />

m<strong>et</strong>eorologic<strong>al</strong> station after correction to match throughf<strong>al</strong>l<br />

(black) and with stronger ENSO sign<strong>al</strong> added (gray). b,<br />

Monthly mean rainf<strong>al</strong>l intensity from BWN, resc<strong>al</strong>ed to match<br />

throughf<strong>al</strong>l at site as used in base simulation (black) and<br />

with stronger season<strong>al</strong> sign<strong>al</strong> (gray), showing medians, upper<br />

and lower quartiles, outliers (+) and means. c, Rainf<strong>al</strong>l<br />

intensity used for simulations of past peat dome evolution.<br />

Oxygen isotope data from speleothem data at Mulu Park,<br />

Sarawak (Partin <strong>et</strong> <strong>al</strong> 2007, Moerman <strong>et</strong> <strong>al</strong> 2013) was used to<br />

approximate fluctuations in past rainf<strong>al</strong>l (SI Text, Sensitivity<br />

to environment<strong>al</strong> and anthropogenic change). d,e, Bias in<br />

stable surface Laplacian as a function of rainf<strong>al</strong>l averaging<br />

time. Using a coarser grid for rainf<strong>al</strong>l time series results in<br />

overestimation of the stable Laplacian of the surface because<br />

re<strong>al</strong> fluctuations in the water table driven by intermittent rain<br />

increase the average transmissivity, reducing long-term peat<br />

accumulation.<br />

Figure S2 | Flow tube discr<strong>et</strong>ization used for peat growth and<br />

hydrologic c<strong>al</strong>culations. a, Discr<strong>et</strong>ization of flow tube for finite<br />

volume hydrologic simulation. b, Piecewise-constant and<br />

piecewise-linear approximations of variables for hydrologic<br />

simulation.<br />

Figure S3 | Simulated past and future morphology and fluxes<br />

of Mendaram peat dome. a, Simulated past and future development<br />

of Mendaram peat dome towards its stable shape with<br />

uniform Laplacian (dashed line). b, Current modeled CO 2<br />

sequestration rate vs. distance from groundwater divide at<br />

Mendaram peat dome. c, Modeled CO 2 sequestration rate<br />

vs. position and time at Mendaram peat dome. d, Spati<strong>al</strong><br />

average of modeled CO 2 sequestration rate of Mendaram peat<br />

dome vs. time.<br />

Figure S4 | Convergence to uniform surface Laplacian. a,<br />

Simulated development of a peat ridge growing on an initi<strong>al</strong>ly<br />

flat substrate under simulated precipitation (average n<strong>et</strong><br />

precipitation 1477 mm / y, solid lines) or under constant n<strong>et</strong><br />

precipitation that yields the same stable shape, only 190 mm<br />

/ y (dashed lines), and their convergence to a surface with a<br />

uniform Laplacian (dotted line). The factor-of-eight di erence<br />

in mean rainf<strong>al</strong>l required to generate the same peat dome<br />

morphology with steady (dashed) and intermittent (solid) precipitation<br />

illustrates the strong e ect of fluctuations in rainf<strong>al</strong>l<br />

on long-term peat accumulation. Colors from time color bar,<br />

bottom. b, Laplacian from (a); with simulated rainf<strong>al</strong>l (solid)<br />

and with constant n<strong>et</strong> precipitation (dashed), and convergence<br />

to ultimate Laplacian (horizont<strong>al</strong> dotted line). c, Surface with<br />

a uniform negative Laplacian (dashed) and with a positive<br />

Laplacian (solid). d,e, Simulated water table at three points<br />

on surfaces from (c): uniform negative Laplacian (d) and<br />

positive Laplacian (e). The water table behaves the same at<br />

<strong>al</strong>l three locations shown in (c) with a uniform negative land<br />

surface Laplacian (d), but behaves di erently at the three<br />

locations if the surface Laplacian is positive (e).<br />

DRAFT<br />

1055<br />

1056<br />

1057<br />

1058<br />

1059<br />

1060<br />

1061<br />

1062<br />

1063<br />

1064<br />

1065<br />

1066<br />

1067<br />

1068<br />

1069<br />

1070<br />

1071<br />

1072<br />

1073<br />

1074<br />

1075<br />

1076<br />

1077<br />

1078<br />

1079<br />

1080<br />

1081<br />

1082<br />

1083<br />

1084<br />

1085<br />

1086<br />

1087<br />

1088<br />

1089<br />

1090<br />

1091<br />

1092<br />

1093<br />

1094<br />

1095<br />

1096<br />

1097<br />

1098<br />

1099<br />

1100<br />

1101<br />

1102<br />

1103<br />

1104<br />

1105<br />

1106<br />

1107<br />

1108<br />

1109<br />

1110<br />

1111<br />

1112<br />

1113<br />

1114<br />

1115<br />

1116<br />

PNAS | April 13, 2017 | vol. XXX | no. XX | 9


a<br />

Rainf<strong>al</strong>l intensity<br />

R, mm / d<br />

c<br />

Rainf<strong>al</strong>l intensity<br />

R, mm / d<br />

Year<br />

b<br />

Rainf<strong>al</strong>l intensity R, mm / d<br />

d<br />

Year<br />

e


a<br />

b<br />

node<br />

face


a<br />

Land surface<br />

elevation p, m<br />

b<br />

Current CO 2 flux,<br />

t ha -1 y -1<br />

12<br />

8<br />

4<br />

0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Time, y<br />

-2000 0<br />

2000 4000 6000<br />

6 4 2<br />

0<br />

Distance from river, km<br />

Distance from<br />

river, km<br />

Average CO 2 flux,<br />

t ha -1 y -1<br />

CO 2 flux, t ha -1 y -1<br />

0 2<br />

4<br />

c<br />

6<br />

4<br />

2<br />

0<br />

d<br />

3.0<br />

1.5<br />

0<br />

0 2500 5000<br />

Time, y


Land surface elevation p, m<br />

a<br />

b<br />

Laplacian of land surface<br />

elevation , km -1 x 1000<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

0<br />

−5<br />

−10<br />

1000<br />

750<br />

500 250<br />

0<br />

Distance from river, m<br />

Land surface elevation p, m<br />

Water table height , cm<br />

c<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

1000 750<br />

500<br />

250<br />

0<br />

Distance from river, m<br />

20<br />

0<br />

−20<br />

20<br />

d<br />

e<br />

0<br />

−20<br />

0 30 60 90<br />

Elapsed time, d<br />

−3000<br />

−2000<br />

Time, y<br />

−1000<br />

0

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!