Earth and Planetary Science Letters 430 (2015) 149–159

Contents lists available at ScienceDirect

Earth and Planetary Science Letters

www.elsevier.com/locate/epsl

Can oxygen stable isotopes be used to track precipitation moisture source in vascular -dominated peatlands? ∗ Matthew J. Amesbury a, , Dan J. Charman a, Rewi M. Newnham b, Neil J. Loader c, Jordan Goodrich d,e, Jessica Royles f, David I. Campbell d, Elizabeth D. Keller g, W. Troy Baisden g, Thomas P. Roland a, Angela V. Gallego-Sala a a Geography, College of Life and Environmental Sciences, University of Exeter, Exeter, UK b School of Geography, Earth and Environmental Sciences, Victoria University of Wellington, Wellington, New Zealand c Department of Geography, College of Science, Swansea University, Swansea, UK d School of Science, University of Waikato, Hamilton, New Zealand e Global Change Research Group, San Diego State University, San Diego, USA f Department of Plant Sciences, University of Cambridge, Cambridge, UK g National Isotope Centre, GNS Science, Lower Hutt, Wellington, New Zealand a r t i c l e i n f o a b s t r a c t

Article history: Variations in the isotopic composition of precipitation are determined by fractionation processes Received 10 February 2015 which occur during temperature- and humidity-dependent phase changes associated with evaporation Received in revised form 25 June 2015 and condensation. Oxygen stable isotope ratios have therefore been frequently used as a source of Accepted 14 August 2015 palaeoclimate data from a variety of proxy archives, which integrate this signal over time. Applications Available online 28 August 2015 from ombrotrophic peatlands, where the source water used in cellulose synthesis is derived solely from Editor: D. Vance precipitation, have been mostly limited to Northern Hemisphere Sphagnum-dominated bogs, with few Keywords: in the Southern Hemisphere or in peatlands dominated by vascular . New Zealand (NZ) provides stable isotopes an ideal location to undertake empirical research into oxygen isotope fractionation in vascular peatlands oxygen because single taxon analysis can be easily carried out, in particular using the preserved root matrix of precipitation moisture source the restionaceous wire rush ( spp.) that forms deep Holocene peat deposits throughout the Empodisma country. Furthermore, large gradients are observed in the mean isotopic composition of precipitation restiad peatland across NZ, caused primarily by the relative influence of different climate modes. Here, we test whether New Zealand δ18O of Empodisma α-cellulose from ombrotrophic restiad peatlands in NZ can provide a methodology for developing palaeoclimate records of past precipitation δ18O. Surface plant, water and precipitation ◦ samples were taken over spatial (six sites spanning >10 latitude) and temporal (monthly measurements over one year) gradients. A link between the isotopic composition of root-associated water, the most likely source water for plant growth, and precipitation in both datasets was found. Back-trajectory modelling of precipitation moisture source for rain days prior to sampling showed clear seasonality in the temporal data that was reflected in root-associated water. The link between source water and plant cellulose was less clear, although mechanistic modelling predicted mean cellulose values within published error margins for both datasets. Improved physiological understanding and modelling of δ18O in restiad peatlands should enable use of this approach as a new source of palaeoclimate data to reconstruct changes in past atmospheric circulation. © 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

1. Introduction 1964; Craig and Gordon, 1965; Gat, 2000). The isotopic composi- tion of precipitation is also known to reflect rainout history and Variations in the stable isotopic composition of precipitation atmospheric circulation patterns along given air mass trajectories (δ18O and δD) are determined by multiple fractionation processes (Cole et al., 1999; Araguas-Araguas et al., 2000). The subsequent which occur during temperature- and humidity-dependent phase incorporation of this signal into plants, sediments and ice has led 18 changes associated with evaporation and condensation (Dansgaard, to the use of oxygen stable isotope ratios (δ O) for palaeoclimate reconstructions from a variety of proxy archives including tree * Corresponding author. Tel.: +44 0 1392 725892. rings, lakes, speleothems (see Leng, 2004 and references therein), E-mail address: [email protected] (M.J. Amesbury). ice cores (Brook, 2007)and, less frequently, in peatlands. Ex- http://dx.doi.org/10.1016/j.epsl.2015.08.015 0012-821X/© 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 150 M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 ploitation of the δ18O record from ombrotrophic peatlands, where the source water used in cellulose synthesis is derived solely from precipitation, has been mostly limited to Northern Hemi- sphere Sphagnum-dominated bogs (e.g. Brenninkmeijer et al., 1982; Ménot-Combes et al., 2002; Zanazzi and Mora, 2005; Daley et al., 2010). There has been limited application in the Southern Hemi- sphere (SH; Royles et al., 2013) or in -dominated peatlands (Hong et al., 2000, 2009), despite considerable potential for such records to improve understanding of SH climate dynamics. The oxygen stable isotope content of vascular plant cellulose 18 (δ OC) is determined by kinetic and equilibrium fractionation pathways from source water (Barbour, 2007). Leaf water is used for sucrose synthesis in plants and is generally isotopically en- 18 riched relative to meteoric source water (δ OSW) due to evap- orative enrichment during transpiration (Yakir and DeNiro, 1990; 18 Buhay et al., 1996). δ OC depends on both the isotopic composi- tion of the sucrose and that of the water at the site of cellulose synthesis, which may be spatially and temporally separated from sucrose synthesis. Models to describe these fractionation processes 18 18 and predict δ OC from δ OSW have been developed, initially for trees (e.g. Waterhouse et al., 2002), but have been widely ap- plied, including to peatlands dominated by different vegetation types (e.g. Ménot-Combes et al., 2002; Zanazzi and Mora, 2005; Daley et al., 2010). Vascular and non-vascular plants record differ- ent δ18O from a given water source (Ménot-Combes et al., 2002; Nichols et al., 2010), largely due to the presence/absence respec- tively of stomata capable of regulating moisture/gas exchange. 18 In addition, although δ OC records have previously been inter- preted as ‘palaeo-thermometers’ driven by surface air temperature (e.g. Hong et al., 2000), this link is indirect (Barbour, 2007), can be complicated by the uncoupling of air and leaf temperatures (Helliker and Richter, 2008) and is often untested (e.g. Oldfield, 2001). Quantitative palaeo-temperature estimates from peatland 18 18 δ OC data are therefore problematic and instead, δ OC has been interpreted as a proxy for precipitation moisture source (Daley Fig. 1. Site locations. Kauri line represents geographical split of the two Empodisma 18 ◦ ◦ et al., 2010). If palaeoclimate indices are developed from δ OC species E. robustum (to the north of 38 ) and E. minus (to the south of 38 ) (Wagstaff records in novel geographical or ecological contexts, it is criti- and Clarkson, 2012). Base map is isoscape modelled annually-weighted 1997–2013 mean δ18O in precipitation (h; see Section 2.2 for details) generated using the cal that relationships between the isotopic compositions of pre- relationships reported in Table S1. For a full colour version of this figure, the reader cipitation, source waters and plant cellulose are empirically and is referred to the web version of this article. mechanistically grounded and calibrated with modern climate data (e.g. Daley et al., 2012), or understood via probabilistic statisti- and form a dense canopy of standing living and dead shoots, re- cal approaches such as inverse proxy modelling (Yu et al., 2011; ducing rates of surface evaporation (Agnew et al., 1993; Campbell Yu, 2013). and Williamson, 1997; Wagstaff and Clarkson, 2012). NZ is there- The primary influences on the stable oxygen isotopic compo- fore an ideal location to develop understanding of oxygen isotope sition of plant cellulose are 1) the isotopic composition of source fractionation in vascular plant peatlands because the dominance of water and of water at the site of cellulose synthesis; 2) isotopic a single taxon in both surface vegetation and peat avoids mixed enrichment of water in the leaf due to evaporation (i.e. equilib- species effects (Ménot-Combes et al., 2002; Nichols et al., 2010). rium and kinetic physical fractionation effects); and 3) biochemi- Many of the peatlands also preserve high-resolution full-Holocene cal fractionations during cellulose synthesis. Whilst physical frac- sequences (e.g. Newnham et al., 1995; Vandergoes et al., 2005) tionation effects are temperature- and humidity-dependent, the and are situated in the climatically sensitive SH mid-latitudes (e.g. ± h biochemical fractionation of 27 3 (Yakir and DeNiro, 1990; Ummenhofer and England, 2007; Fletcher and Moreno, 2012). Cru- Aucour et al., 1996) has generally been considered to be indepen- cially, large gradients exist in the annual mean isotopic compo- dent of temperature, although some evidence exists to the con- sition of precipitation across NZ (Fig. 1), caused primarily by the 18 trary (Sternberg and Ellsworth, 2011). In studies of δ OC from relative influence of northern and southern air masses driven by Sphagnum-dominated peatlands, physical fractionation processes different climate modes (e.g. ENSO and the southern westerlies, have been disregarded due to the simpler physiological water-use driven by the Southern Annular Mode; Ummenhofer and England, strategies of mosses compared to vascular plants, such that only a 2007), as well as altitude and rainout effects across inland areas. biochemical fractionation factor has been applied to test the link Here, we test the potential for δ18O in ombrotrophic restiad 18 between moss cellulose and δ OSW (Daley et al., 2010). peatlands to provide a basis for developing palaeoclimatic records Throughout NZ, ombrotrophic peatlands are dominated by the that track precipitation moisture source. Specifically, we test the restionaceous wire rush ( and Empodisma robus- following hypotheses: tum; Wagstaff and Clarkson, 2012). Both species, occurring sepa- ◦ rately north and south of 38 S (known as the ‘Kauri Line’), have H1: The oxygen stable isotopic composition of the source water 18 reduced, scale-like leaves, high water-use efficiency, a surface clus- (δ OSW) used by Empodisma spp. in cellulose synthesis tracks the ter root matrix with a similar base exchange and water holding isotopic composition of precipitation, both spatially and tempo- capacity to Sphagnum, can reduce water loss via stomatal control rally. M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 151

) H2: The oxygen stable isotopic composition of the peat form- )

71 18 64 . . annual

ing parts of Empodisma spp. ( O ; i.e. the cluster root matrix; δ D 5 C 4 ( ( δ (SD)

mean Wagstaff and Clarkson, 2012) reflect spatial and temporal varia- are

28 39 . . ppt tion in the source water, subject to fractionation and enrichment 24 47 – – – – in annual GNIP (ppt) − − effects.

)

) ) ) ) )

2. Methods 23 . 17 94 81 95 05 . . . . . mean Din 8 6 7 7 8 10

( ( ( ( ( ( δ

2.1. Sites and field sampling precipitation

81 18 46 35 62 37 (SD) ......

30 50 30 44 23 43 Din annual ppt Isoscape δ − − − − − Six Empodisma-dominated ombrotrophic peatlands were se- − lected to span variability in the isotopic gradient of precipitation and ) )

O O 93 across NZ (Fig. 1, Table 1). Divided by the Kauri Line, the three 76 . . 18 18 0 0 deviation. δ ( ( δ

(SD) northern sites, Tangonge (TNG), Kopuatai (KOP) and Moanatuatua

mean

48 01 for . (MOA) contain E. robustum whereas the three southern sites, Kaipo .

ppt 4 7

− − – – – – annual in (KAI), Okarito (OKA) and Otautau (OTT), contain E. minus (Wagstaff GNIP standard and Clarkson, 2012). All sites have been defined as ombrotrophic values

=

) ) ) ) ) (e.g. Newnham et al., 1995; Elliot, 1998; Newnham and Lowe, ) SD

O 04 16 23 18 34 23 . . . . . 2000; Campbell et al., 2014) or display characteristics associated . mean 18 1 1 1 1 1 1

Isoscape ( ( ( ( ( ( δ with ombrotrophy, such as typically diagnostic vegetation commu- (OTT).

32 32 29 53 09 34 ...... 7 4 5 6 5 nities, domed or raised profiles in relation to surrounding land and 7 − − − − − − details. annual in ppt (SD) Isoscape an acidic substrate. This selection criteria was applied to provide for confidence that the only water source available to plants was pre- Invercargill

cipitation, though we acknowledge that our sites fall on a gradient text

for from those where we can be very confident of ombrotrophy (KOP, see MOA), to intermediate sites where less information is available Relative humidity (%) (TNG, OKA, OTT) to one site where, given its topographic setting, 1977–1991 the influence of other water sources must be deemed a possibility Network,

(KAI). and

summer We employed both spatial and temporal approaches to sam- Total Station

pling and analysis. For the spatial approach, we sampled six lo- (TNG) cations on each site in November 2012, providing a total dataset Climate

Total (mm) annual of 36 sampling points. At each sampling point, we took two Precipitation Kaitaia Empodisma spp. plant samples (surface cluster roots and recent for

Virtual

shoot growth), two potential external source water samples (one squeezed from the interstitial cluster root water, hereafter referred

summer Mean NIWA to as ‘root-associated water’ and one from the below ground water- 1972–1991 table, hereafter referred to as ‘bog water’ – see Table 2 for mean, the

for

96 14.70 2665 521 82.4 85 19.21 1318 259 84.9 63 18.04 1186 256 82.8 72 15.19 3338 887 84.3 72 19.05 1186 242 80.5 minimum and maximum bog water sampling depths) and a water- 98 14.09 1053 276 82.8 ...... C) from 9 9 ◦

11 13 Temperature 14 15 table depth measurement to test the potential effect of microto- ( Mean annual means pography on evaporative enrichment (Ménot-Combes et al., 2002). Sampling root-associated water provides only an approximation of annual

the internal root water but we use it here in addition to pre- downloaded

cipitation to examine potential sources of external source water. are

E. minus E. minus E. minus E. robustum Empodisma species E. robustum E. robustum was We also took plant samples of recent growth of the most com- (ppt) mon other plant species present at each site, including Sphagnum spp., Machaerina teretifolia, , Leptospermum scopar- humidity

Altitude ium and Dracophyllum longifolium. (m) For the temporal approach, aimed at understanding seasonal precipitation

relative changes, we took monthly samples of root and recent shoot

Din and growth, root-associated and bog water and monthly integrated E) δ

( precipitation at two locations on one site (KOP) from November Long and

2012 to November 2013. Precipitation over the year was measured O 18 using a tipping bucket rain gauge (TB5, Hydrological Services, NSW, δ

S) ◦ precipitation Lat ( 37.3879 175.5552 5 35.1048 173.2156 10 for Australia) situated 0.8 m (mean canopy height) above the sur- face. Wind direction was measured with a three dimensional sonic

anemometer (CSAT3, Campbell Scientific, Utah, USA) mounted on a values tower 4.25 m above the peat surface with continuous half-hourly GNIP temperature, wind statistics collected and stored on a data logger (CR3000, for

Campbell Scientific, Utah, USA), with daily (24 h) means calculated 2012. for analysis. data

2.2. Climate and isoscape data and modelling 1997 to

Climate

from

Climate data for mean monthly temperature, total monthly pre- data. cipitation and monthly relative humidity were downloaded from OKA West Coast 43.2471 170.2205 75 KAI Te Urewera National Park 38.6849 177.2011 1000 OTT Southland 46.1311 168.0324 65 MOA Waikato 37.9235 175.3701 60 Site Region TNGKOP Northland Waikato the NZ National Institute for Water and Atmospheric Research Table 1 Site means 152 M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159

Table 2 Mean values and ranges of root-associated and bog water δ18O(h) and water table depth for all sites in the spatial dataset. WTD = water table depth. Site Mean Range WTD mean WTD range (h) (h) (cm) (min–max, Root water Bog water Root water Bog water cm) TNG −3.3 −5.91.3 1.6 30.5 7 (27–34) KOP −1.6 −6.14.4 2.3 8.3 8 (4–12) MOA −4.5 −5.32.1 0.6 59.7 23 (50–73) KAI −4.4 −5.32.3 2.2 19.3 24 (5–29) OKA −4.8 −4.71.1 1.4 2.516(−5*–11) OTT −7.0 −6.62.0 1.7 7.5 14 (3–17)

Mean all sites −4.3 −5.72.2 1.6 – –

* Minimum value of −5cm for Okarito indicates that cluster roots were 5 cm below the standing water table at the bog surface.

(NIWA) National Climate Database (cliflo.niwa.co.nz; data accessed 2010), but up to 10 equivalent bleaching stages were required for ◦ ◦ May 2014). The 0.05 × 0.05 gridded Virtual Climate Station Net- shoot samples to obtain a product of similar visual purity. work (VSCN; Tait et al., 2006, 2012) cells containing each of our Oxygen (δ18O) and hydrogen (δD) stable isotope ratios of wa- sites were used to calculate long term annual or growing season ter samples were measured at the Lancaster Environment Centre means, or sums for precipitation, from 1972 to 2013. (LEC), Lancaster University, UK, using a Vario Pyrocube with Vario To examine spatial variation in the isotopic composition of pre- Liquid Autosampler (Elementar Analysensysteme, Germany), inter- cipitation we used isoscape modelled values for δ18O and δDfrom faced with an Isoprime 100 isotope ratio mass spectrometer (Iso- the VSCN grid square containing each of our sites. The isoscape prime, UK) set in continuous flow mode. All water samples were model is based on δ18O and δD measurements of monthly pre- filtered prior to analysis to remove small quantities of particulates cipitation samples collected between 2007 and 2009 from 51 sites introduced to the sample during collection. For δD, 1μl of wa- ◦ across NZ (Frew and Van Hale, 2011), which were used to develop ter was reduced to hydrogen gas at 1080 C over a chrome-based a multiple linear regression model (including two geographic and catalyst. δ18O was determined by pyrolysis of 0.5 μl of water at ◦ five daily meteorological VSCN variables; Ehtesham et al., 2013). 1450 C over glassy carbon. Internal (LEC tap water #2) and in- All meteorological variables were weighted by daily precipitation ternational standards (VSMOW, GISP) were run with each batch of amounts to generate monthly-weighted climate means correspond- samples with analytical precision typically better than 0.4h and ing to the collection times of the precipitation samples. These 1h and for δ18O and δD respectively. weighted means, as well as latitude and elevation, were regressed, Oxygen (δ18O) and hydrogen (δD) stable isotope ratios of cellu- using JMP v.8 (SAS Institute, North Carolina, USA), against all avail- lose samples were measured at Swansea University, UK, following able monthly δ18O and δDobservations using stepwise backward the simultaneous COH method described by Loader et al. (2014). elimination, supervised to eliminate highly correlated variables (re- Samples of 300–350 mg dried α-cellulose were passed from an au- gression statistics and equation in Table S1). VCSN climate data tosampler (Filot et al., 2006)into an equilibration chamber heated ◦ were then used with the final regression equation to calculate to 110 C, through which steam of known isotopic composition an estimate of daily- and monthly-weighted precipitation δD and (IAEA-CH7 and IAEA-C3) was passed in a flow of helium. Sam- ◦ δ18O for each point on the 0.05 × 0.05 grid, processed in Matlab ples were left to equilibrate for 600 s, dropped into a Flash HT R2012b (The Mathworks, Massachusetts, USA). The modelling and elemental analyser (Thermo GmbH, Germany) and pyrolysed over ◦ mapping procedure was validated by Ehtesham et al. (2013). glassy carbon at 1400 C prior to mass-spectrometric analysis using GNIP (Global Network of Isotopes in Precipitation) data for δ18O a Thermo Delta V mass spectrometer (Thermo GmbH, Germany). and δDin precipitation from Kaitaia (1972–1991) and Invercargill IAEA-CH7 standards were run with each batch of samples. Stan- 18 (1977–1991; IAEA/WMO, 2014; data accessed July 2014) provided dard precision was ±0.30h and ±3.0h for δ O and δDrespec- a lower resolution check on the isoscape data. tively (Loader et al., 2014). 18 NOAA (US National Oceanic and Atmospheric Administration) Stable oxygen (δ O) and hydrogen (δD) isotope ratios were re- h HYSPLIT (HYbrid Single Particle Lagrangian Integrated Trajectory; ported as permil ( ) deviations from the VSMOW standard where: Draxler and Rolph, 2014) back-trajectory models were run to ex- Rsample amine precipitation moisture source regions over the course of the δ18OorδD (h) = − 1(Coplen, 2011)(1) year-long temporal dataset at KOP. We modelled air mass trajec- Rstandard −1 tory for the 72 h before the rain day (>5mmday ) prior to each where R is the ratio of 18O/16O or 2H(D)/1H in the sample and sampling day. We used trajectory type 1 (ensemble), the GDAS standard. A mean of three isotope measurements was calculated global archived meteorological dataset (2006 – present) and a level for each sample. Since neither δ18O or δDare independent vari- height of 250 m above ground level. ables, all water lines presented in δ18O/δDbiplots (Fig. 2) are calculated using Deming regression, a form of orthogonal linear 2.3. δ18O and δD analyses regression which assumes error for both x and y variables.

Root and shoot samples were prepared to α-cellulose using a 3. Results standard methodology (Loader et al., 1997; Daley et al., 2010). Roots were cleaned with distilled water and manually separated 3.1. Precipitation and cellulose source water from other plant material; each sample was wet-sieved to remove fine detritus and all other plant remains then removed using a Mean annual δ18O in precipitation from isoscape modelled data low-powered microscope at ×10 magnification. Roots and shoots for NZ and GNIP sites at Kaitaia and Invercargill provided an in- were cut into short (approximately 1 mm) sections prior to chemi- dication of the spatial variability in the isotopic composition of cal treatment (bleaching). Five 1 h bleaching stages were sufficient precipitation (Table 1, Fig. 1). Strong gradients exist that represent to form pure white cellulose for the root samples (Daley et al., source and temperature effects on δ18O (Barbour, 2007), with more M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 153

Fig. 2. (A) δ18O/δDbiplot of mean monthly 1997–2013 isoscape modelled precipitation from all sites and GNIP data from Kaitaia and Invercargill compared to root-associated water and bog water samples from all sites in the spatial dataset. Global meteoric water line (GMWL; blue) and a local meteoric line derived from the combined isoscape and GNIP data (red) are shown. (B) δ18O/δDbiplot of mean monthly 1997–2013 isoscape modelled precipitation from all sites and monthly measured precipitation compared to root-associated water and bog water (means of two monthly replicate values are plotted for root-associated and bog water) samples from KOP in the temporal dataset. GMWL = global meteoric water line. For (A) and (B), Pearson’s r values are given for the correlation between δ18O and δDfor the relevant datasets. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

18O enriched values occurring in the Northland and Waikato re- This offset is primarily driven by three sites (TNG, OKA and OTT), gions where a warmer climate is more influenced by sub-tropical whereas root-associated waters at KOP, MOA and KAI are more moisture sources and more 18O depleted values in the South Is- consistent with the slope and intercept of the LMWL (Fig. 2A). land where a cooler climate is more dominated by the prevailing Mean site root-associated water and isoscape precipitation val- southern ocean westerly winds. GNIP data support this general ues are strongly positively correlated but this is not significant pattern with precipitation from the southerly, cooler Invercargill (r = 0.689, p = 0.13). being more 18O depleted than Kaitaia precipitation (Table 1). Site- Bog water δ18O and δDvalues from the spatial dataset are sig- specific isoscape data show more 18O enriched values for TNG, nificantly but weakly positively correlated (r = 0.384, p = 0.02) KOP and MOA, with OKA and OTT more 18O depleted (Table 1). and occur over a much narrower range of values than the root- In addition, altitude (more 18O depleted values with increasing associated water samples, suggesting that bog water is more iso- altitude) and, to a lesser extent, rainout (progressively more 18O topically homogeneous. The weak response of bog water to the depleted precipitation moving inland from the coast) effects also isotopic composition of precipitation is also evident in its relation- contribute to the overall variability in δ18O observed, particularly ship to latitude. Whereas root-associated water shows a clear 18O at KAI (Fig. 1). A local meteoric water line (LMWL) for NZ derived depletion effect with increasing southerly latitude (significant neg- from the combined isoscape and GNIP datasets is highly significant ative correlation r =−0.72, p < 0.001), bog water values remain (r = 0.97, p < 0.001) and conforms to the global meteoric water relatively stable across the latitudinal gradient sampled (r =−0.11, line (GMWL) with a slope of 7.97 (Fig. 2A). p = 0.53). Full results for root-associated and bog water δ18O for the spa- In the temporal dataset, root-associated water and bog water tial and temporal datasets are given in Tables S2 and S3. In the δ18O and δDare both significantly correlated (root-associated wa- spatial dataset, root-associated water δ18O and δDare significantly ter r = 0.94, p < 0.001; bog water r = 0.80, p < 0.001) and water positively correlated (r = 0.89, p < 0.001) and also broadly con- lines conform to the slope and intercept of the measured and form to the expected pattern along the GMWL with waters from isoscape-modelled precipitation (Fig. 2B), as well as the GMWL, the warmer, lower latitude North Island sites generally positioned without the offset evident in the spatial dataset (Fig. 2A). Root- higher compared to the cooler, higher latitude South Island sites associated water δ18O co-varied strongly with precipitation over (Fig. 2A). Waters from KOP and TNG are the most 18O and D en- the year of sampling whereas bog water again showed less vari- riched, suggesting evaporative enrichment at these sites, with more ability, with bog water values occurring over a narrower range overlap and a less clear gradient between the remaining sites, (Fig. 3B). Furthermore, while both root-associated and bog water though, on average, water from OTT is the most 18O and D de- were positively correlated with precipitation, r values were higher pleted. A water line derived from root-associated waters is broadly for root-associated water–precipitation (r = 0.85, p < 0.001) and parallel to (slope of 7.43), but offset above the GMWL and LMWL. the bog water–precipitation relationship was not significant at 154 M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159

Table 3 Rainfall events prior to sampling days in the temporal dataset at KOP for which NOAA HYSPLIT back trajectory models were run to determine precipitation moisture source. Precipitation and wind data are from climate station within 100 metres of sampling points (see Section 2.1). Moisture source regions are defined sensu Sturman and Tapper (2006). See also Fig. S1. Sampling date Date of previous Rainfall amount Wind direction Wind Moisture source region rainfall event (mm) on rain day compass ◦ ( ) 08/11/2012 29/10/2012 6.9 128 SE Tropical maritime Tasman 11/12/2012 06/12/2012 12.4 269 W Tropical maritime Pacific 11/01/2013 30/12/2012 6.3 247 WSW Tropical maritime Pacific 08/02/2013 05/02/2013 7.7 241 WSW Tropical maritime Pacific 12/03/2013 03/03/2013 3.2 244 WSW Southern maritime 09/04/2013 05/04/2013 10.6 172 S Southern maritime 22/05/2013 21/05/2013 7.3 156 SSE Tropical maritime Pacific 11/06/2013 09/06/2013 47.7 144 SE Modified Polar maritime 23/07/2013 14/07/2013 9.3 170 S Modified Polar maritime 21/08/2013 19/08/2013 12.0 162 SSE Modified Polar maritime 27/09/2013 24/09/2013 42.6 139 SE Tropical maritime Tasman 05/11/2013 31/10/2013 18.5 279 W Tropical maritime Pacific

α = 0.05 (r = 0.55, p = 0.08). The root-associated water samples from the summer months (Dec.–Apr.) were more 18O enriched than most of the precipitation values from the same period, sug- gesting that evaporative enrichment was occurring (Fig. 3B). This is also evident as an evaporative trend from the GMWL at more enriched values (Fig. 2B). Previous work on Sphagnum peatlands has shown that microto- pography can have a significant effect on evaporative enrichment, leading to variability in source water for mosses growing at dif- ferent points along a hummock to hollow gradient (Ménot-Combes et al., 2002). Although microtopographical variation on Empodisma peatlands is less than typically observed on Sphagnum peatlands, we tested whether this had any effect at our sites. There were no significant correlations between water-table depth and root- associated or bog water for either δ18O or δDalthough the δ18O variability in root-associated water (7.75h) was greater than re- ported in surface waters across a hummock to hollow gradient in a Sphagnum peatland (4h; Ménot-Combes et al., 2002). Mean site δ18O and δDvalues for root-associated water were consis- tently more 18O and D enriched and showed more variability than mean site values for bog water, suggesting a degree of evaporative enrichment in root-associated waters and a dampening of the pre- cipitation signal in bog waters (Table 2). Neither of these effects were related to water-table depth. The difference between bog and root-associated waters was highest in the warmer northern sites of KOP and TNG, whereas other sites had more similar or overlapping 18 18 bog and root-associated water ranges, suggesting that the evapora- Fig. 3. (A) δ O in root and shoot cellulose. Note different y axis scales. (B) δ O in root-associated water, bog water and precipitation with indication of precipitation tive effect is highest in the northern, warmer sites, but is unrelated moisture source. Data in both panels as measured from monthly samples taken at to water-table depth. KOP from November 2012 to November 2013. For root and bog waters and root and To examine the different moisture sources for precipitation over shoot cellulose, the values plotted are means of results from two sampling locations. the year of sampling at KOP, we ran HYSPLIT back-trajectory mod- Sampling dates and moisture source regions are shown in Table 3. For a full colour − els (Draxler and Rolph, 2014)for the rain days (>5mmday 1) version of this figure, the reader is referred to the web version of this article prior to each sampling day and characterised moisture source re- gions sensu Sturman and Tapper (2006) in Table 3. Tropical mois- 3.2. Source waters and plant cellulose ture source regions are characterised by a more heavy isotope 18 enriched signature, whereas those with a more distant, cooler Full results for Empodisma root and shoot δ OC for the spatial 18 southerly moisture source originating in the Southern Ocean are and temporal datasets are given in Tables S4 and S5. Shoot δ OC 18 characterised by a more heavy isotope depleted signature. It is values were consistently higher than root δ OC (Fig. 3A; mean = clear from both the wind direction and moisture source regions 39.7 ± 2.4h (1σ ) for shoots versus 31.3 ± 0.9h for roots in the that there is a seasonal shift in weather patterns delivering precipi- spatial dataset (n = 36) and 44.9 ±2.7h for shoots versus 33 ±1h tation to KOP over the year, with more westerly winds and tropical for roots in the temporal dataset (n = 24)). The wide range of 18 moisture sources dominating during the summer months and more δ OC values (3.3h and 8h for spatial roots and shoots and 3.3h southerly winds and southern/polar moisture sources dominating and 9h for temporal roots and shoots respectively) reflected the during the winter months. This is reflected in both the precipi- wide range of values for the isotopic composition of precipitation tation and root-associated water results in the temporal dataset both over the site transect and the year of sampling (Table 1). (Fig. 3) which show more 18O and D enriched values when tropical The difference in magnitude of permil variation between roots and moisture sources were dominant and relatively 18O and D depleted shoots may be a genuine reflection of relative source water val- values when southern/polar moisture sources were dominant. ues if internal shoot water is more evaporatively enriched than M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 155

Table 4 Correlation coefficients (r values) for δ18O root and bog water with root and shoot cellulose. p values <0.1 are shown in parentheses. Spatial dataset correlations are for mean δ18O of 6 site within site replicates for water and cellulose samples and GPP-weighted annual mean values for precipitation (see Section 3.3 for details). Temporal correlations are for mean δ18O of two monthly replicate measurements for water and cellulose samples. Precipitation was collected for 11 months. Root cellulose Shoot cellulose n δ18O δ18O Spatial dataset Root water −0.278 −0.068 6 Bog water −0.274 −0.498 6 Precipitation −0.124 −0.019 6

Spatial Root water −0.487 −0.141 5 dataset Bog water −0.148 −0.481 5 (without KAI) Precipitation −0.911 (0.032) −0.603 5

Temporal dataset Root water −0.405 −0.153 12 Bog water −0.379 −0.464 12 Precipitation −0.597 (0.053) 0.043 11 internal root water when cellulose synthesis takes place, or may only reflect sampling regime (see below). Root, and particularly 18 shoot, δ OC values were significantly higher than those reported for other peatland vascular plants (e.g. 27–30h, Brenninkmeijer et al., 1982; about 24h, Ménot-Combes et al., 2002; 20–25h, Hong et al., 2009). Mean δ18O values for the samples of other vascular C Fig. 4. Epsilon/water biplot of δ18O in potential source waters (precipitation, root- = plants growing at our sites (n 20) were also higher than previ- associated water, bog water) with enrichment (ε) between potential source water ous studies at 30.7h ± 2.4h (1σ ), equivalent to Empodisma spp. and root cellulose for the spatial dataset. For root-associated water and bog water, 18 18 means are of six within site replicates, for precipitation, 1997–2013 mean monthly root δ OC, but again much lower than shoot δ OC (Table S6). Root-associated or bog water and root or shoot δ18O were not values are used. For a full colour version of this figure, the reader is referred to the C web version of this article significantly correlated in either the spatial or temporal datasets (Table 4), though in the temporal dataset, precipitation and root tween source waters and cellulose, with higher ε where source cellulose were relatively strongly negatively correlated, which was water is more 18O depleted. almost significant at α = 0.05 (r =−0.59, p = 0.053). When KAI However, δ18O/δDbiplots for the spatial (Fig. 5A) and tempo- was removed from the spatial dataset as a potential outlier (Sec- ral (Fig. 5B) datasets show that cellulose values were considerably tion 2.1), a strong and significant negative correlation between pre- more scattered than root-associated water and precipitation. In cipitation and root cellulose was evident (r =−0.911, p = 0.032; the temporal dataset, outlying values occurred in both the sum- Table 4). Fig. 4 further supports the strong negative correlation be- mer and winter months suggesting a lack of obvious seasonality

Fig. 5. δ18O/δDbiplots (standardised z-scores) of the isotopic composition of precipitation (mean monthly 1997–2013 values from the modelled isoscape data for spatial dataset, measured precipitation for temporal dataset; open circles), root-associated water (closed circles), root cellulose (dark grey circles) and shoot cellulose (light grey circles) for (A) the spatial dataset and (B) the temporal dataset. Outlying root cellulose values in (B) are labelled with sampling month. 156 M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159

Fig. 6. δ18O offset (h) between measured and modelled root cellulose values for the spatial (A) and temporal (B) datasets. Site names are defined in Section 2.1. Black bars = source water in model is precipitation; white bars = source water in model is root-associated water. in the discrepancy between cellulose and source water values. In transpiration and is dependent on the relative humidity over the the temporal dataset (Fig. 3), shoot cellulose broadly tracked root- evaporating surface and the airflow dynamics in the leaf bound- associated water and precipitation, with more 18O enriched values ary layer, which is controlled by leaf size and morphology, with in late summer (March and April) and a decline to more 18O plants with smaller, segmented leaves having a lower component depleted values over the winter months. Conversely, root cellu- of turbulence, resulting in higher εk values (Buhay et al., 1996; lose δ18O showed an antiphase relationship with precipitation and Araguas-Araguas et al., 2000). Several different values have been 18 root-associated water, with relatively O depleted values in the published for εk ranging from 16h for a turbulent boundary layer summer months when source waters were more 18O enriched and (Dongmann et al., 1974; Yu et al., 2011)to 28h for a stagnant 18 an apparent O enrichment of root cellulose in the winter months boundary layer (e.g. Aucour et al., 1996; Ménot-Combes et al., 18 when source waters were more O depleted, though the varia- 2002). Buhay et al. (1996) present a range of 18.4 ± 6.1h, de- tion of root cellulose values occurred over a narrower permil range pendent primarily on leaf morphology. Given the scale-like leaves when compared to shoot cellulose. This disconnect between shoot and high water use efficiency of Empodisma spp. (Wagstaff and and root cellulose may partly reflect sampling regime. Shoot sam- Clarkson, 2012), we used the upper limit of 28.5h for εk, sensu ples were formed from well-defined recent growth that reflected Ménot-Combes et al. (2002). We use the ±3h error associated monthly precipitation variation, as the cellulose was newly synthe- with biochemical fractionation (Yakir and DeNiro, 1990; Table S7) sised, whereas the root cellulose, extracted from samples in which to assess whether model predicted δ18O is an accurate reflection new growth was difficult to precisely delineate, reflected a longer- of measured δ18O. term average of source water inputs. We used mean annual temperature for TNG, KOP and MOA as these sites have year round growing seasons (e.g. Campbell et 3.3. Mechanistic modelling al., 2014; Goodrich et al., 2015) and mean November to March temperature for KAI, OKA and OTT to represent the shorter grow- 18 A mechanistic model was applied to predict δ OC from ing season in these locations. To reflect these differences and 18 δ OSW, relative humidity and the equilibrium (εe), kinetic (εk) better represent the seasonality of Empodisma spp. growth, we and biochemical (εb) fractionation processes. A range of mod- used a gross primary productivity (GPP) weighted mean of pre- els have been presented by different authors (e.g. Ménot-Combes cipitation δ18O. This was calculated by using the strong, signif- et al., 2002; Waterhouse et al., 2002; Zanazzi and Mora, 2005; icant relationship between monthly air temperature and GPP at Barbour, 2007; Daley et al., 2010; Yu et al., 2011; Yu, 2013). Such KOP (monthly measurements from October 2012 to January 2014; models have been criticised due to their simplicity (Barbour, 2007) r = 0.9, p < 0.001; Fig. S2) to predict monthly GPP at all sites but are used here since more detailed physiological models require based on mean monthly (1972–2013) NIWA VSCN temperature input values for processes for which measurements were not avail- data. We then used monthly proportions of total annual GPP and able. Despite minor differences in form, the general approaches isoscape modelled mean monthly precipitation δ18O to calculate of these models are similar and are summarised by Zanazzi and a GPP-weighted mean annual value for precipitation δ18O at each Mora (2005) and Daley et al. (2010) as (definitions of all terms in site. Table S7): For both the spatial (Table S8) and temporal (Table S9) datasets, 18 18 δ OC = δ OS + εb + (εe + εk)(1 − h) (2) using root-associated water rather than precipitation as the source water for cellulose synthesis resulted in predicted δ18O values where 18     closer to measured δ O on average (Fig. 6), though in the spatial 1137 0.4156 dataset, both potential water sources resulted in mean offsets less εe = exp − − 0.0020667 − 1 h T2 T than the 3 error associated with biochemical fractionation. That root-associated water is a marginally better predictor of δ18O may × 1000 (Yu, 2013)(3) C support the earlier assertion that, despite the low levels of evap-

Calculated values for εe for each site (Eq. (3), Table S8) were oration from the surface of Empodisma-dominated bogs (Campbell consistent with those produced using globally distributed GNIP and Williamson, 1997), some degree of evaporative enrichment of ◦ ◦ data for temperatures between 15 and 20 C (10.3h at 15 C to precipitation has taken place in our root-associated water samples. ◦ 9.8h at 20 C) by Araguas-Araguas et al. (2000). Equation (2) as- Root-associated water was more 18O enriched than precipitation at sumes equilibrium between source water and atmospheric water all sites with the exception of OTT. To further examine the relative vapour. This was assumed for the growing season in central Eu- roles of root-associated water and precipitation as potential source rope (Ménot-Combes et al., 2002) and is a reasonable assumption waters, we applied the mechanistic model to surface plant sam- in our case given the demonstrated correlation between precipita- ples of the most common other vascular plant species at each site tion and root-associated water. Kinetic fractionation occurs during (Table S6) and a simplified model for mosses (Daley et al., 2010), M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 157

which excludes εe and εk, to Sphagnum samples taken at four sites months when source water is most enriched and stomata are (Table S10), where: only open for a relatively short part of the morning when evap- oration potential is low (Sharp, 1995, quoted in Campbell and 18 18 δ OC(moss) = δ OS + εb (4) Williamson, 1997), there may be relatively less evaporative enrich- ment of leaf water, with the summer water signal more tightly For both vascular plants and Sphagnum, mean modelled–measured impressed. In contrast, in winter when source water is more heavy offsets for both potential source waters fell within the ±3h er- isotope depleted, shoots may be less water-stressed and so assimi- ror range for biochemical fractionation (Tables S6 and S10), though lating with open stomata for longer periods, resulting in compara- root-associated water was once again a marginally better predictor tively more evaporative enrichment, driving the observed negative of δ18O than precipitation. C correlation. Secondly, in winter months, when growth conditions In the spatial dataset, using either root-associated water or pre- may be marginal, shoots may remobilise stored starches from roots cipitation as the input values for source water resulted in average to form new sucrose. If this translocation favoured 16O over 18O, modelled–measured offset for root δ18O < 3h. In both cases, the then roots would become more 18O enriched and shoots more 18O four North Island sites fell within this margin, whereas the two depleted, as observed (Fig. 3A). In order to drive the observed neg- South Island sites had offset values >3h (Table S8; Fig. 6). The ative relationship, this hypothesis would be dependent on a low modelled–measured offset for shoot δ18O was consistently higher level of exchange occurring between organically bound oxygens (mean = 10.31h for source water = root-associated water) as a with those in source water during photosynthesis. This figure has result of the much higher δ18O for shoots compared to roots. C been estimated at 40–100% for the aquatic plant Lemna gibba un- In the temporal dataset (Table S9), the annual mean modelled– der different growth conditions (Yakir and DeNiro, 1990), but the measured offset for root δ18O (using root-associated water as the figure for Empodisma spp. specifically remains unknown. The ab- input value) was within acceptable error (2.25h), but the equiva- sence of detailed physiological studies on, and understanding of, lent value for shoot δ18O was again much higher (14.28h). Annual Empodisma spp. is therefore highlighted and must be addressed if mean values hid distinct seasonality in the temporal data since its isotopic fractionation pathways are to be fully understood, in offset values were consistently lower from November to April com- pared to the winter months (May to September; Table S9; Fig. 6). order to maximise the potential for palaeoclimate proxy develop- ment. In both the spatial and temporal datasets, there was a nar- 4. Discussion: δ18O relationships and development of a rower range of δ18O values for root cellulose compared to root- palaeoclimate proxy 18 associated water (spatial: root δ OC range = 3.3h, root-associated water range = 7.8h; temporal: root δ18O range = 3.3h, root- The observed relationships between the isotopic composition C associated water range = 14.4h), an effect which has been pre- of precipitation, root-associated water and bog water show that viously observed along an altitudinal transect of bogs in Switzer- root-associated water reflects the isotopic composition of precipi- land (Ménot-Combes et al., 2002). One possible explanation is tation, and therefore the precipitation moisture source, over time that δ18O values reflect an integrated longer-term average of and space (Fig. 2). Our first hypothesis is therefore upheld. Con- C versely, bog water, which represents a larger water volume, which more variable source water conditions over the growing season is likely to have a longer turnover time than root-associated water, (Ménot-Combes et al., 2002), as the range is close to that observed h forms a more isotopically homogeneous pool in which the contem- for the site averages for precipitation (3.02 , Table 1). Further- porary precipitation signal is dampened. This result is consistent more, results from the modelling of temporal data showed bet- with Ménot-Combes et al. (2002) who demonstrated, using short ter performance during the growing season, when active growth vertical profiles of δ18O in a Sphagnum-dominated peatland, that and cellulose synthesis effectively tracks changes in source water, water from depth was more isotopically homogeneous than sur- which is also seen in the tracking of source waters by shoot cellu- face water. lose in the temporal dataset (Fig. 3). Conversely, higher modelled- 18 However, the relationships between precipitation/root-associ- measured δ OC values were observed during winter months (Ta- ated water and root/shoot cellulose were less clear (Table 4), ble S9), when growth slows or ceases (Goodrich et al., 2015). How- meaning that our second hypothesis (that root cellulose reflects ever, rather than an effect of sensitivity, this seasonal differential in spatial and temporal variation in the source water) cannot cur- model performance for the temporal dataset may reflect the more 18 rently be confidently upheld. Although we selected sites that have negative δ O values for root-associated water and precipitation in 18 been defined as ombrotrophic in previous publications or display the winter months, which are related to dominant polar (more O characteristics associated with ombrotrophy, the strong and sig- depleted) moisture sources, as opposed to the dominant tropical 18 nificant negative relationship observed between precipitation and (more O enriched) moisture sources during the summer months root cellulose in the spatial dataset when KAI was removed (Ta- (Table 3). ble 4) suggests that this site may be subject to external hydrolog- The observed relationship between root-associated water and ical influences, most likely either surface run-off from surround- precipitation is convincing in both spatial and temporal datasets. ing upland slopes or heavy isotope depleted snowmelt, and may This, in addition to biochemical enrichment coefficients falling not therefore be fully ombrotrophic. The negative relationship be- within error for the simplified modelling of Sphagnum cellulose tween precipitation and root cellulose evident in the temporal samples (Table S10), suggests that root-associated water is a po- and reduced spatial datasets (Figs. 3 and 4; Table 4) is in con- tential indicator of the internal leaf water. However, understand- trast to modelled predictions but may offer some insight into the ing the relationship between the external source water (i.e. root- link between source waters and cellulose. We propose two po- associated water as measured here) and the internal water used tential mechanisms that may explain this observation. Firstly, Em- in cellulose synthesis is essential for understanding the relation- podisma-dominated peatlands have previously been described as a ship between source water and plant cellulose. Measurements of ‘wet desert’ (Campbell and Williamson, 1997), with large canopy δ18O of extracted ‘true’ root and leaf water (e.g. by vacuum dis- resistance causing a hydrological separation of the dry shoots and tillation; West et al., 2006) would allow the extent of any frac- wet roots, whereby new growth shoots in the upper canopy may tionation between external source water and internal root and become water-stressed even though roots below the canopy re- leaf water to be quantified, but this was not undertaken in this main saturated. If shoots were most water-stressed in summer study. 158 M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159

The good performance of the mechanistic modelling is partly Acknowledgements dependent on the values calculated/used for different fractionation processes. Calculated values for εe fell within published ranges This research was funded by the UK Natural Environment Re- and εb is widely accepted and has been consistently applied at search Council small grant NE/J013595/1. We gratefully acknowl- 27 ± 3h in many isotopic studies (e.g. Ménot-Combes et al., 2002; edge Sue Rouillard in the University of Exeter Geography drawing Waterhouse et al., 2002; Zanazzi and Mora, 2005; Daley et al., office for producing Fig. 1. Thanks to Dave Hughes at the Lancaster 2010; Yu et al., 2011; Yu, 2013). εk has the highest uncertainty University Environment Centre for running stable isotope analyses with a value of 28.5h chosen based on a review of literature and on our water samples. We gratefully acknowledge all landowners reasonable assumptions on the leaf form and physiology of Em- for their permission to access the sites used for this research; the podisma spp., but if kinetic fractionation was lower (e.g. Buhay et Te Rarawa and Ngai Takoto iwi at Tangonge, the Ngati Apakura 18 al., 1996 suggest a mean value of 18.4h) then modelled δ OC val- iwi at Moanatuatua and the NZ Department of Conservation at all ues would fall outside of the ±3h error when compared to mea- other sites. 18 sured root δ OC. In addition, the broadly applied assumption of the temperature independence of εb has also recently been ques- Appendix A. Supplementary material tioned (Sternberg and Ellsworth, 2011). Experiments generating cellulose at different temperatures showed that εb was consistent Supplementary material related to this article can be found on- ◦ at a mean value of 26h above temperatures of 20 C but increased line at http://dx.doi.org/10.1016/j.epsl.2015.08.015. ◦ below 20 C to up to 31h (Sternberg and Ellsworth, 2011). Our sites are located in a climate space that spans similar tempera- References ture gradients (Table 1) so it may be reasonable to expect that εb is higher at the higher latitude, lower temperature sites (OKA and Agnew, A.D.Q., Rapson, G.L., Sykes, M.T., Bastow Wilson, J., 1993. The functional OTT). This would reduce the higher modelled–measured δ18O off- ecology of Empodisma minus (Hook, f.) Johnson & Cutler in New Zealand om- brotrophic mires. New Phytol. 124, 703–710. set seen for these sites in the spatial dataset (Table S8), resulting Araguas-Araguas, L., Froehlich, K., Rozanski, K., 2000. Deuterium and oxygen-18 18 from higher mean root/shoot δ OC at OTT and OKA compared to isotope composition of precipitation and atmospheric moisture. Hydrol. Pro- the North Island sites. Our data therefore lend tentative empirical cess. 14, 1341–1355. support to the experimentally observed temperature dependence Aucour, A.-M., Hillaire-Marcel, C., Bonnefille, R., 1996. Oxygen isotopes in cellulose of ε and provide impetus for this issue to be addressed in field from modern and quaternary intertropical peatbogs: implications for palaeohy- b drology. Chem. Geol. 129, 341–359. studies. Barbour, M.M., 2007. Stable oxygen isotope composition of plant tissue: a review. There was a consistent offset of mean value 10.2h (range Funct. Plant Biol. 34, 83–94. 18 6.1h–16h) between root and shoot δ OC across both spatial and Brenninkmeijer, C.A.M., van Geel, B., Mook, W.G., 1982. Variations in the D/H and 18 16 temporal datasets, which may reflect the hydrological separation of O/ O ratio in cellulose extracted from a peat core. Earth Planet. Sci. Lett. 61, 283–290. shoots and roots caused by large canopy resistance (Campbell and Brook, E.J., 2007. Ice core methods: stable isotopes. In: Encyclopedia of Quaternary Williamson, 1997). Mechanistic models predicted Empodisma spp. Science, pp. 1212–1219. 18 root, but not shoot, δ OC values to within the error margin as- Buhay, W.M., Edwards, T.W.D., Aravena, R., 1996. Evaluating kinetic fractionation fac- sociated with biochemical fractionation, strengthening the case for tors used for ecologic and paleoclimatic reconstructions from oxygen and hydro- 18 gen isotope ratios in plant water and cellulose. Geochim. Cosmochim. Acta 60, root δ OC as the primary target for proxy development. Results 2209–2218. of mechanistic modelling indicate that improved understanding Campbell, D.I., Williamson, J.L., 1997. Evaporation from a raised peat bog. J. Hy- of oxygen isotope fractionation in Empodisma-dominated restiad drol. 193, 142–160. peatlands could enable use of this approach as a new source of Campbell, D.I., Smith, J., Goodrich, J.P., Wall, A.M., Schipper, L.A., 2014. Year-round palaeoclimate data to reconstruct changes in past atmospheric cir- growing conditions explains large CO2 sink strength in a New Zealand raised peat bog. Agric. For. Meteorol. 192–193, 59–68. culation. However, our research has raised several important, as Cole, J.E., Rind, D., Webb, R.S., Jouzel, J., Healy, R., 1999. Climatic controls on in- yet untested hypotheses regarding the plant physiology and iso- terannual variability of precipitation δ18O: simulated influence of temperature, topic fractionation of Empodisma spp., which should be addressed precipitation amount, and vapor source region. J. Geophys. Res., Atmos. 104, to enhance the potential of this proxy archive to provide valu- 14223–14235. able palaeoclimatic reconstructions in a climatically sensitive re- Coplen, T.B., 2011. Guidelines and recommended terms for expression of stable isotope-ratio and gas-ratio measurement results. Rapid Commun. Mass Spec- gion lacking in other suitable proxies. trom. 25, 2538–2560. Craig, H., Gordon, L.I., 1965. Deuterium and oxygen-18 variations in the ocean and 5. Conclusions the marine atmosphere. In: Tongiorgi, E. (Ed.), Stable Isotopes in Oceanographic Studies and Paleotemperatures. Consiglio Nazionale Delle Richerche, Spoleto, Italy, pp. 9–130. When oxygen stable isotope ratios are used as a proxy for Daley, T.J., Barber, K.E., Street-Perrott, F.A., Loader, N.J., Marshall, J.D., Crowley, S.F., past changes in new geographical or ecological contexts, it is criti- Fisher, E.H., 2010. Holocene climate variability revealed by oxygen isotope anal- cal that interpretation is empirically and mechanistically grounded ysis of Sphagnum cellulose from Walton Moss, northern England. Quat. Sci. and calibrated with modern climate data, with a clear under- Rev. 29, 1590–1601. standing of the links between climate, source water and cellulose Daley, T.J., Mauquoy, D., Chambers, F.M., Street-Perrott, F.A., Hughes, P.D.M., Loader, N.J., Roland, T.P., van Bellen, S., Garcia-Meneses, P., Lewin, S., 2012. Investigat- δ18O values. To test the development of stable isotope palaeocli- ing late Holocene variations in hydroclimate and the stable isotope composition mate records from vascular plant-dominated peatlands, we sam- of precipitation using southern South American peatlands: an hypothesis. Clim. pled precipitation, surface and sub-surface waters and root and Past 8, 1457–1471. shoot plant material over spatial and temporal gradients in NZ. Dansgaard, W., 1964. Stable isotopes in precipitation. Tellus 16, 436–468. Dongmann, G., Nürnberg, H.W., Förstel, H., Wagener, K., 1974. On the enrichment of Variation in root-associated water tracked precipitation over both 18 H2 O in the leaves of transpiring plants. Radiat. Environ. Biophys. 11, 41–52. spatial and temporal gradients. The link between source water and Draxler, R.R. Rolph, G.D. 2014. HYSPLIT (HYbrid Single-Particle Lagrangian plant cellulose was less clear, although mechanistic modelling pre- Integrated Trajectory) model access via NOAA ARL READY website 18 dicted mean root δ OC values within published error margins for (http://www.arl.noaa.gov/HYSPLIT.php). NOAA Air Resources Laboratory, College both datasets. Improved understanding of Empodisma physiology Park, MD. Ehtesham, E., Baisden, W.T., Keller, E.D., Hayman, A.R., Van Hale, R., Frew, R.D., 2013. and isotope fractionation should enable the development of this Correlation between precipitation and geographical location of the δ2H values approach as a new source of palaeoclimate data to reconstruct of the fatty acids in milk and bulk milk powder. Geochim. Cosmochim. Acta 111, changes in past atmospheric circulation. 105–116. M.J. Amesbury et al. / Earth and Planetary Science Letters 430 (2015) 149–159 159

Elliot, M.B., 1998. Late Quaternary pollen records of vegetation and climate change brotrophic peatlands as a quantitative proxy for precipitation—evaporation bal- from Kaitaia Bog, far northern New Zealand. Rev. Palaeobot. Palynol. 99, ance. Geochim. Cosmochim. Acta 74, 1407–1416. 189–202. Oldfield, F., 2001. A question of timing: a comment on Hong, Jiang, Lui, Zhou, Beer, Filot, M.S., Leuenberger, M., Pazdur, A., Boettger, T., 2006. Rapid online equilibration Li, Leng, Hong and Qin. Holocene 11, 123–124. method to determine the D/H ratios of non-exchangeable hydrogen in cellulose. Royles, J., Sime, L.C., Hodgson, D.A., Convey, P., Griffiths, H., 2013. Differing source Rapid Commun. Mass Spectrom. 20, 3337–3344. water inputs, moderated by evaporative enrichment, determine the contrasting 18 Fletcher, M.-S., Moreno, P.I., 2012. Have the Southern Westerlies changed in a zon- δ OCELLULOSE signals in maritime Antarctic moss peat banks. J. Geophys. Res., ally symmetric manner over the last 14 000 years? A hemisphere-wide take on Biogeosci. 118, 184–194. a controversial problem. Quat. Int. 253, 32–46. Sternberg, L., Ellsworth, P.F.V., 2011. Divergent biochemical fractionation, not conver- Frew, R.D., Van Hale, R.J., 2011. A stable isotope rainfall map for the protection of gent temperature, explains cellulose oxygen isotope enrichment across latitudes. New Zealand’s biological and environmental resources. Report to New Zealand PLoS ONE 6, e28040. http://dx.doi.org/10.1371/journal.pone.0028040. Ministry for Primary Industries, Biosecurity. 34 pp. Sturman, A.P., Tapper, N.J., 2006. The Weather and Climate of Australia and New Gat, J.R., 2000. Atmospheric water balance – the isotopic perspective. Hydrol. Pro- Zealand, second edition. Oxford University Press, Melbourne. cess. 14, 1357–1369. Tait, A., Henderson, R., Turner, R., Zheng, X.G., 2006. Thin plate smoothing spline Goodrich, J.P., Campbell, D.I., Schipper, L.A., Clearwater, M.J., Rutledge, S., 2015. interpolation of daily rainfall for New Zealand using a climatological rainfall sur- Vapour pressure deficit is a critical climate control on GPP and the light re- face. Int. J. Climatol. 26, 2097–2115. sponse of NEE at a southern hemisphere bog. Agric. For. Meteorol. 203, 54–63. Tait, A., Sturman, J., Clark, M., 2012. An assessment of the accuracy of interpolated Helliker, B.R., Richter, S.L., 2008. Subtropical to boreal convergence of tree-leaf tem- daily rainfall for New Zealand. J. Hydrol. (N.Z.) 51, 25–44. peratures. Nature 454, 511–515. Ummenhofer, C.C., England, M.H., 2007. Interannual extremes in New Zealand pre- Hong, Y.T., Jiang, H.B., Liu, T.S., Zhou, L.P., Beer, J., Li, H.D., Leng, X.T., Hong, B., Qin, cipitation linked to modes of Southern Hemisphere climate variability. J. Cli- X.G., 2000. Response of climate to solar forcing recorded in a 6000-year δ18O mate 20, 5418–5440. time-series of Chinese peat cellulose. Holocene 10, 1–7. Vandergoes, M.J., Newnham, R.M., Preusser, F., Hendy, C.H., Lowell, T.V., Fitzsimons, Hong, Y.T., Hong, B., Lin, Q.H., Shibata, Y., Zhu, Y.X., Leng, X.T., Wang, Y., 2009. Syn- S.J., Hogg, A.G., Kasper, H.U., Schluechter, C., 2005. Regional insolation forcing chronous climate anomalies in the western North Pacific and North Atlantic of Late Quaternary climate change in the Southern Hemisphere. Nature 436, regions during the last 14 000 years. Quat. Sci. Rev. 28, 840–849. 242–245. IAEA/WMO, 2014. Global network of isotopes in precipitation. The GNIP database. Wagstaff, S., Clarkson, B., 2012. Systematics and ecology of the Australasian genus Accessible at: http://www.iaea.org/water. Empodisma (Restionaceae) and description of a new species from peatlands in Leng, M.J., 2004. Isotopes in quaternary palaeoenvironmental reconstruction northern New Zealand. PhytoKeys 13, 39–79. (ISOPAL). Quat. Sci. Rev. 23, 739–741. Loader, N.J., Robertson, I., Barker, A.C., Switsur, V.R., Waterhouse, J.S., 1997. An im- Waterhouse, J.S., Switsur, V.R., Barker, A.C., Carter, A.H.C., Robertson, I., 2002. Oxygen proved technique for the batch processing of small wholewood samples to and hydrogen isotope ratios in tree rings: how well do models predict observed α-cellulose. Chem. Geol. 136, 313–317. values? Earth Planet. Sci. Lett. 201, 421–430. Loader, N.J., Street-Perrott, F.A., Daley, T.J., Hughes, P.D.M., Kimak, A., Levanic,ˇ T., Mal- West, A.G., Patrickson, S.J., Ehleringer, J.R., 2006. Water extraction times for plant lon, G., Mauquoy, D., Robertson, I., Roland, T.P., van Bellen, S., Ziehmer, M.M., and soil materials used in stable isotope analysis. Rapid Commun. Mass Spec- Leuenberger, M., 2014. Simultaneous determination of stable carbon, oxygen and trom. 20, 1317–1321. hydrogen isotopes in cellulose. Anal. Chem. 87, 376–380. Yakir, D., DeNiro, M.J., 1990. Oxygen and hydrogen isotope fractionation during cel- Ménot-Combes, G., Burns, S.J., Leuenberger, M., 2002. Variations of 18O/16O in plants lulose metabolism in Lemna Gibba L. Plant Physiol. 93, 325–332. from temperate peat bogs (Switzerland): implications for paleoclimatic studies. Yu, S.-Y., 2013. Quantitative reconstruction of mid- to late-Holocene climate in NE Earth Planet. Sci. Lett. 202, 419–434. China from peat cellulose stable oxygen and carbon isotope records and mech- Newnham, R.M., Lowe, D.J., 2000. Fine-resolution pollen record of late-glacial cli- anistic models. Holocene 23, 1507–1516. mate reversal from New Zealand. Geology 28, 759–762. Yu, S.-Y., Kang, Z., Zhou, W., 2011. Quantitative palaeoclimate reconstruction as an Newnham, R.M., de Lange, P.J., Lowe, D.J., 1995. Holocene vegetation, climate and inverse problem: a Bayesian inference of late-Holocene climate on the eastern history of a raised bog complex, northern New Zealand based on palynology, Tibetan Plateau from a peat cellulose δ18O record. Holocene 22, 405–412. plant macrofossils and tephrochronology. Holocene 5, 267–282. Zanazzi, A., Mora, G., 2005. Paleoclimatic implications of the relationship between Nichols, J., Booth, R.K., Jackson, S.T., Pendall, E.G., Huang, Y., 2010. Differential oxygen isotope ratios of moss cellulose and source water in wetlands of lake hydrogen isotopic ratios of Sphagnum and vascular plant biomarkers in om- superior. Chem. Geol. 222, 281–291.