• Aucun résultat trouvé

Terrestrial nitrogen feedbacks may accelerate future climate change

N/A
N/A
Protected

Academic year: 2021

Partager "Terrestrial nitrogen feedbacks may accelerate future climate change"

Copied!
6
0
0

Texte intégral

(1)

HAL Id: hal-03201202

https://hal.archives-ouvertes.fr/hal-03201202

Submitted on 18 Apr 2021

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Terrestrial nitrogen feedbacks may accelerate future

climate change

Sönke Zaehle, Pierre Friedlingstein, Andrew Friend

To cite this version:

Sönke Zaehle, Pierre Friedlingstein, Andrew Friend. Terrestrial nitrogen feedbacks may accelerate future climate change. Geophysical Research Letters, American Geophysical Union, 2010, 37 (1), pp.n/a-n/a. �10.1029/2009GL041345�. �hal-03201202�

(2)

Terrestrial nitrogen feedbacks may accelerate future climate change

So¨nke Zaehle,1 Pierre Friedlingstein,2,3 and Andrew D. Friend4

Received 13 October 2009; revised 25 November 2009; accepted 10 December 2009; published 5 January 2010.

[1] The effects of nitrogen (N) constraints on future terrestrial carbon (C) dynamics are investigated using the O-CN land surface model. The model’s responses to elevated [CO2] and soil warming agree well with observations made in ecosystem manipulation studies. N dynamics reduce terrestrial C storage due to CO2fertilization over the period 1860 – 2100 by 50% (342 Pg C) mainly in mid-high latitude ecosystems, compared to a simulation not accounting for N dynamics. Conversely, N dynamics reduce projected losses of land C due to increasing temperature by 16% (49 Pg C); however, this effect is prevalent only in mid-high latitude ecosystems. Despite synergistic interactions, the balance of these opposing effects is a significant reduction in future net land C storage. Terrestrial N dynamics thereby consistently increase atmospheric [CO2] in the year 2100 with a median value of 48 (41 – 55) ppmv, corresponding to an additional radiative forcing of 0.29 (0.28 – 0.34) W m2.

Citation: Zaehle, S., P. Friedlingstein, and A. D. Friend (2010), Terrestrial nitrogen feedbacks may accelerate future climate change, Geophys. Res. Lett., 37, L01401, doi:10.1029/ 2009GL041345.

1. Introduction

[2] While there is good evidence that terrestrial eco-system productivity and C storage are responding posi-tively to increasing atmospheric [CO2] [Norby et al., 2005], it is expected that N feedbacks will strongly constrain such behavior in the future [Hungate et al., 2003; Luo et al., 2004]. Conversely, increased soil organic matter decomposition resulting from higher soil temperatures could, while reducing soil C through increased respiration, increase net N mineralization and thereby enhance future plant productivity and C storage [Melillo et al., 2002]. Predicting the relative importance of these responses at the global scale is important as the outcome will strongly impact future terrestrial C storage; and hence climate. Two recent global modeling studies have suggested that N dynamics significantly alter the future behavior of the terrestrial C cycle, reversing the terrestrial carbon-climate interaction from strongly positive to weakly negative, with

important consequences for future climate [Sokolov et al., 2008; Thornton et al., 2009].

2. Methods

[3] We first use observations from ecosystem manipula-tion studies to evaluate the sensitivity of C dynamics to N constraints as simulated by the O-CN land-surface model [Zaehle and Friend, 2010], a component of the Institut Pierre Simon Laplace’s (IPSL) climate model [Marti et al., 2005]. We then use this model to assess the significance of terrestrial N dynamics on the interactions between the global C cycle and climate. O-CN has been developed from ORCHIDEE [Krinner et al., 2005], and combines a com-prehensive treatment of C, water, and energy fluxes with current understanding of N effects on plant and soil pro-cesses, including a mechanistic simulation of N losses to leaching and trace gas emissions. Zaehle et al. [2010] found that O-CN simulates historical C and N fluxes and stocks that are consistent with current understanding, in particular those patterns associated with seasonal and interannual variability in net land-atmosphere C exchanges. We present results from two sets of simulations, one explicitly account-ing for N dynamics (‘‘O-CN’’), and a second usaccount-ing the same model but with N dynamics switched off (‘‘O-C’’) to isolate the marginal effects of N dynamics. Supplementary material is available describing details of the modeling procedure.

3. Results

3.1. Model Sensitivity

[4] The realism of the sensitivity of O-CN to anticipated changes in atmospheric [CO2] was assessed by comparing site-specific simulations with observations from the Duke free air CO2enrichment experiment [Oren et al., 2001]. The model agrees well with both observed productivity in the control plots and the observed enhancement of net primary production (NPP; Figure 1a) due to increased [CO2] [Norby et al., 2005]. The observed increase of N stress with time, exhibited by a strong sensitivity of the NPP CO2response to N addition, falling foliar N content, and increasing below-ground allocation [Palmroth et al., 2006] is also reproduced by the model (see Figure S1).5

[5] Model responses to soil warming were tested by comparing predictions with the results of the Harvard Forest soil warming experiment [Melillo et al., 2002]. Simulated enhancement of soil respiration and cumulative mineraliza-tion of C (950 g C m2) and N (40 g N m2) over 10 years of soil warming agree well with the observations. Previous

5Auxiliary materials are available in the HTML. doi:10.1029/

2009GL041345. Here for Full Article 1

Max Planck Institute for Biogeochemistry, Jena, Germany.

2QUEST, Department of Earth Sciences, University of Bristol, Bristol,

UK.

3

LSCE, UMR, CEA, CNRS, Gif-sur-Yvette, France.

4

Department of Geography, University of Cambridge, Cambridge, UK. Copyright 2010 by the American Geophysical Union.

0094-8276/10/2009GL041345$05.00

(3)

work has shown that O-CN’s predictions of the fate of increases in mineral N supply, and the response of net C storage to chronic increases in N inputs, are in close agreement with monitoring and manipulative studies [Zaehle and Friend, 2010]. O-CN reproduces accurately the observed fraction of mineralized N transferred to woody biomass, and the observed stimulation of vegetation growth. The latter compensates for the enhanced soil organic matter decomposition [Magill et al., 2004] (Figures 1b and S2). This important response cannot be captured unless terres-trial N dynamics and their effects are considered.

[6] Despite these encouraging model-data comparisons we acknowledge that the set of available ecosystem manip-ulation experiments is limited with respect to spatial and ecosystem coverage as well as duration. We therefore apply our model to the global scale with important caveats, while nevertheless noting that if these warming and CO2 experi-ments are representative of larger scale responses, then so is our model. A global sensitivity experiment assessing the short-term response of O-CN to increased atmospheric [CO2] and soil warming suggests that N dynamics play a key role in boreal, and a strong role in temperate ecosys-tems, whereas tropical ecosystems show a rather low sensitivity to N dynamics (see supplementary material). These latitudinal differences in the role of N dynamics for C exchange are entirely consistent with ecological under-standing of the spatial distribution of nutrient controls on primary production [Vitousek and Howarth, 1991]. 3.2. Global Simulations

[7] To assess the effects of N dynamics on global land C dynamics, we undertook factorial global transient simula-tions driven by increasing atmospheric [CO2] [Nakicenovic et al., 2000], N deposition [Dentener et al., 2006], and by changing climate. Climate forcing used anomalies for 1860 to 2100 obtained from IPSL-CM4 [Marti et al., 2005] under the A2 emission scenario [Nakicenovic et al., 2000] applied to a 1961 – 1990 observational climatology [Mitchell et al., 2004].

Figure 1. Observed (OBS) and simulated responses from the O-CN and O-C model versions of (a) net primary production (NPP) to atmospheric [CO2] enrichment and (b) cumulative net C uptake to soil warming. The response to elevated [CO2] is the mean enhancement (in percent) of annual NPP in elevated (566 ppmv) relative to ambient (367 ppmv) plots during 6 years of enrichment (1997 – 2002) at Duke Forest (35°580N, 79°050W, predominant vegetation type: temperate needleleaved evergreen forest). Observations are taken from Norby et al. [2005]. Vertical lines denote between-year standard deviation, including the variance between replicate plots. The total ecosystem C accumulation [kg C m2] in response to a soil warming of 5 K is expressed as the difference between the perturbed and control plots at Harvard Forest (42°300N, 72°W, predominant vegetation type: temperate broadleaved deciduous forest) after 10 years of treatment. Observations are taken from Melillo et al. [2002] and Magill et al. [2004]. See auxiliary material for details.

Figure 2. (a) Cumulative global land C storage from 1860 with (red; O-CN) and without (blue; O-C) accounting for terrestrial N dynamics and (b) effect of N dynamics on cumulative global land C storage by 2100 (expressed as O-CN minus O-C), simulated in response to historical and projected future changes in atmospheric [CO2] only (CO2), climate change only (CLIM), N deposition only (NDEP), and all three factors combined (ALL). ‘‘SYN’’ refers to the additional C storage in land due to the synergistic interactions of individual forcing factors [Zaehle et al., 2010]. See auxiliary material for details.

L01401 ZAEHLE ET AL.: CARBON-NITROGEN FEEDBACKS AFFECT CLIMATE L01401

(4)

[8] Predictions by both O-CN and O-C are equally compatible with current estimates of global terrestrial C stocks and fluxes (Table S1), with the terrestrial biosphere predicted to have accumulated 83 Pg C (O-CN) or 108 Pg C (O-C) between 1860 and 2000. This difference in C sequestration would have led to a difference in atmospheric [CO2] of 6 ppmv, based on a three-box global C cycle model incorporating the responses of the IPSL coupled carbon-climate model [Friedlingstein et al., 2006]. These results imply land sink fractions of total anthropogenic C emissions of 0.26 (O-C) and 0.20 (O-CN) for the period 1960 – 2000, both compatible with current understanding of the dynamics of the contemporary global C cycle (0.20 – 0.27) [Raupach et al., 2008].

[9] N dynamics, expressed as O-CN minus O-C, cause projected additional land C accumulation between 1860 and 2100 to fall by 37%, from 388 Pg C to 244 Pg C (Figure 2). The simulations using CO2 forcing alone show that N availability strongly limits terrestrial C accumulation in response to rising CO2 concentrations (so-called ‘‘CO2 fertilization’’) in northern boreal and temperate ecosystems, but has little effect within the tropics (Figure S4). Global terrestrial C accumulation by 2100 is 50% (342 Pg C) lower than when not explicitly accounting for N dynamics (Figure 2), resulting in a reduction of the land CO2 sensitivity parameter bl from 1.27 to 0.63 Pg C ppmv1 (Figure 3a). The simulations with only climate change forcing show that the stimulation of plant growth due to warming induced soil net N mineralization reduces temper-ate and boreal C losses in response to climtemper-ate change by

16% (49 Pg C) (Figures 2 and S4). However, this reduction is not sufficient to compensate for the heat-stress and drought related C losses from tropical ecosystems. These opposing effects of temperature increases in different regions due to N dynamics result in a small global reduction in the magnitude of the land temperature sensitivity (gl) from 61 to 51 Pg C K1 (Figure 3b). Increasing N deposition, although important in regions of high deposition rates, has only minor impacts at the global scale (Figure 2), leading to a comparatively small increase in net land C uptake (27 Pg C), mostly in the northern extra-tropics, in agreement with previous findings [Townsend et al., 1996; Nadelhoffer et al., 1999].

[10] The simulated effect of N dynamics on net land C storage in temperate and boreal ecosystems is consistent with overwhelming evidence that N availability strongly controls terrestrial C cycling in temperate and boreal eco-systems [Vitousek and Howarth, 1991], limiting their response to elevated [CO2] [Hyvonen et al., 2007]. N dynamics increase simulated C storage in boreal ecosystems due to warming induced net N mineralization as suggested by Jarvis and Linder [2000]; however, N availability also limits net C gains in Arctic regions expected from increased growing season length and air temperature, as observed by Shaver et al. [1998]. O-CN results qualitatively agree with those of Sokolov et al. [2008] and Thornton et al. [2009] in that the strength of CO2 fertilization is reduced and that warming-induced N mineralization increases vegetation growth, but the three studies differ in the relative impor-Figure 3. Sensitivity of projected net land-atmosphere C flux by O-CN (red) or O-C (blue) in response to (a) atmospheric [CO2] and (b) climate changes relative to the respective simulated 1860 conditions. (c) Projected atmospheric [CO2] taking only the CO2fertilization effect on land and ocean net C fluxes into account using a 3-box model. (d) Increase in projected atmospheric [CO2] relative to Figure 3c as a result of additionally accounting for the sensitivity of the net land-ocean-atmosphere C exchanges to climatic changes (without synergistic effects). Open diamonds (Figures 3a – 3b)) show sensitivities obtained with coupled carbon-cycle climate models [Friedlingstein et al., 2006] that do not account for N dynamics. Solid circles are results from the CLM model [Thornton et al., 2009] and solid triangles from the IGSM model [Sokolov et al., 2008] with (red) and without (blue) explicitly accounting for N dynamics. Land C flux sensitivities (Figures 3a – 3b) are derived from the simulated cumulative land-atmosphere C flux in response to changes in either atmospheric [CO2] or climate (expressed as global mean surface temperature) from 1860 (see text). Projected atmospheric [CO2] (Figures 3c – 3d) are based on these sensitivities combined with the net atmosphere-ocean C flux sensitivities of the IPSL ocean model [Friedlingstein et al., 2006]. See auxiliary material for details.

(5)

tance of the two effects as the strength of the N constraint varies geographically.

[11] Available evidence suggests that the productivity of tropical ecosystems is limited more by the availability of phosphorus (P) than of N [Davidson et al., 2004; Cleveland and Townsend, 2006]; the lower importance of N dynamics for tropical C exchange predicted by O-CN appears there-fore reasonable. However, this finding is counter to the results of Sokolov et al. [2008], who show a strong attenuation of the tropical C cycle response due to N dynamics, and Thorton et al. [2009], who found a large increase in tropical C storage in response to warming due to reduced N limitation. Our simulated increase in productivity and widening of foliar C:N ratios in response to elevated [CO2] have been observed in tropical rain forests [Lloyd and Farquhar, 2008], suggesting that elevated [CO2] may affect plant growth, although the evidence is inconclusive. Obser-vational evidence suggests a negative effect of temperature increases on vegetation growth in tropical forests [Clark, 2004], exacerbated by increasing drought stress related to positive phases of ENSO [Phillips et al., 2009]. This evidence supports the trend towards reduced tropical bio-mass under moderate warming predicted by both O-C and O-CN.

[12] What do our results imply for future atmospheric [CO2] and climate? N dynamics induce a major constraint on the response of land C uptake to atmospheric CO2, coupled with a small increase in land carbon uptake due to increased temperature. This N constraint on CO2 fertiliza-tion would lead to an addifertiliza-tional 71 – 99 ppmv of CO2 remaining in the atmosphere by 2100 (Figure 3c), depend-ing on the assumed ocean sensitivity to atmospheric [CO2] [Friedlingstein et al., 2006]. The combined effect of N dynamics on both land CO2 (bl) and temperature sensitiv-ities (gl), but without allowing for any synergistic interac-tions between them, results in an additional increase in atmospheric CO2of 36 – 111 ppmv (relative to 28 – 85 ppmv for O-C), depending on the assumed ocean carbon and atmosphere sensitivity [Friedlingstein et al., 2006]. This increase is due to the effect of reduced CO2‘‘fertilization’’ outweighing the reduction in carbon loss from climate change. Taking account of the complete set of forcings (i.e., atmospheric [CO2], climate, and N deposition) and their synergistic interactions (Figure 2b) [see also Zaehle et al., 2010], the results of O-CN imply that atmospheric [CO2] will be in the range 834 – 1014 (median: 916) ppmv by the year 2100, whereas the C cycle-only model (O-C) implies an atmospheric concentration of 793 – 960 (median: 868) ppmv. Terrestrial N dynamics are therefore predicted to increase atmospheric [CO2] in year 2100 by 41 – 54 (medi-an: +48) ppmv, with the range of values reflecting the different transient linear climate and ocean-atmosphere C flux sensitivities across the C4MIP coupled carbon-cycle climate models [Friedlingstein et al., 2006]. This additional amount of CO2in the atmosphere due to N feedbacks would cause an additional radiative forcing of 0.28 – 0.34 (median: 0.29) Wm2, corresponding to 0.18 – 0.37 (median: 0.31) K of additional mean global surface warming. While acknowl-edging that other important factors in the global C cycle, such as land use or P dynamics, would likely impact on the response of the terrestrial biosphere to N dynamics, the study presented here represents a key step towards

realisti-cally accounting for, and so better understanding, the effects of N dynamics on C cycle-climate feedbacks.

4. Conclusions

[13] These results demonstrate that modeling studies not accounting for terrestrial N dynamics [Friedlingstein et al., 2006] have likely substantially overestimated future C storage in terrestrial ecosystems in response to elevated atmospheric CO2, and thus underestimated C accumulation in the atmosphere. Contrary to previous studies [Sokolov et al., 2008; Thornton et al., 2009], we find that N dynamics have little effect on tropical C dynamics and that the stimulation of plant growth in N-limited ecosystems due to future warming of soils is only prevalent in mid- to high-latitude ecosystems. This effect is too small to preclude the existence of a positive feedback between the global C cycle and the climate system by the end of this century. Our model simulations reported here suggest that N dynamics consistently accelerate C accumulation in the atmosphere and thereby imply greater rates of climate change than estimated previously.

[14] Acknowledgments. We thank I. C. Prentice for useful discus-sions. This work was supported by the Marie Curie RTN GREENCYCLES (MRTN-CT-2004-512464) and ERG JULIA (PERG02-GA-2007-224775). Computing time was provided by the Centre d’Energie Atomique, Saclay, France.

References

Clark, D. A. (2004), Sources or sinks? The responses of tropical forests to current and future climate and atmospheric composition, Philos. Trans. R. Soc. London, Ser. B, 359, 477 – 491, doi:10.1098/rstb.2003.1426. Cleveland, C., and A. Townsend (2006), Nutrient additions to a tropical rain

forest drive substantial soil carbon dioxide losses to the atmosphere, Proc. Natl. Acad. Sci. U. S. A., 103, 10,316 – 10,321, doi:10.1073/ pnas.0600989103.

Davidson, E. A., et al. (2004), Nutrient limitation of biomass growth in a tropical secondary forest: early results of a nitrogen and phosphorus amendment experiment, Ecol. Appl., 14, suppl., 150 – 163, doi:10.1890/ 01-6006.

Dentener, F., et al. (2006), Nitrogen and sulfur deposition on regional and global scales: A multimodel evaluation, Global Biogeochem. Cycles, 20, GB4003, doi:10.1029/2005GB002672.

Friedlingstein, P., et al. (2006), Climate-carbon cycle feedback analysis: Results from the C4MIP model intercomparison, J. Clim., 19, 3337 –

3353, doi:10.1175/JCLI3800.1.

Hungate, B. A., J. S. Dukes, M. R. Shaw, Y. Luo, and C. B. Field (2003), Nitrogen and climate change, Science, 302, 1512 – 1513, doi:10.1126/ science.1091390.

Hyvonen, R., et al. (2007), The likely impact of elevated [CO2], nitrogen

deposition, increased temperature and management on carbon sequestra-tion in temperate and boreal forest ecosystems: A literature review, New Phytol., 173, 463 – 480, doi:10.1111/j.1469-8137.2007.01967.x. Jarvis, P., and S. Linder (2000), Botany: Constraints to growth of boreal

forests, Nature, 405, 904 – 905, doi:10.1038/35016154.

Krinner, G., N. Viovy, N. de Noblet-Ducoudre´, J. Oge´e, J. Polcher, P. Friedlingstein, P. Ciais, S. Sitch, and I. C. Prentice (2005), A dynamic global vegetation model for studies of the coupled atmosphere-biosphere system, Global Biogeochem. Cycles, 19, GB1015, doi:10.1029/ 2003GB002199.

Lloyd, J., and G. D. Farquhar (2008), Effects of rising temperatures and [CO2] on the physiology of tropical forest trees, Philos. Trans. R. Soc. B,

363, 1811 – 1817, doi:10.1098/rstb.2007.0032.

Luo, Y., et al. (2004), Progressive nitrogen limitation of ecosystem re-sponses to rising atmospheric carbon dioxide, BioScience, 54, 731 – 739, doi:10.1641/0006-3568(2004)054[0731:PNLOER]2.0.CO;2. Magill, A., J. Aber, W. Currie, K. Nadelhoffer, M. Martin, W. McDowell,

J. Melillo, and P. Steudler (2004), Ecosystem response to 15 years of chronic nitrogen additions at the Harvard Forest LTER, Massachusetts, USA, For. Ecol. Manage., 196, 7 – 28, doi:10.1016/j.foreco.2004.03.033. Marti, O., et al. (2005), The New IPSL Climate System Model: IPSL-CM4,

84 pp., Inst. Pierre-Simon Laplace, Paris, France.

L01401 ZAEHLE ET AL.: CARBON-NITROGEN FEEDBACKS AFFECT CLIMATE L01401

(6)

Melillo, J. M., et al. (2002), Soil warming and carbon-cycle feedbacks to the climate system, Science, 298, 2173 – 2176, doi:10.1126/ science.1074153.

Mitchell, T. D., T. R. Carter, P. D. Jones, M. Hulme, and M. New (2004), A comprehensive set of high-resolution grids of monthly climate for Europe and the globe: the observed record (1901 – 2000) and 16 scenarios (2001 – 2100), Tyndall Cent. Working Pap. 55, 33 pp., Univ. of East Anglia, Norwich, U. K.

Nadelhoffer, K., B. Emmet, P. Gunderson, O. Kjonaas, C. Koopmans, P. Schleppi, A. Tietma, and R. Wright (1999), Nitrogen deposition makes a minor contribution to carbon sequestration in temperate forests, Nature, 398, 145 – 148, doi:10.1038/18205.

Nakicenovic, N., J. Alcamo, G. Davis, B. de Vries, and J. Fenhann (2000), Special Report on Emissions Scenarios: A Special Report of Working Group III of the Intergovernmental Panel on Climate Change, Cambridge Univ. Press, Cambridge, U. K.

Norby, R. J., et al. (2005), Forest response to elevated CO2is conserved

across a broad range of productivity, Proc. Natl. Acad. Sci. U. S. A., 102, 18,052 – 18,056, doi:10.1073/pnas.0509478102.

Oren, R., et al. (2001), Soil fertility limits carbon sequestration by forest ecosystems in a CO2-enriched atmosphere, Nature, 411, 469 – 472,

doi:10.1038/35078064.

Palmroth, S., R. Oren, H. R. McCarthy, K. H. Johnsen, A. C. Finzi, J. R. Butnor, M. G. Ryan, and W. H. Schlesinger (2006), Aboveground sink strength in forests controls the allocation of carbon below ground and its [CO2]-induced enhancement, Proc. Natl. Acad. Sci. U. S. A., 103,

19,362 – 19,367, doi:10.1073/pnas.0609492103.

Phillips, O. L., et al. (2009), Drought sensitivity of the Amazon rainforest, Science, 323, 1344 – 1347, doi:10.1126/science.1164033.

Raupach, M. R., J. Canadell, and C. Le Que´re´ (2008), Anthropogenic and biophysical contributions to increasing atmospheric CO2growth rate and

airborne fraction, Biogeosciences, 5, 1601 – 1613.

Shaver, G. R., L. C. Johnson, D. H. Cades, G. Murray, J. A. Laundre, E. B. Rastetter, K. J. Nadelhoffer, and A. E. Giblin (1998), Biomass and CO2

flux in wet sedge tundras: Responses to nutrients, temperature, and light, Ecol. Monogr., 68, 75 – 97.

Sokolov, A. P., D. W. Kicklighter, J. M. Melillo, B. S. Felzer, C. A. Schlosser, and T. W. Cronin (2008), Consequences of considering car-bon-nitrogen interactions on the feedbacks between climate and the ter-restrial carbon cycle, J. Clim., 21, 3776 – 3796, doi:10.1175/ 2008JCLI2038.1.

Thornton, P. E., et al. (2009), Carbon-nitrogen interactions regulate climate-carbon cycle feedbacks: Results from an atmosphere-ocean general cir-culation model, Biogeosciences, 6, 2099 – 2120.

Townsend, A., B. H. Braswell, E. Holland, and J. Penner (1996), Spatial and temporal patterns in potential terrestrial carbon storage due to deposi-tion of fossil fuel derived nitrogen, Ecol. Appl., 6, 806 – 814, doi:10.2307/ 2269486.

Vitousek, P. M., and R. W. Howarth (1991), Nitrogen limitation on land and in the sea: How it can occur, Biogeochemistry, 13, 87 – 115, doi:10.1007/ BF00002772.

Zaehle, S., and A. D. Friend (2010), Carbon and nitrogen cycle dynamics in the O-CN land surface model: 1. Model description, site-scale evaluation, and sensitivity to parameter estimates, Global Biogeochem. Cycles, doi:10.1029/2009GB003521, in press.

Zaehle, S., A. D. Friend, F. Dentener, P. Friedlingstein, P. Peylin, and M. Schulz (2010), Carbon and nitrogen cycle dynamics in the O-CN land surface model: 2. The role of the nitrogen cycle in the historical terrestrial carbon balance, Global Biogeochem. Cycles, doi:10.1029/ 2009GB003522, in press.



P. Friedlingstein, QUEST, Department of Earth Sciences, University of Bristol, Queens Road, Bristol BS8 1RJ, UK.

A. D. Friend, Department of Geography, University of Cambridge, Downing Place, Cambridge CB2 3EN, UK.

S. Zaehle, Max Planck Institute for Biogeochemistry, Hans-Kno¨ll-Str. 10, D-07745 Jena, Germany. (soenke.zaehle@bgc-jena.mpg.de)

Figure

Figure 1. Observed (OBS) and simulated responses from the O-CN and O-C model versions of (a) net primary production (NPP) to atmospheric [CO 2 ] enrichment and (b) cumulative net C uptake to soil warming

Références

Documents relatifs

Dans cette hypothèse (Figure IV-1.3), la surface d'érosion à la base de l'Oligocène se serait mise en place en régime subaérien sur la plate-forme et jusqu'en pied de pente à la

As the fourth industrial revolution accelerates, it is crucial to establish an innovative strategy for Government that utilizes these technologies. Notably, to keep

Figure 5.75: Mean energy expenditure over all AUVs versus swarm centroid travel distance while constructing and maintaining a square lattice formation in simulated Red Sea

Different LSLBI with different outcomes Independent farmer model Associative farmer model Cooperative farmer model Speculative 1000-day model Asset management and

Land Matrix Initiative Land Matrix Global Observatory LPI Observatory Madagascar … Thematic Rangelands Observatory … Land Observatories. • Each Observatory –

Large-scale land acquisitions as aggravator Drivers Food production Biofuels Industrial production Forest/fibre production Ecosystemic Services/Turisme Speculation Triggers. Food

6 CIBSS - Centre for Integra4ve Biological Signalling Studies, University of Freiburg 7 Signalling Research Centres BIOSS and CIBSS, University of Freiburg!.

There have been obvious failures when brick was painted and several factors appear responsible: (1) older types of coatings such as oil paints did not perform well on bricks