<<

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ doi:10.5194/bg-13-341-2016 © Author(s) 2016. CC Attribution 3.0 License.

Multiple between plants, microbes, and surfaces: model development, parameterization, and example applications in several tropical forests

Q. Zhu, W. J. Riley, J. Tang, and C. D. Koven Climate Sciences Department, Earth Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA

Correspondence to: Q. Zhu ([email protected])

Received: 9 January 2015 – Published in Biogeosciences Discuss.: 5 March 2015 Revised: 18 December 2015 – Accepted: 21 December 2015 – Published: 18 January 2016

Abstract. Soil is a complex system where biotic (e.g., plant or calibration. Under soil inorganic nitrogen and phospho- roots, micro-organisms) and abiotic (e.g., mineral surfaces) rus elevated conditions, the model accurately replicated the consumers compete for resources necessary for life (e.g., ni- experimentally observed competition among nutrient con- trogen, phosphorus). This competition is ecologically signif- sumers. Although we used as many observations as we could icant, since it regulates the dynamics of soil and obtain, more nutrient addition experiments in tropical sys- controls aboveground plant . Here we develop, tems would greatly benefit model testing and calibration. In calibrate and test a nutrient competition model that accounts summary, the N-COM model provides an ecologically con- for multiple soil nutrients interacting with multiple biotic and sistent representation of nutrient competition appropriate for abiotic consumers. As applied here for tropical forests, the land BGC models integrated in Earth System Models. Nutrient COMpetition model (N-COM) includes three pri- + − mary soil nutrients (NH4 , NO3 and POx; representing the 3− 2− − sum of PO4 , HPO4 and H2PO4 ) and five potential com- petitors (plant roots, decomposing microbes, nitrifiers, deni- 1 Introduction trifiers and mineral surfaces). The competition is formulated with a quasi-steady-state chemical equilibrium approxima- Atmospheric CO2 concentrations have risen sharply since the tion to account for substrate (multiple substrates share one pre-industrial era, primarily due to anthropogenic ) and consumer (multiple consumers compete for combustion and land use and land cover change (Houghton, one substrate) effects. N-COM successfully reproduced ob- 2003; Le Quéré et al., 2013; Marland et al., 2008). Terrestrial served soil heterotrophic respiration, N2O emissions, free mitigate the increasing atmospheric CO2 trend + by absorbing roughly a quarter of anthropogenic CO emis- phosphorus, sorbed phosphorus and NH4 pools at a tropical 2 forest site (Tapajos). The overall model uncertainty was mod- sions (Le Quéré et al., 2009). However, it is still an open erately well constrained. Our sensitivity analysis revealed question whether the terrestrial CO2 sink can be sustained that soil nutrient competition was primarily regulated by (Sokolov et al., 2008; Zaehle et al., 2010), given that plant consumer–substrate affinity rather than environmental fac- productivity is generally limited by soil nutrients (Elser et al., tors such as soil temperature or soil moisture. Our results also 2007; LeBauer and Treseder, 2008; Vitousek and Howarth, imply that under strong nutrient limitation, relative competi- 1991) and soil nutrients could be quickly depleted through tiveness depends strongly on the competitor functional traits biogeochemical (Chauhan et al., 1981; Nordin et al., 2001; (affinity and nutrient carrier ). We then ap- Shen et al., 2011) and hydrological (Dise and Wright, 1995; plied the N-COM model to analyze field nitrogen and phos- Perakis and Hedin, 2002) processes. Therefore, a holistic phorus perturbation experiments in two tropical forest sites representation of soil nutrient dynamics is critically impor- (in Hawaii and Puerto Rico) not used in model development tant to model the responses of terrestrial CO2 up- take to climate change.

Published by Copernicus Publications on behalf of the European Geosciences Union. 342 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Until recently, land models integrated in Earth system tain sufficient nutrients to sustain their biological processes models (ESMs) have largely ignored the close coupling be- (e.g., , respiration). Therefore, it is critical to tween soil nutrient dynamics and the , although improve the representation of nutrient competition to accu- the impacts of soil nutrients (primarily nitrogen and phos- rately model how terrestrial ecosystems will respond to per- phorus) regulating carbon–climate feedback are clearly re- turbations in soil nutrient dynamics (e.g., from elevated nitro- quired in ecosystem and land models (Za- gen deposition or CO2 fertilization-induced nutrient require- ehle and Dalmonech, 2011; Zhang et al., 2011). For example, ments). none of the land models in C4MIP (Coupled Climate Carbon Intense competition between plants and Cycle Model Intercomparison Project phase 4) had coupled is a well-observed phenomenon in nutrient-limited systems carbon and nitrogen dynamics (Friedlingstein et al., 2006). (Hodge et al., 2000a; Johnson, 1992; Kaye and Hart, 1997). The current generation of CMIP5 (Anav et al., 2013) models Previously, plants were thought to be initial losers in nutrient used for the recent IPCC (Intergovernmental Panel on Cli- competition, due to the fact that microbes are more intimately mate Change) assessment had only two members (CLM4CN: associated with substrates (Woodmansee et al., 1981). How- Thornton et al., 2007; and BNU-ESM: Ji et al., 2014) that ever, increasing observational evidence indicates that plants considered nitrogen regulation of terrestrial carbon dynam- compete effectively with soil microorganisms (Schimel and ics. However, as discussed below, several recent studies have Bennett, 2004) under certain circumstances, sometimes even shown that these models had large biases in most of the in- outcompeting them and suppressing microbial growth (Hu et dividual processes important for simulating nutrient dynam- al., 2001; Wang and Lars, 1997). 15N isotope studies have ics. We therefore believe that, at the global scale, no credible also demonstrated that plants can capture a large fraction of representation of nutrient constraints on terrestrial carbon cy- added nitrogen (Hodge et al., 2000b; Marion et al., 1982). cling yet exists in ESMs. In the short term (days to months), plants maintain their Further, none of the CMIP5 ESMs included a phospho- competitiveness mainly through (1) establishing mycorrhizal rus cycle, which is likely important for tropical forest car- fungi associations (Drake et al., 2011; Rillig et al., 1998), bon budgets (Vitousek and Sanford, 1986). The recent IPCC which help plants acquire organic and inorganic forms of ni- report highlights the importance of nitrogen and phosphorus trogen (Hobbie and Hobbie, 2006; Hodge and Fitter, 2010) availability on land carbon storage, even though the phospho- and (2) root exudation of extracellular that decom- rus limitation effect is uncertain (Stocker et al., 2013). Since pose soil (Phillips et al., 2011). In the next generation of ESMs participating in the CMIP6 syn- the relatively longer term (months to years), morphological thesis will continue to focus on the impacts of a changing adjustment occurs; for example, plants allocate more carbon climate on terrestrial CO2 and abiotic exchanges with the at- to fine roots to explore laterally and deeper (Iversen et al., mosphere (Provides, 2014), developing ecologically realistic 2011; Jackson et al., 2009). Finally, over the course of years and observationally constrained representations of soil nu- to decades, plant succession can occur (Medvigy et al., 2009; trient dynamics and carbon–nutrient interactions in ESMs is Moorcroft et al., 2001) and the new plant demography will critical. need to be considered to represent nutrient controls on this The importance of nutrient limitations in terrestrial timescale. ecosystems has been widely demonstrated by nitrogen and Given these patterns from the observational literature, nu- phosphorus fertilization experiments (Elser et al., 2007). For trient competition is either absent or over-simplified in ex- instance, plant net (NPP) is enhanced in isting ESMs. One common representation of plant–microbe plots with nutrient addition (LeBauer and Treseder, 2008). competition is that plants compete poorly against microbes in Similarly, plant growth can be stimulated due to atmospheric acquisition. For example, the O-CN land model (Za- nitrogen deposition (Matson et al., 2002). Boreal forests ehle and Friend, 2010) assumes that soil-decomposing mi- are strongly limited by nitrogen availability (Vitousek and crobes have the priority to immobilize soil mineral nitrogen. Howarth, 1991), because low temperatures reduce nitrogen After microbes meet their demands, the remaining nitrogen mineralization (Bonan and Cleve, 1992) and N2 fixation is then available for plant uptake. (DeLuca et al., 2002, 2008). In contrast, tropical forests are Another treatment in ESM land models is that microbial often phosphorus limited (Vitousek et al., 2010), since trop- and plant nutrient acquisition competitiveness is based on ical are old and phosphorus derived from parent ma- their relative demands. For example, CLM4CN (Thornton terial weathering has been depleted through long-term pe- et al., 2007) assumes that the plant and microbial nitrogen dogenesis processes (Vitousek and Farrington, 1997; Walker demands are satisfied simultaneously. Under nitrogen infer- and Syers, 1976). In natural ecosystems without external nu- tile conditions, all nitrogen demands in the system are down- trients inputs (e.g., N deposition), soil nitrogen or phospho- regulated proportional to the individual demands and subject rus (or both) are likely insufficient to satisfy both plant and to available soil mineral nitrogen. This approach led to un- demands (Vitousek and Farrington, 1997). realistic diurnal cycles of gross primary production (GPP), Plants have to compete with microorganisms and mineral with midday depressions in GPP occurring because of pre- surfaces (Kaye and Hart, 1997; Schimel et al., 1989) to ob- dicted diurnal depletion of the soil mineral nitrogen pool.

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/

Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 343

- Plant NO3 uptake CO2 Plant POx uptake Plant C:N:P Plant NH4+ uptake

C P N2 fixation N Litter Litter Litter

Soil org N Soil org C Soil org P

N2O

immob +

NH4 deposition Dust P deposition immob +

4 N2O, N2 x Mineralization Mineralization PO NH + NH4 PP POx Weathering

POx Leaching NO3- deposition Mineral surface Sorb esorb D immob

- 3

NO - NO3 PO Occluded PS

NO3- Leaching

Figure 1. Model structure. Boxes represent pools, solid arrows represent aqueous fluxes and dashed arrows represent gaseous pathways out or into the system. Three essential chemical elements (carbon, C; nitrogen, N; and phosphorus, P) are simulated in N-COM (blue, red and green represent C, N and P pools and processes, respectively).

Emergent impacts of this conceptualization of nutrient con- We extend on that work here by presenting an implemen- straints on GPP resulted in poor predictions compared to ob- tation of ECA kinetics to represent competition for multi- servations, with smaller than observed plant C growth re- ple soil nutrients in a multiple consumer environment. We sponses to N deposition (Thomas et al., 2013a) and larger note that this paper demonstrates a method to handle instan- than observed responses to N fertilization (Thomas et al., taneous competition in the complex soil–plant network, but 2013b). Further, most biogeochemistry models not integrated a robust competition representation for climate-scale models in ESMs also adopt one of these approaches. For instance, will require representation of dynamic changes in plant allo- Biome-BGC (Running and Coughlan, 1988), CENTURY cation and plant composition. (Parton et al., 1988), CASA (Carnegie–Ames–Stanford Ap- The aim of this study is to provide a reliable nutrient com- proach; Potter et al., 1993) and the Terrestrial Ecosystem petition approach applicable for land models integrated in Model – TEM (McGuire et al., 1992) assume that available ESMs. However, before integration into an ESM, the com- nutrients preferentially satisfy the soil microbial immobiliza- petition model needs to be carefully calibrated and inde- tion demand. pendently tested against observational data. This paper will We believe the two conceptualizations of competition used therefore focus on model development and evaluation at sev- in ESMs substantially over-simplify competitive interactions eral tropical forest sites where observations are available. between plants and microbes and lead to biases in carbon cy- Our objectives are to (1) develop a soil biogeochemistry + − cle predictions. To begin to address the problems with these model with multiple nutrients (i.e., NH4 , NO3 and POx; rep- simplified approaches, Tang and Riley (2013) showed that 3− 2− − resented as the sum of PO4 , HPO4 and H2PO4 ) and mul- complex consumer–substrate networks can be represented tiple nutrient consumers (i.e., decomposing microbes, plants, with an approach – called equilibrium chemical approxima- nitrifiers, denitrifiers, and mineral surfaces) competition us- tion (ECA) kinetics – that simultaneously resolves multiple ing ECA kinetics (Tang and Riley, 2013; Zhu and Riley, demands for multiple substrates, and demonstrated that the 2015); (2) constrain the model with in situ observational data approach was consistent with observed litter sets of , nitrogen, and phosphorus dynamics us- observations. ECA kinetics has also recently been applied to ing a Markov chain Monte Carlo (MCMC) approach; and (3) analyze the emergent temperature response of soil organic test model performance against nitrogen and phosphorus fer- matter (SOM) decomposition, considering equilibrium, non- tilization studies. equilibrium and enzyme temperature sensitivities and abiotic interactions with mineral surfaces (Tang and Riley, 2014). www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 344 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

2 Method ously, soil organic N and P changes follow C decomposition:

N N N 2.1 Model development dNj X X X = −F dec + F move + F immob + F immob (5) dt N,j N,ij NH4,ij NO3,ij i=1 i=1 i=1 The Nutrient COMpetition model (N-COM) is designed as N N dPj X X a soil biogeochemistry model (Fig. 1) to simulate soil car- = −F dec + F move + F immob, (6) dt P,j P,ij P,ij bon decomposition, nitrogen and phosphorus transforma- i=1 i=1 tions, abiotic interactions and plant demands. Although our move move ultimate goal is to incorporate N-COM into a decomposi- where FN,ij and FP,ij are fluxes of nitrogen and phospho- tion model that represents active microbial activity as the pri- rus moving from the upstream (i) to downstream (j) pools. F immob, F immob and F immob are immobilization fluxes of soil mary driver of decomposition, we start here by presenting the NH4,ij NO3,ij P,ij dec dec N-COM approach using a Century-like (Koven et al., 2013; mineral nitrogen and phosphorus. FN,j and FP,j represent Parton et al., 1988) structure, with additions to account for decomposition losses. phosphorus dynamics. In our approach, we calculate poten- Equations (5) and (6) state that changes in the jth or- tial immobilization using literature-derived parameters (e.g., ganic N or P pool are the summation of three terms: (1) or- VMAX, KM ) in a Michaelis–Menten (MM) kinetics frame- ganic N and P lost during soil organic matter mineralization − dec − dec work. The potential immobilization is subsequently modified ( FN,j and FP,j ); (2) a fraction of the ith organic N or P using the ECA competition method. move pool (upstream) enters into the jth pool (downstream; FN,ij Five pools of soil organic carbon (C), nitrogen (N) and and F move); and (3) soil microbial immobilization (F immob, phosphorus (P) are considered: coarse wood debris (CWD), P,ij NH4,ij F immob and F immob). Immobilization occurs only when the litter, fast soil organic matter (SOM) pool, medium SOM NO3,ij P,ij pool, and slow SOM pool. Litter is further divided into three newly entering organic N is insufficient to sustain the soil C : N (or C : P) ratio (more details described in Appendix A). sub-groups: metabolic, and lignin. The soil organic + − dec dec dec The inorganic nitrogen pools (NH4 and NO3 ; Eqs. 7–8) C, N and P decomposition (FC,j , FN,j , FP,j ) follow first- order decay: are altered by production (organic N mobilized by microbes), consumption (uptake by plants and microbes, gaseous or dec aqueous losses) and transformation (nitrification and denitri- FC,j = kj Cj rθ rT (1) fication). Inorganic P (POx) is assumed to be either taken up dec FN,j = kj Nj rθ rT (2) by plants and decomposing microbes or adsorbed to mineral dec surfaces (Eq. 9). Plants utilize all forms of phosphate (e.g., FP,j = kj Pj rθ rT , (3) 3− 2− − PO4 , HPO4 , and H2PO4 ), but for simplicity we use the symbol POx to represent the sum of all possible phosphate where kj is the rate constant of soil organic matter decay −1 −2 forms throughout the paper: (s );Cj ,Nj and Pj are pool sizes (g m ) of carbon, ni- trogen and phosphorus, respectively (j from 1 to 7 repre- N N d[NH4] XX mob nit plant sents the soil organic matter pools: CWD, metabolic litter, = F − F − F dt NH4,ij NH4 NH4 cellulose litter, lignin litter, fast soil organic carbon (SOC), j=1 i=1 dep median SOC, slow SOC); rT and rθ (dimensionless) are soil − immob + BNF + FNH F FNH (7) temperature and moisture environmental regulators. 4 4 d[NO3] Decomposed carbon (F dec; upstream ith pool) either (1) = −F den + (1 − f N2O)F nit − F plant − F immob C,i dt NO3 NH4 NO3 NO3 enters a downstream pool (jth) or (2) is lost as CO2. Soil or- − F leach + F dep (8) ganic carbon (downstream jth pool) temporal change is cal- NO3 NO3 culated as d[PO ] N N x = XX mob − plant − immob FP,ij FP FP N dt dCj X j=1 i=1 = −F dec + F move, (4) dt C,j C,ij − surf − leach + weather i=1 FP FP F , (9) where F mob and F mob are gross mineralization rates for N NH4,ij P,ij P move where F is the summation of carbon fluxes that move nitrogen and phosphorus. F nit is the nitrification flux, part C,ij NH4 i=1 N O from the upstream pool (i) to the downstream pool (j) due to of which is lost through a gaseous pathway (f 2 ) and the rest is incorporated into the NO− pool. F den is the den- the decomposition of upstream SOC. For each upstream car- 3 NO3 = itrification flux, which transforms nitrate to N2O and N2 bon pool (i 1,2,...,7), the fractions integrated into down- + stream pools (j = 1,2,...,7) is summarized in a 7 × 7 matrix which then leave the soil system. Plant uptake of soil NH4 , f (Table 2). The percentage of decomposed carbon that is NO− and PO are represented as F plant, F plant and F plant, ij 3 x NH4 NO3 P respired as CO2 is represented by gi (Table 2). Simultane- respectively. Soil-decomposing microbial immobilization of

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 345

Table 1. A summary of the modeled consumer–resource competi- substrate competition reaction is represented by tion network. + + k1 k S + E C −→2 P + E. (13) Resources Consumers − k1 + NH Plant Decomposing microbe Nitrifier −4 The enzyme (E: e.g., nutrient carrier enzyme produced by NO Plant Decomposing microbe Denitrifier + − 3 plants and microbes) and substrate (S: e.g., NH , NO ) re- PO Plant Decomposing microbe Mineral surface 4 3 x action (reversible reaction) forms a substrate-enzyme com- plex (C). The following irreversible reaction leads to product + − immob immob (P: meaning the nutrients has been taken up) and releases soil NH , NO and POx are represented as F , F 4 3 NH4 NO3 enzyme (E) back into soil media. For the whole complex re- and F immob. The F leach and F leach are leaching losses of P NO3 P action network, nutrient uptakes are formulated as − soil NO and POx. External inputs into soil inorganic N plant 3 [NH ]·[E ] F plant = kplant · 4 N dep NH4 NH4 (3) (14) (1) (2) [ plant ] [ mic ] (4) [ nit ] (5) plant,NH4 [NH4 ] [NO3 ] EN EN EN pools include atmospheric ammonia deposition (F ), at- K (1 + plant,NH + plant,NO + plant,NH + mic,NH + nit,NH ) NH4 M K 4 K 3 K 4 K 4 K 4 dep M M M M M [NH ]·[Emic ] mospheric nitrate deposition (F ) and biological nitrogen F immob = kimmob · 4 N NO NH4 NH4 [ plant ] [ mic ] [ nit ] (15) 3 mic,NH4 [NH4 ] [NO3 ] EN EN EN K (1 + mic,NH + mic,NO + plant,NH + mic,NH + nit,NH ) BNF M K 4 K 3 K 4 K 4 K 4 fixation (F ). External sources of phosphate come from M M M M M nit [NH4 ]·[E ] weather F nit = knit · NH4 parent material weathering (F ). NH4 NH4 [ plant ] [ mic ] [ nit ] (16) nit,NH4 [NH4 ] EN ( EN EN K (1 + nit,NH + plant,NH + mic,NH + nit,NH ) M K 4 K 4 K 4 K 4 Finally, the dynamics of sorbed P (PS), occluded P (PO) M M M M [NO ]·[Eplant ] plant = plant · 3 N FNO kNO plant (17) and parent material P (PP) are modeled as 3 3 [E ] [Emic ] [Eden ] Kplant,NO3 (1 + [NH4 ] + [NO3 ] + N + N + N ) M plant,NH4 plant,NO3 plant,NO3 mic,NO3 den,NO3 KM KM KM KM KM [ ] mic d PS [NO3 ]·[E ] surf occl immob = immob · N FNO kNO plant (18) = F − F (10) 3 3 [E ] [Emic ] [Eden ] P P Kmic,NO3 (1 + [NH4 ] + [NO3 ] + N + N + N ) M mic,NH4 mic,NO3 plant,NO3 mic,NO3 den,NO3 dt KM KM KM KM KM [ ]·[ den ] d[P ] NO3 ENO O occl F den = kden · 3 = NO3 NO3 [ plant ] [ mic ] [ den ] (19) F (11) den,NO3 [NO3 ] EN EN EN P K (1 + den,NO + plant,NO + mic,NO + den,NO ) M K 3 K 3 K 3 K 3 dt M M M M [ ] plant d PP weather dep plant plant [POx ]·[E ] = − + F = k · P F FP , (12) P P plant (20) [E ] [Emic] [Esurf] dt plant,P + [POx ] + P + P + P KM (1 ) plant,P plant,P Kmic,P Ksurf,P where the pool of sorbed P is balanced by the adsorption flux KM KM M M mic surf occl [POx ]·[E ] (F ) and occlusion flux (F ). Parent material is lost by F mic = kmic · P P P P P plant (21) weather [E ] [Emic] [Esurf] weathering (F ) and is slowly replenished by external mic,P + [POx ] + P + P + P KM (1 ) dep Kmic,P plant,P Kmic,P Ksurf,P atmospheric phosphorus inputs (F , such as dust). More M KM M M P mic [POx ]·[E ] detailed information on the modeled C, N and P fluxes is F surf = ksurf · P P P plant , (22) [E ] [Emic] [Esurf] documented in Appendix A. surf,P + [POx ] + P + P + P KM (1 ) Ksurf,P plant,P Kmic,P Ksurf,P M KM M M 2.2 Multiple-consumer multiple-resource competition where the F represent the nutrient uptake fluxes and k is network the base reaction rate that enzyme–substrate complex forms product (k+ in Eq. 13). [E] and K denote enzyme abun- The soil biogeochemistry model presented in Sect. 2.1 has 2 M dance and half-saturation constants (substrate–enzyme affin- multiple potential nutrient consumers (plants, SOM decom- ity). Superscripts and subscripts refer to consumers and sub- posing microbes, nitrifiers, denitrifiers, mineral surfaces) and + − strates, respectively. These equations account for the effect multiple soil nutrients (NH , NO , PO ). The consumer– 4 3 x of (1) multiple substrates (e.g., NH+ and NO−) sharing one resource network is summarized in Table 1. As in many 4 3 consumer, which inhibits the effective binding between any land BGC models (CLM, Century, etc.), we have not ex- (1) (2) + specific substrate and the consumer (terms and in plicitly included the mineral surface adsorption of NH and 4 Eq. 14) and (2) multiple consumers (e.g., plants, decom- NO−, since we assume ammonia is quickly protected by 3 posing microbes and nitrifiers) sharing one substrate (e.g., mineral surfaces from leaching (no leaching term in Eq. 7) + NH ), which lowers the probability of effective binding be- but then released for plant and microbial uptake when the 4 tween any consumer and NH+ (terms (3), (4), and (5) in biotic demand arises. An improved treatment of these dy- 4 Eq. 14). namics would necessitate a prognostic model for pH, which For our reaction network (Eqs. 13–22), we make the is beyond the scope of this analysis. Unlike sorbed P (which following four assumptions: (1) Plant roots and decom- can be occluded), there is no further abiotic loss of sorbed posing microbes possess two types of nutrient carrier en- ammonia. Therefore, the free ammonia pool is interpreted in zymes (nutrient transporters). One is for nitrogen (NH+ and the current model structure as a potential free ammonia pool 4 − plant mic (free + sorbed). NO3 ; EN ,EN ), and the other is for phosphorus, includ- plant mic Competition between different consumers in acquiring dif- ing different forms of phosphate (EP ,EP ). (2) Nutri- ferent resources is summarized in Table 1. Each consumer– ent carrier enzyme abundance is scaled with (fine www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 346 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Table 2. Model parameters and baseline values.

C associated

gi Percentage of carbon remains in the – (1.0; 0.45; 0.5; 0.5; Koven et al. (2013) soil after decomposition of ith SOM 0.83; 0.45; 0.45) fij fraction of SOM leave from ith – (0, 0, 0.76, 0.24, 0, 0, 0; Koven et al. (2013) pool and enter into jth pool 0, 0, 0, 0, 1, 0, 0; 0, 0, 0, 0, 1, 0, 0; 0, 0, 0, 0, 0, 1, 0; 0, 0, 0, 0, 0, 0.995, 0.005; 0, 0, 0, 0, 0.93, 0, 0.07; 0, 0, 0, 0, 1, 0, 0) CN Soil organic matter CN ratio – (13,16,7.9) Parton et al. (1988) CP Soil organic matter CP ratio – (110,320,114) Parton et al. (1988) TURNSOM Soil organic matter turn over year (4.1, 0.066, 0.25, 0.25, Koven et al. (2013) (CWD, metabolic lit, cellulose lit, 0.17, 5, 270) lignin lit, fast SOM, medium SOM, slow SOM) N associated plant + k Reaction rate of plant NH carrier enzyme day−1 120a Jackson et al. (1997); NH4 4 Min et al. (2000) plant,NH4 + −2 KM Half-saturation constant for plant NH4 uptake g m 0.09 Kuzyakov and Xu (2013) Kmic,NH4 Half-saturation constant for g m−2 0.02 Kuzyakov and Xu (2013) M + decomposing microbe NH4 immobilization + −1 knit Maximum fraction of NH4 day 10 % Parton et al. (2001) pool that could be utilized by nitrifiers nit,NH4 + −2 KM Half-saturation constant for nitrifier NH4 consumption g m 0.076 Drtil et al. (1993) plant − k Reaction rate of plant NO carrier enzyme day−1 2a Jackson et al. (1997); NO3 3 Min et al. (2000) plant,NO3 − −2 KM Half-saturation constant for plant NO3 uptake g m 0.07 Kuzyakov and Xu (2013) − Kmic,NO3 Half-saturation constant for decomposing microbe NO g m−2 0.04 Kuzyakov and Xu (2013) M − 3 microbe NO3 immobilization den,NO3 − −2 KM Half-saturation constant for denitrifier NO3 consumption g m 0.011 Murray et al. (1989) [ plant] −2 · a EN Plant nitrogen carrier enzyme abundance for nitrogen uptake g m Cfroot 0.0000125 Tang and Riley (2013); Trumbore et al. (2006) b F immob,pot [ mic] −2 N EN Decomposing microbes nitrogen carrier g m 1000 Tang and Riley (2013) enzyme abundance for nitrogen immobilization [Enit] Nitrifier nitrogen carrier enzyme g m−2 1.2E−3 Raynaud et al. (2006) N + abundance for NH4 assimilation [Eden] Denitrifier nitrogen carrier enzyme g m−2 1.2E−3 Raynaud et al. (2006) N − abundance for NO3 assimilation N O −4 f 2 Fraction of nitrification flux lost as N2O – 6E Li et al. (2000) P associated −2 −1 kweather Parent material P weathering rate g P m year 0.004 Wang et al. (2010) −1 −6 koccl P occlude rate month 1.0E Yang et al. (2014) plant −1 a kP Reaction rate of plant POx carrier enzyme day 12 Colpaert et al. (1999) plant,P −2 KM Half-saturation constant for plant POx uptake g m 0.067 Cogliatti and Clarkson (1983) mic,P −2 KM Half-saturation constant for g m 0.02 Chen (1974) decomposing microbe POx immobilization surf −2 VMAXP Maximum mineral surface POx adsorption g m 133 Wang et al. (2010) surf,P −2 KM Half-saturation constant for mineral surface POx adsorption g m 64 Wang et al. (2010) [ plant] −2 · a EP Plant phosphorus carrier enzyme abundance for POx uptake g m Cfroot 0.0000125 Tang and Riley (2013); Trumbore et al. (2006) b F immob,pot [ mic] −2 P EP Decomposing microbes phosphorus g m 400 Tang and Riley (2013) carrier enzyme abundance for POx immobilization [ surf] −2 surf − [ ] EP Mineral surface “effective enzyme” abundance for POx adsorption g m VMAXP PS Tang and Riley (2013) a The scaling factor for plant nutrient enzyme abundance is 0.0000125. This number is inferred by assuming that growing season plant nutrient carrier enzymes are roughly the same order of magnitude compared with decomposing microbes’. Typical values for soil decomposing microbe biomass and tropical forest fine root biomass are 0.1 (Tang and Riley, 2013) and 400 gC m−2 (Trumbore et al., 2006). A typical value of scaling factor that scales microbial biomass to enzyme abundance is 0.05 (Tang and Riley, 2013). Therefore,Cfroot · x = Cmic · 0.05 or 400 · x = 0.1 · 0.05. We have x = 0.0000125. Further, we have plant plant plant plant plant plant k ·[E ] = VMAX . We know that typical values for VMAX and (E ) are 0.6 g m−2 day−1 (Min et al., 2000) and 0.005 g m−2. Then we have k = 120 day−1. Similarly, we have NH4 N NH4 NH4 N NH4 plant plant plant plant plant plant plant plant k · (E ) = VMAX , k ·[E ] = VMAX . Knowing that typical values for VMAX and VMAX are 0.01 (Min et al., 2000) and 0.06 (Colpaert et al., 1999) g m−2 day−1, we have NO3 N NO3 P P P NO3 P plant plant k = 2 and k = 12 day−1. b For decomposing microbes, we have VMAXmic = kmic ·[Emic]. Typical values for VMAXimmob and [Emic] are 5 g m−2 day−1 (Kuzyakov and Xu, 2013) and 0.005 g m−2 (Tang NO3 P N N N N N mic = mic and Riley, 2013). Therefore, we have kN 1000. Since our model calculates potential N immobilization rates and approximates them as VMAXN . The changes of potential N immobilization rates at each time step immob,pot immob,pot F F [ mic] = N = N immob [ mic] −2 −1 −2 mic = imply the changes of enzyme abundance through EN mic 1000 . Similarly, we have that VMAXP and EP are 2 g m day (Chen, 1974) and 0.005 g m . Therefore, kP 400 and kN immob,pot F mic = P EP 400 .

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 347 root or microbial biomass). Scaling factors are 0.0000125 Appendix Eqs. A10–11); (4) annual NPP (NPPannual in Ap- (for plants) and 0.05 (for decomposing microbes; Table 2). + dep − pendix Eq. A13); (5) NH4 deposition (FNH ); (6) NO3 de- surf] 4 (3) Mineral surface “effective enzyme” abundance [EP dep position (F ); and (7) hydrologic discharge (Qdis in Ap- is approximated by the available sorption surface area NO3 pendix Eq. A14). External inputs of mineral phosphorus are (VMAXsurf−[PS]). (4) Nitrifiers and denitrifiers are not ex- P derived from Mahowald et al. (2005, 2008). plicitly simulated, therefore we assume that their biomass and associated nutrient transporter abundance are fixed 2.3 Model parameterization and sensitivity analysis nit denit (EN ,EN ). For simplicity, we group the “decomposing mi- We constrained model parameters and performed sensitivity crobes/nitrifier/denitrifier/mineral surface nutrient carrier analyses using a suite of observations distinct from the ob- enzyme (E)” and their “base reaction rate k” into one single servations we used subsequently to test the model against the variable “VMAX” (see Appendix B for full derivation). N and P manipulation experiments. Because tropical systems Furthermore, we defined “potential rates (potential immobi- can be either nitrogen or phosphorous limited (or both; Elser lization, nitrification, denitrification, adsorption rates)” and et al., 2007; Vitousek et al., 2010), we chose observations used them as proxies of “VMAX”. Therefore, Eqs. (15), from a tropical forest site to constrain the N and P compe- (16), (18), (19), (21), and (22) become tition in our model (Tapajos National Forest, Para, Brazil; Table 3). [NH ] F immob = F immob,pot · 4 NH4 NH4 [ plant ] [ mic ] [ nit ] (23) In the parameter estimation procedure, several data mic,NH4 [NH4 ] [NO3 ] EN EN EN K (1 + mic,NH + mic,NO + plant,NH + mic,NH + nit,NH ) M K 4 K 3 K 4 K 4 K 4 M M M M M streams are assimilated into the N-COM model, including [NH ] nit = nit,pot · 4 + FNH FNH plant (24) 4 4 [E ] [Emic ] [Enit ] Knit,NH4 (1 + [NH4 ] + N + N + N ) measurements of soil NH concentrations, soil free phos- M nit,NH4 plant,NH4 mic,NH4 nit,NH4 4 KM KM KM KM [NO ] phate concentrations, sorbed phosphate concentrations and immob = immob,pot · 3 FNO FNO plant (25) 3 3 [E ] [Emic ] [Eden ] Kmic,NO3 (1 + [NH4 ] + [NO3 ] + N + N + N ) N O and CO flux measurements. The data sets are summa- M mic,NH4 mic,NO3 plant,NO3 mic,NO3 den,NO3 2 2 KM KM KM KM KM [NO ] rized in Table 3 and cover a wide range of N and P biogeo- den = den,pot · 3 FNO FNO plant (26) 3 3 [E ] [Emic ] [Eden ] Kden,NO3 (1 + [NO3 ] + N + N + N ) M den,NO3 plant,NO3 mic,NO3 den,NO3 chemistry dynamics. A set of model parameters is selected KM KM KM KM immob,pot [POx ] for calibration (Table 4), which comprise nutrient compe- F mic = F · P P plant (27) [E ] [Emic] [Esurf] k K ) mic,P + [POx ] + P + P + P tition kinetics parameters ( and M as well as the fast KM (1 ) Kmic,P plant,P Kmic,P Ksurf,P M KM M M soil carbon turnover time (TURNSOM). Because we had only surf,pot [POx ] F surf = F · a short-term CO2 respiration flux record, we were unable P P plant . (28) [E ] [Emic] [Esurf] surf,P + [POx ] + P + P + P to calibrate the longer turnover time parameters. However, KM (1 ) Ksurf,P plant,P Kmic,P Ksurf,P M KM M M since we test the calibrated model against short-term fertil- ization responses, this omission will not affect our evalua- In this case, the potential rates are treated as maximum reac- tion. Longer records from eddy covariance flux towers and tion rates (VMAX), because they are calculated without nu- 14C soil measurements are required to constrain the longer trient constraints or biotic and abiotic interactions. For exam- immob,pot turnover time pool values. ple, potential P immobilization rate (FP ) is based on We employed the MCMC approach (Ricciuto et al., 2008) the total phosphorus demand that can perfectly maintain the to assimilate the observations into N-COM. MCMC directly soil CP stoichiometry during soil organic matter decomposi- draws samples from a pre-defined parameter space and tries tion (Eq. A9). This potential immobilization rate represents to minimize a pre-defined cost function: the maximum phosphorus influx that the soil could take up at surf,pot J = (M(θ) − D)T R−1(M(θ) − D), (29) that moment. The maximum adsorption rate (FP ) is the time derivative of the Langmuir equation (Eq. A12), which where M(θ) and D are vectors of model outputs and ob- is a theoretically maximal adsorption rate excluding all other servations including time series of different simulated vari- biotic and abiotic interactions. The potential rates (VMAX) ables (e.g., soil CO2 and N2O effluxes and soil concentra- + are updated by the model rather than calibrated, except for tions of NH , free POx and sorbed POx); θ is a vector of surf surf 4 VMAXP . VMAXP denotes the maximum adsorption ca- model parameters (θi); and i from 1 to 20 represents the pacity (not maximum adsorption rate), which affects the po- parameters that are calibrated (Table 4). R−1 is the inverse surf,pot tential adsorption rate (FP ). of data error covariance matrix. We assumed that diagonal The model is run on an hourly time step, initialized with elements are 40 % of observed values and off-diagonal ele- state variables and critical parameters (Table 2). Since the ments are zeros. We further assumed that the prior parame- model is designed to be a component of the and ter follows a lognormal distribution. µ and σ were 0.91 and ACME Land Models (CLM, ALM; which are essentially cur- 0.95 of their initial values, respectively (Table 4). We then rently equivalent), we used CLM4.5 site-level simulations ran MCMC to sample 50 000 parameter pairs (Fig. A1). The to acquire temporally-resolved: (1) soil temperature factors second half of the samples was fit to a Gaussian distribution. on decomposition (rT ); (2) soil moisture factors on decom- We also employed the Gelman–Rubin criterion to quantita- anox position (rθ ); (3) the anoxic fraction of soil pores (f in tively show whether or not the MCMC chain converged. The www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 348 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Table 3. Observational data sets used for calibration. Number of observations for each data stream is included in brackets.

Processes Data sets Location References C associated Soil heterotrophic Tapajos National Forest, (Silver et al., 2012) respiration (20) Para, Brazil + N associated Soil NH4 (5) N2O efflux (20) Tapajos National Forest, (Silver et al., 2012) Para, Brazil P associated Soil free phosphate (3) Sorb phosphate (3) Tapajos National (McGroddy et al., 2008) Forest, Para, Brazil

Table 4. Calibrated parameters are reported in terms of (1) mean/standard deviation by fitting to a Gaussian distribution; (2) 25 and 75 % quantile. Both variance-based and quantile-based parameters uncertainty reduction are provided; (3) Gelman–Rubin convergence criterion.

25 75 25 75 Parameters µprior σprior µposterior σposterior UR Qprior Qprior Qposterior Qposterior UR Gelman–Rubin criterion

TURNSOM (3.7, 0.06, (3.9, 0.06, (5.2, 0.07, (0.33, 0.01, (92, 83, (5.33, 0.086, (19.32, 0.31, (5.05, 0.63, (5.39, 0.076, (97, 94, (1.69, 1.03, (CWD, metabolic, cellulose, 0.23, 0.23, 0.24, 0.24, 0.17, 0.17, 0.01, 0.005, 96, 98, 0.33, 0.33, 1.18, 1.18, 0.16, 0.17, 0.18, 0.18, 97, 99, 1.75, 1.01, lignin lit, fast, medium SOM) 0.16, 4.6) 0.18, 4.8) 0.14, 3.6) 0.008, 0.37) 96, 92) 0.22, 6.5) 0.8, 23.5) 0.13, 3.2) 0.14, 3.9) 98, 96 1.06, 1.55) plant k 109 114 58 14 88 156.1 565.4 52.8 60.0 98 1.87 NH4 plant,NH4 KM 0.082 0.086 0.173 0.018 79 0.12 0.42 0.16 0.18 93 2.86 mic,NH4 KM 0.018 0.019 0.071 0.0067 65 0.026 0.094 0.065 0.076 85 1.94 knit 0.091 0.095 0.37 0.038 60 0.13 0.47 0.36 0.39 91 2.02 nit,NH4 KM 0.069 0.072 0.082 0.012 83 0.10 0.36 0.07 0.09 94 1.95 plant k 1.8 1.9 7.6 1.7 13 2.60 9.42 6.11 9.14 56 1.01 NO3 plant,NO3 KM 0.064 0.067 0.085 0.0064 90 0.09 0.33 0.08 0.09 97 3.17 mic,NO3 KM 0.036 0.038 0.096 0.014 63 0.05 0.19 0.09 0.10 92 2.64 den,NO3 KM 0.0101 0.0105 0.022 0.0034 68 0.014 0.052 0.019 0.024 87 1.03 plant kP 11 11.5 59 0.75 93 15.61 56.54 58.86 59.81 98 1.06 plant,P KM 0.061 0.064 0.11 0.015 77 0.09 0.32 0.10 0.12 94 1.52 mic,P KM 0.018 0.019 0.037 0.0047 75 0.026 0.094 0.034 0.039 93 2.86 surf VMAXP 121 127 182 30 76 173.0 626.6 156.5 206.3 89 2.25 surf,P KM 64 58 200 50 18 83.2 301.5 162.6 233.0 68 1.05 calibrated model parameters are reported in terms of means (by 5 ◦C); and (3) elevated soil moisture (by 50 %). SOBOL and standard deviations. Uncertainty reduction (UR) is calcu- sampling (Pappas et al., 2013), a global sensitivity technique, lated based on (1) variance (Eq. 30a) and (30b) 25 and 75 % is employed to calculate the sensitivities of output variables quantile (Eq. 30b): with respect to various inputs:

σposterior UR = (1 − ) · 100% (30a) σ | σprior VARpi (Ep∼i (Y pi)) Si = , (31) 75 − 25 VAR(Y ) Qposterior Qposterior UR = (1 − ) · 100%, (30b) Q 75 − 25 Qprior Qprior where Si is the first-order sensitivity index of the ith param- eter and ranges from 0 to 1. By comparing the values of Si, where σprior is prior parameter uncertainty, which is 95 % of we were able to evaluate which processes affect the pattern the parameter initial value. The σposterior is calibrated param- of nutrient competition. Y represents the model outputs of eter uncertainty, which is calculated by fitting the calibrated + − plant NH , NO or POx uptake; pi is the target parame- model parameters to a Gaussian distribution. Q75 and Q25 4 3 ter; p∼i denotes all parameters that are associated with nutri- are 75 and 25 % percentage quantile of each parameter. UR ent competition except the target parameter; and VAR(.) and is a useful metric (Zhu and Zhuang, 2014), because it quan- E(.) represent variance and mean, respectively. titatively reveals the reduction in the range of a particular pa- rameter after calibration with MCMC. It does not, however, 2.4 Model application indicate that the parameter itself is more consistent with ob- served values of the parameter. A large value of UR implies After calibration, we applied the N-COM model to several a more robust model. tropical forest nutrient fertilization studies not included in In addition, we conducted a sensitivity study to identify the calibration data set, where isotopically labeled nitrogen the dominant controlling factors regulating nutrient competi- or phosphorous was injected into the soil. The fer- tion in N-COM. Three scenarios were considered: (1) base- tilization experiments measured the fate of added nutrients; line climate and soil conditions; (2) elevated soil temperature for example, identifying the fraction of added N or P that

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 349

+ − 3− Table 5. Short-term (24 or 48 h) fertilization experiments off NH4 , NO3 or PO4 additions used to evaluate the performance of the N-COM competition scheme.

Data sets Added Competitors Duration References nutrient (hour) 3− −1 PO4 fertilization 10 µg g I. Mineral II. Decomposing 48 Olander and Vitousek (2005) surface microbe + −1 NH4 fertilization 4.6 µg g I. Plant II. Decomposing III. Nitrifier 24 Templer et al. (2008) microbe − −1 NO3 fertilization 0.92 µg g I. Plant II. Decomposing 24 Templer et al. (2008) microbe goes into the plant, is immobilized by microbes, or is stabi- 3 Results and discussion lized by mineral surfaces. These measurements offer an ef- fective baseline to test whether the N-COM model captures 3.1 Calibrated model parameters short-term nutrient competition. Because we have focused in this paper on applications Our best estimates (second half of the MCMC chain) of the in tropical forests, we choose three tropical forest fertiliza- 3− + − selected model parameters based on the observations at the tion experiments with (1) PO4 , (2) NH4 and (3) NO3 ad- Tapajos National Forest, Para, Brazil are shown in Fig. 2. 3− ditions (Table 5). The PO4 fertilization experiment (Olan- We found that calibrated parameter samples were not heav- der and Vitousek, 2005) was conducted in three Hawaiian ily tailed and they generally follow Gaussian distributions tropical forests along a soil chronosequence (300, 20 000 and (Fig. A3). In order to quantitatively compare the calibrated −1 4 100 000 year old soils) that were fertilized with 10 µg g parameter distributions with prior distributions, we fit pa- 32 3− PO4 , respectively, and microbial demand vs. soil sorption rameter samples to a Gaussian distribution and estimated its was measured. We did not evaluate the role of plants in phos- means and standard deviations (Table 4). phorus competition for the Hawaii sites, since plant phospho- Even though the parameter mean was improved, the un- rus uptake was not measured in those field studies. Our model certainty may still be relatively large. In other words, a prog- discriminates the Hawaii sites along the chronosequence by nostic prediction based on these calibrated parameters could setting distinct initial pool sizes (derived from Olander and be relatively uncertain (Scholze et al., 2007), due to large Vitousek, 2004, 2005) of soil organic carbon, nitrogen and uncertainty associated with the calibrated parameters. There- phosphorus, and soil parent material phosphorus. + − fore, we calculated the variance-based UR (URσ ; Eq. 30a) We also used measurements from NH4 and NO3 fertiliza- to evaluate model improvement in terms of parameter uncer- tion studies located at the Luquillo tropical forest in Puerto tainty. We found that parameters’ uncertainties were reduced −1 15 + Rico (Templer et al., 2008). In that study, 4.6 µg g NH4 by 13–98 %. This calculation might either overestimate or was added into the highly weathered tropical forest soil and underestimate the URσ , due to the fact that the calibrated pa- 15 + the consumption of NH4 by plant roots, decomposing rameters did not strictly follow Gaussian distributions. But microbes and nitrifiers were measured. In the same study, the actual URσ should not be far from our estimates, be- −1 15 − 0.92 µg g NO3 was added to the soil and the plant uptake cause these samples were not widely spread across the poten- and microbial immobilization was measured. The measure- tial parameter space (Fig. 2). The least constrained parame- ments were made 24 or 48 h after the were added. ter was kplant (reaction rate of plant nitrogen carrier enzyme For the model scenarios, we (1) spun up the N-COM NO3 with NO− substrate). Two other NO− dynamics related pa- model for 100 years; (2) perturbed the soil nutrient pool by 3 3 mic,NO3 rameters were also not well constrained: URσ of K the same amount as the fertilization; (3) ran the model for 24 M− or 48 h and calculated how much of the added nutrients were (half-saturation constant for decomposing microbe NO3 im- absorbed by plants, microbes or mineral surfaces; and (4) mobilization) and Kden,NO3 (half-saturation constant for den- − M compared our model simulations with the observed data to itrifier NO3 consumption) were only 63 and 68 %, respec- assess model predictability. The 100-year spin-up simulation + tively. Compared with NH4 or POx competition-related pa- aimed at eliminating the effects of imposed initial inorganic − rameters, we concluded that parameters associated with NO3 pool sizes on fertilization experiments, rather than accumu- competition were the least constrained in the model. This re- lating soil organic matter in the system, since we initialized − sult was primarily due to the lack of NO3 pool size data, the soil organic carbon pools from CLM4.5 steady-state pre- − and secondarily due to the fact that NO3 was not the ma- dictions. jor nitrogen source for plant or decomposing microbes. We also provide quantile-based UR for reference (Table 4). The www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 350 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

k k k turn k CWD mat lit cel lit lig lit fast SOM 5000 5000 5000 5000 5000

0 0 0 0 0 5 10 15 20 0.1 0.2 0.3 0.2 0.4 0.6 0.8 1 1.2 0.2 0.4 0.6 0.8 1 1.2 0.2 0.4 0.6 0.8 k kplant Kplant,NH4 Kmic,NH4 k med SOM NH4 M M nit 5000 5000 5000 5000 5000

0 0 0 0 0 5 10 15 20 25 100 200 300 400 500 600 0.1 0.2 0.3 0.4 0.02 0.04 0.06 0.08 0.1 0.1 0.2 0.3 0.4 0.5 Knit,NH4 kplant Kplant,NO3 Kmic,NO3 Kden,NO3 M NO3 M M M 5000 5000 5000 5000 5000

0 0 0 0 0 0.1 0.2 0.3 2 4 6 8 10 0.1 0.2 0.3 0.05 0.1 0.15 0.2 0.01 0.02 0.03 0.04 0.05 kplant Kplant,P Kmic,P VMAXsurf Ksurf,P P M M P M 5000 5000 5000 5000 5000

0 0 0 0 0 10 20 30 40 50 60 0.1 0.2 0.3 0.02 0.04 0.06 0.08 0.1 200 400 600 100 200 300

Figure 2. Distribution of prior and calibrated model parameters. above-mentioned conclusions still hold with quantile-based ECAplant NH4 URQ, although the quantile-based URQ is generally higher [NH ]·[Eplant ] = 4 N . [ plant ] [ mic ] [ nit ] (33) plant,NH4 [NH4 ] [NO3 ] EN EN EN K (1 + plant,NH + plant,NO + plant,NH + mic,NH + nit,NH ) than variance-based URσ . One parameter was calibrated to M K 4 K 3 K 4 K 4 K 4 plant M M M M M be at the upper boundary of its prior ranges (kP ), implying that this tropical plant is highly efficient in phosphorus up- Other “consumer-substrate reactions” have similar forms. take. Although we do not have direct kinetic parameter obser- Under a nutrient-abundant situation (e.g., fertilized agricul- vations for the specific tropical species involved in our study, ture ecosystem), the relative competitiveness of each con- an inferred high phosphorus uptake efficiency is reasonable sumer (ECA) is dominated by its specific enzyme abundance for tropical species that have adapted to these phosphorus- [E]. Under such conditions, substrate affinity is no longer deficient environments (Begum and Islam, 2005; Föhse et al., a controlling factor. In contrast, under nutrient-limited con- 1988). ditions (e.g., many natural ecosystems), ECA is dominated Convergence of model parameters is reported with the by the specific enzyme abundance as well as the substrate Gelman–Rubin criterion (univariate potential scale reduction affinity ([E]/KM ). Therefore, consumers could either en- factor; Table 4 and Fig. A2). Using this criterion, seven (out able an alternative high-affinity nutrient transporter system of twenty) parameters are found to converge (Gelman–Rubin (low KM ) or exude more enzyme to enhance competitive- ≤ 1.1). The lack of convergence (in addition, 20-dimensional ness. For example, at the whole-soil scale it has been shown multivariate potential scale reduction factor is 12.04) of the that root spatial occupation (Cfroot) determines a plant’s com- remaining parameters is partly due to data paucity. In par- petitiveness when low soil nutrient diffusivity is limiting nu- ticular, starting from different initial values, MCMC calibra- trient supply (Raynaud and Leadley, 2004). Consistently, our tions may result in different models that give rise to simi- results highlighted the dominant role of nutrient carrier en- lar model–data misfit (i.e., “equifinality”; Tang and Zhuang, zyme abundance (E proportional to Cfroot) in controlling 2008). In this regard, high-frequency measurements may im- competition. If we further assumed that plants, decompos- prove model calibration (see more discussion in Sect. 3.3). ing microbes and nitrifiers enzyme abundances were approx- The non-convergence of model parameters implies an im- imately equal, we will have that their relative competitive- + : : plant,NH4 : perfect model. Therefore, for large-scale model application, ness in acquiring NH4 was about 4 10 9 (1/KM more work on data collection, parameter tuning and uncer- mic,NH4 : nit,NH4 1/KM 1/KM ). However, such results could not tainty analysis is needed. However, even with these caveats, be easily generalized to other ecosystems, because they heav- the model predictability is reasonably good when applied ily relied on the traits (affinity) of specific competitors. For to the tropical forest fertilization experiments described in a different ecosystem, those traits would be drastically dif- Sect. 3.4. ferent due to the change of, e.g., plant species composi- We re-organize the right-hand sides of Eqs. (14)–(22) to tion and microbial community structure. Even for the same be the product of potential nutrient uptake rate and an ECA + ecosystem, those traits could be highly heterogeneous. For limitation term; for example for plant NH4 uptake, example, the community structure of decomposing microbes could be different in rhizosphere and bulk soil (with different F plant = kplant · ECAplant (32) K ). However, in this work we assumed a well-mixed en- NH4 NH4 NH4 M

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 351 vironment (one soil column), in order to be consistent with consistently sensitive to the same parameters across all tem- large-scale ecosystem models. Although beyond the scope perature and moisture conditions. The environment affects of the current study, the consequences of ignoring the rhi- the nutrient competition primarily through altering the nu- zosphere vs. bulk soil heterogeneity warrants further investi- trient abundance. Enhanced soil temperature and soil mois- gation. Large-scale models aim to quantify ecosystem level ture accelerated soil organic carbon turnover, thereby releas- dynamics, although they are usually driven by parameters ing more inorganic nutrient into the soil (gross mineraliza- inferred from in situ field observations. In the absence of tion). However, the impacts on plant nutrient uptake are lim- a model that explicitly represents this spatial heterogeneity, ited (Fig. 3) because the enhanced soil organic matter de- it is difficult to quantify the impacts of using inferred rhi- cay also requires higher immobilization fluxes to sustain the zosphere affinities on model predictions of the soil organic matter CNP stoichiometry. The enhancement of whole soil (Schimel et al., 1989). Furthermore, the assump- net mineralization would be limited, and therefore would not tion of well-mixed environment in large-scale model is an change soil nutrient status dramatically. inevitable flaw, because of large computational demands and a lack of scale-aware parameters and model structures for 3.3 Model performance large-scale models to run fine-scale simulations. Although in this study ECA was applied to a large-scale The prior and calibrated models were compared against ob- model, the competition framework is readily applicable to servational data sets of pool sizes of soil free phosphate, fine-scale models that consider soil heterogeneity. In fine- + sorbed phosphate, and NH4 , CO2 efflux and N2O efflux scale models, bulk soil nutrient competition can occur only (Fig. 4). We note that although we attempted to acquire as among different microbes because they are ubiquitous in the many data sets that contained these five observations as pos- soil (e.g., nitrifier vs. microbial decomposer), while rhizo- sible, more observations in tropical ecosystems would clearly sphere nutrient competition occurs among plants and mi- improve the parameter estimates. For example, in the ex- crobes (e.g., nitrifier vs. microbial decomposer vs. roots). periment we analyzed, only three measurements of soil free This distinction implies that the competitiveness parameters phosphate were made during 1999. Many detailed dynam- we infer here for N-COM, which does not currently explic- ics are therefore missing and could impact our parameter itly represent bulk versus rhizosphere processes, subsume estimates. The prior model predicted an increasing trend of the range of fine-scale processes controlling nutrient uptake. soil free POx, which resulted from underestimates of plant More research is required to link these different model spatial plant P uptake (by underestimating of kP ) and soil microbial scales, theory and parameterizations. mic,P Our modeling framework highlights the important concept P immobilization (by overestimating KM ). The calibrated that “competitiveness” is a dynamic property of the competi- model captured the seasonal dynamics of soil free POx rea- tion network, and more importantly that it is linked to com- sonably well: increases during the wet season and grad- petitor functional traits (affinity and nutrient carrier enzyme ual decreasing during the dry season (August to Novem- abundance). This concept is in contrast to the prevailing as- ber). The prior model also largely underestimated the sea- sonal variability of nitrogen dynamics and underestimated sumption underlying all major large-scale ecosystem models, + + which either assume “relative demand competitiveness for the NH4 pool size due to overestimation of plant NH4 up- take (kplant). In addition, it also underestimated the denitri- different nutrient consumers” (Thornton et al., 2007) or “soil NH4 microbes outcompete plants” (McGuire et al., 1992; Parton fication N2O emissions, because of an underestimation of + − et al., 1988). Imposing such pre-defined orders of competi- NH4 to NO3 transformation rate (knit). Consequently, there − tiveness neglects the diversity of nutrient competitors (plants was not enough NO3 substrate to react with denitrifiers and and microbes) and their differences in nutrient uptake capac- release N2O. The calibrated model, however, accurately re- + ity expressed by relevant functional traits. Our model frame- produced the seasonal dynamics of both NH4 pool sizes and work offers a theoretically consistent approach to account for soil N2O emissions. There were small differences between the diversity of nutrient competition in different competitor the prior and calibrated model predictions of soil CO2 emis- networks. sions. The CO2 and N2O effluxes were more frequently ob- served at Tapajos National Forest during 1999 to 2001, com- 3.2 Model sensitivity analysis pared with phosphorus data. Most of the measurements were collected during the wet season. Therefore the modeled CO2 Through sensitivity analysis, we separately investigated the and N2O emissions were largely improved by assimilating + − factors controlling plant NH4 , NO3 and POx competition these data sets. (Fig. 3). Each sensitivity analysis consisted of three scenar- The model performance implies that after assimilating ios: (1) normal conditions (control); (2) elevated soil temper- multiple data sets, our model predictions were improved over ature (+Ts); and (3) elevated soil moisture (+θ). The sen- the prior model. However, it is clear that more observations sitivity analysis indicates that the model is highly sensitive of the metrics applied in our MCMC approach would bene- to kinetics parameters (e.g., KM ). Furthermore, the model is fit the model calibration. Unfortunately, because of our focus www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 352 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Plant NH4 uptake Plant NO3 uptake Plant PO4 uptake

0.4 0.4 0.5 +ST +ST +ST

0.2 0.2 0.25

Control Control Control

+SM +SM +SM

plant k plant plant NH4 k k NO3 P plant E plant plant NH4 E E NO3 P plant,NH4 K plant,NO3 plant,P M K K M M k mic,NO3 nit K k M fast SOM Kmic,NH4 M fast SOM CP fast SOM CN surf den,NO3 VMAX fast SOM CN K P nit,NH4 M K Ksurf,P M k M fast SOM k Kmic,P fast SOM M

Figure 3. Model sensitivity analysis with SOBOL sampling. For each metric, three scenarios are shown: baseline (Control), elevated soil ◦ temperature by 5 C(+Ts), and elevated soil moisture by 50 % (+θ), respectively. The length of bar (plot in polar coordinate) is the sensitivity (unit-less) of model output with respect to model input variables. Our results showed that the plant nutrient uptake was mostly regulated by internal consumer–substrate uptake kinetics rather than the external environmental conditions (e.g., Ts, θ).

PO4 SP NH4 0.6 2.5 0.8

0.7 0.5 2 0.6 0.4 1.5 0.5 2 2 2 −

− 0.3 − 1 0.4 gP m gP m 0.2 gN m 0.3 0.5 0.1 0.2 0 0 0.1

−0.1 −0.5 0 1999 1999 2000 2000 2001

−3 CO2 x 10 N2O 6 8 Model posterior uncertainty 7 5 Model posterior mean 6 Model prior

1 4 1 Observation − − 5 day day 2 2

− 3 − 4 gN m 3 gC m 2 2 1 1

0 0 2000 2001 2000 2001

Figure 4. Model performance at Tapajos National Forest, Para, Brazil. Overall, the calibrated model (blue line) improved predictions over the prior model (gray line) when compared to observations. Green areas indicate the calibrated model uncertainties.

on tropical sites, we were unable to acquire more data sets sampling of those sparse measurements (e.g., POx) would that had the full suite of measurements required. Data sets significantly benefit the model uncertainty reduction. − of soil nutrient pool sizes (e.g., NO3 ) and higher frequency

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 353

(a) Hawaii 300 yr soil (b) Hawaii 20K yr soil (c) Hawaii 4.1M yr soil 0.8 0.8 2 ) ) ) −2 −2 −2 1.5 0.6 0.6 1 0.4 0.4 0.5 uptake (g m uptake (g m uptake (g m 0.2 0.2 3− 4 3− 4 3− 4 0 PO PO PO 0 0 −0.5 Decomp. Mineral Decomp. Mineral Decomp. Mineral microbe microbe microbe (d) Puerto Rico (Luquillo) (e) Puerto Rico (Luquillo) 0.2 0.06 ) ) −2 −2 0.05 Model 0.15 Obs 0.04

0.1 0.03

0.02 uptake (g m uptake (g m − 3 + 4 0.05 0.01 NO NH 0 0 Plant Decomp. Nitrifier Plant Decomp. microbe microbe

Figure 5. Model perturbation experiments compared with nitrogen and phosphorus fertilization field experimental data. The blue dots show the difference between control and perturbed simulations, which mean how much newly added nutrient each consumer takes up. The red circles are recovered isotopically labeled nutrient within each consumer. Since plants phosphorus uptake was not measured at Hawaii sites, we did not include the plants in the perturbation study.

3.4 Model testing against nitrogen and phosphorus improved substrate quality (Kaspari et al., 2008; Wieder et fertilization studies al., 2009). In the Puerto Rican Luquillo forest nitrogen addition ex- periments, partitioning of added ammonium between plants To test the calibrated N-COM model, we conducted short- and heterotrophic was well captured by the N-COM term numerical competition experiments (24 or 48 h simula- model, with no significant differences between model predic- tions) by manually imposing an input flux into nutrient pools tions and observations (Fig. 5d). However, the model under- + − equivalent to the N and P fertilization experiments described estimated nitrifier NH4 uptake. NO3 competition in this site above and in Table 5. The simulated results were compared was also relatively accurately predicted (Fig. 5e), although with observations from the field manipulations. the measurements did not include denitrification. Model es- − − In the P addition experiments across the Hawaiian timates of plant NO3 uptake and microbial NO3 immobi- chronosequence, the partitioning of phosphate between mi- lization were consistent with the observed ranges, but we crobes and mineral surfaces was well represented by the N- highlight the large observational uncertainties, particularly − COM model in the intermediate (20 K yr) and old (4.1 M yr) for microbial NO3 uptake. sites (Fig. 5b and c), with no significant differences be- In the pseudo-first-order decomposition model we applied tween model predictions and observations. In the youngest here to demonstrate the ECA competition methodology, the Hawaiian site (300 years; Fig. 5a), the relative partitioning soil organic matter C : N : P ratio also limited microbial N 3− and P uptake. For this type of decomposition model, stoichio- was correctly simulated, but the predicted PO4 magnitudes were lower than observations. Our simulations indicated that metric differences between soil organic matter and microbes at the young soil site the added P exceeded microbial de- are not dynamically simulated. Such a simplification of soil mand, resulting in lower predicted microbial P uptake than and microbial stoichiometry favors large spatial scale model observed. This discrepancy reflected a possible deficiency of structures over long temporal periods, but hampers predic- first-order SOC decay models (as we used here), which im- tion of microbial short-term responses to N and P fertiliza- plicitly treat microbes as a part of soil organic matter. Since tion. For example, the observed difference between microbial microbial nutrient immobilization is strictly regulated by the and soil C : P ratios can be as large as 6-fold (Mooshammer SOC turnover rate in this type of model, external nutrient et al., 2014; Xu et al., 2013). Were that the case in the ob- inputs will no longer affect microbial nutrient uptake if the servations we applied, the potential soil P demand calculated inputs exceed potential microbial demand. We therefore be- based on a fixed soil organic matter C : P ratio could be only lieve that explicit microbe–enzyme models might be able to 17 % of that based on microbial C : P ratio. 3− better explain the strong microbe PO4 uptake signal ob- served at the young Hawaii fertilization experiment site. Mi- 3.5 Implications of ECA competition treatment crobial models explicitly simulate the dynamics of micro- bial biomass, which might be able to capture the expected Terrestrial ecosystem growth and function are continuously rapid growth of microbial communities under conditions of altered by climate (e.g., warming, drought; Chaves et al., www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 354 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

2003; Springate and Kover, 2014), external nutrient inputs biogeochemistry models with traditional treatments of nutri- (e.g., N deposition; Matson et al., 1999, 2002), and atmo- ent competition likely underestimate plant nutrient uptake. spheric composition (e.g., CO2 concentration; Norby et al., Nutrient competition should be treated as a com- 2010; Oren et al., 2001; Reich et al., 2006). Improved under- plex consumer–substrate reaction network: multiple “con- standing of the underlying mechanisms regulating ecosys- sumers”, including plant roots, soil heterotrophic microbes, tem responses to environmental changes has been obtained nitrifiers, denitrifiers and mineral surfaces, each competing through in situ level to large-scale and long-term manipula- for substrates of organic and inorganic nitrogen and phospho- tion experiments. For example, decade-long Free-Air Carbon rus as nutrient supply. In such a model structure, the success Dioxide Enrichment (FACE) experiments have revealed that of any consumer in substrate acquisition is affected by its nitrogen limitation diminished the CO2 fertilization effect of consumer–substrate affinity (Nedwell, 1999). Such competi- forest (Norby et al., 2010) and grassland (Reich and Hobbie, tive interactions have been successfully applied to microbe– 2013) ecosystems. However, fewer efforts have been made microbe and plant–microbe substrate competition modeling towards incorporating the observed process-level knowledge (Bonachela et al., 2011; Lambers et al., 2009; Maggi et al., into ESMs. Therefore, a major uncertainty that has limited 2008; Maggi and Riley, 2009; Moorhead and Sinsabaugh, the predictability of ESMs has been the incomplete repre- 2006; Reynolds and Pacala, 1993) for many years. sentation of soil nutrient dynamics (Zaehle et al., 2014). Even Here, we applied the consumer–substrate network in a though new soil paradigms were proposed dur- broader context of plant, microorganism and abiotic min- ing recent decades (Korsaeth et al., 2001; Schimel and Ben- eral interactions. We analyzed the consumer-substrate net- nett, 2004), they were restricted to either conceptual models work using a first-order accurate equilibrium chemistry ap- or only applied to explain laboratory experiments. proximation (ECA; Tang and Riley, 2013; Zhu and Ri- Many large-scale terrestrial biogeochemistry models (e.g., ley, 2015). Our sensitivity analysis confirmed that the O-CN, CASA, TEM) have adopted the classical paradigm consumer–substrate affinity and nutrient carrier enzyme that microbes decompose soil organic matter and release abundance were the most important factors regulating rel- + NH4 as a “” product (Waksman, 1931). The rate of this atively short-term competitive interactions. The ECA com- process is defined as “net N mineralization”, and is adopted petition treatment represents ecosystem responses to envi- as a “measure” of plant available inorganic N (Schimel and ronmental changes and has the potential to be linked to a Bennett, 2004). This classical paradigm overlooked the fact microbe-explicit land biogeochemistry model. The approach that “net N mineralization” actually comprised two individ- allows competition between plants, microbes and mineral ual processes – gross N mineralization and microbial N im- surfaces to be prognostically determined based on nutrient mobilization. Implicitly, the classical paradigm assumes that status and capabilities of each consumer. the microbes have priority to assimilate as much of the avail- able nutrient pool as possible. Soil nutrients were only avail- able for plant uptake if there were not enough free energy 4 Conclusions materials (e.g., dissolved soil organic carbon) to support mi- crobial metabolism. As a result, soil microbes were consid- In this study, we developed a soil biogeochemistry model (N- ered “victors” in the short-term nutrient competition. Some COM) that resolves the dynamics of soil nitrogen and phos- other large-scale terrestrial biogeochemistry models (e.g., phorus, plant uptake of nutrients, microbial uptake and abi- CLM4CN), simplify the concept of nutrient competition dif- otic interactions. We focused on the implementation, param- ferently. They calculate the plant N uptake and soil N immo- eterization and testing of the nutrient competition scheme bilization separately; and then down-regulate the two fluxes that we plan to incorporate into the ESM land models according to the soil mineral N availability. As a result, plant CLM and ALM. We described the multiple-consumer and and soil microbe competitiveness for nutrients is determined multiple-nutrient competition network with the Equilibrium by their relative demand. Chemical Approximation (ECA; Tang and Riley, 2013) con- Climate-scale land models have over-simplified or ignored sidering two inhibitive effects: (1) multiple substrates (e.g., + − competition between plants, microbes, and abiotic mecha- NH4 and NO3 ) sharing one consumer inhibits the effec- nisms. In reality, under high nutrient stress conditions, plants tive binding between any specific substrate and the consumer can exude nutrient carrier enzymes or facilitate mycorrhizal and (2) multiple consumers (e.g., plants, decomposing mi- + fungi associations to enhance competitiveness for nutrient crobes, nitrifiers) sharing one substrate (e.g., NH4 ) lowers acquisition (Drake et al., 2011; Hobbie and Hobbie, 2006; the probability of effective binding between any consumer Treseder and Vitousek, 2001). In addition, plants can adjust and that substrate. We calibrated the model at a tropical for- C allocation to construct more fine roots, which scavenge nu- est site with highly weathered soil (Tapajos National Forest, trients over larger soil volumes (Iversen et al., 2011; Jackson Para, Brazil), using multiple observational data sets with the et al., 2009; Norby et al., 2004). Soil spatial heterogeneity MCMC approach. The calibrated model compared to mul- might also contribute to the success of plant nutrient com- tiple categories of observational data was substantially im- petition (Korsaeth et al., 2001). Therefore, most ecosystem proved over the prior model (Fig. 4). The seasonal dynam-

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 355 ics of soil carbon, nitrogen and phosphorus were moderately This study is an important step towards implementing well captured. However, our results would likely be more more realistic nutrient competition schemes in complex robust if more temporally resolved observations of carbon, climate-scale land models. Traditional ESMs generally lack nitrogen and phosphorous were available. Although the cali- realistic soil nutrient competition, which likely biases the brated model is the best one we can derive based on limited estimates of terrestrial ecosystem carbon productivity and data, several model parameters were not well converged. We biosphere–climate feedbacks. This study showed the ef- therefore conclude that more work on data collection, param- fectiveness of ECA kinetics in representing soil multiple- eter tuning and uncertainty analysis is needed. consumer and multiple-nutrient competition networks. Of- To test the resulting model using the calibrated parame- fline calibration and independent site-level testing is criti- ters, we applied N-COM to two other tropical forests (Hawaii cally important to ensuring the newly incorporated model tropical forest and Luquillo tropical forest) not used in the will perform reasonably when integrated in a complex ESM. calibration process and conducted nutrient perturbation stud- To this end, we provide a universal calibration approach us- ies consistent with fertilization experiments at these sites. ing MCMC, which could in the future be used to further con- The results showed that N-COM simulated the nitrogen and strain N-COM across plant functional types, climate and soil phosphorus competition well for the majority of the obser- types. + vational metrics. However, the model underestimated NH4 uptake by nitrifiers, probably due to the loosely constrained − nitrification parameters that were the result of NO3 pool size data paucity during calibration at the Brazil site (Ta- ble 4). Data sets of soil nutrient pool sizes and CO2 and N2O effluxes with high-frequency sampling would signifi- cantly benefit the model uncertainty reduction. To date, many terrestrial ecosystem biogeochemistry mod- els assume that microbes outcompete plants and immobilize nutrients first (Wang et al., 2007; Zaehle and Friend, 2010; Zhu and Zhuang, 2013), although CLM currently assumes constant and relative demand competitiveness of plants and microbes. Few models, to our knowledge, consider the role of abiotic interactions in the competitive interactions. In the case of microbes outcompeting plants, the plant is only able to utilize the nutrients that exceed microbial demands during that time step. The leftover nutrients are defined as net min- eralization, which is a widely adopted concept in soil bio- modeling (Schimel and Bennett, 2004). These models oversimplify plant–microbe interactions by impos- ing dubious assumptions (e.g., microbes always win against plants). We showed that (in Sect. 3.1) “competitiveness” is a dynamic rather than fixed property of the competition net- work, and more importantly, it should be linked to competitor functional traits (affinity and nutrient carrier enzyme abun- dance).

www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 356 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Appendix A: CNP fluxes Besides decomposing microbe nutrient immobilization, other potential nutrient uptakes are The fluxes of carbon, nitrogen and phosphorus coming from the upstream pool (i) to the downstream pool (j) due to SOM F nit,pot = ([NH ]· k · r · r · (1 − f anox) (A10) NH4 4 nit θ T decomposition are calculated as F den,pot = min(f (decomp),f ([NO ])) · f anox (A11) NO3 3 move = dec FC,ij fij FC,i gi (A1) VMAXsurf · Ksurf,P d[PO ] surf,pot = P M · x   FP , (A12) move dec 1 gi (Ksurf,P + [PO ])2 dt FN,ij = fij FC,i min , (A2) M x CNi CNj   nit,pot den,pot surf,pot 1 gi where F , F and F are potential rates for move dec NH4 NO3 P FP,ij = fij FC,i min , , (A3) + − CPi CPj NH4 nitrification, NO3 denitrification and mineral surface + POx adsorption. knit is the maximum fraction of free NH4 where gi is the percentage of carbon remaining in the soil pool that could be utilized by nitrifiers. The potential ni- after decomposition of the ith SOM pool (i.e., CUE, with trification rate is controlled by soil temperature (rT ), soil the rest being released as CO2); fij is the fraction of SOM anox moisture (rθ ) and soil oxygen status (1 − f ). The po- leaving the ith pool and entering the jth pool; and F dec is C,i tential denitrification rate (F den,pot) is either constrained NO3 the first-order decay of the ith SOM pool. CN and CP are − soil C : N and C : P ratios, respectively. by substrate availability (f (decomp)) or NO3 availability If the upstream-decomposed soil organic nitrogen (phos- (f ([NO3]); Del Grosso et al., 2000), taking into account the anox surf,pot phorus) is more than enough to sustain the downstream C : N soil anaerobic condition (f ). FP is derived from the (C : P) ratio, then the excess nitrogen (phosphorus) enters Langmuir adsorption model (Barrow, 1978), where adsorbed + 3− surf [POx ] the soil NH (PO ) pool. PO represents the sum of PO , P is equal to VMAX · surf,P . Taking the time deriva- 4 x x 4 P K +[PO ] 2− − M x HPO4 and H2PO4 that could be utilized by plants and mi- tive leads to the adsorption rate (Wang et al., 2010). + croorganisms, and adsorbed by mineral surfaces: Soil NH4 content is altered by inputs from deposition dep BNF   (FNH ) and biological N2 fixation (F ; Cleveland et al., mob dec 1 gi 4 FN,ij = fij FC,i max − ,0 (A4) 1999): CNi CNj  1 g  1 − e−0.003·NPPannual mob dec i BNF = · FP,ij = fij FC,i max − ,0 , (A5) F 1.8 , (A13) CPi CPj 365 · 86400 where NPP is annual net primary production. Controls where F mob and F mob are the nitrogen and phosphorus gross annual N,ij P,ij on biological N fixation are complex and several mod- mineralization rates. Equations (A4) and (A5) ensure that 2 els have been developed for large-scale land BGC mod- gross mineralization is not less than zero. In contrast, if ni- els (Cleveland et al., 1999; Fisher et al., 2010; Hartwig, trogen (phosphorus) is insufficient, soil microbes immobilize 1998; Parton et al., 1993; Running et al., 1989; Vitousek free NH+ and NO− (PO ): 4 3 x and Field, 1999). However, the emergent responses predicted  g 1  across these model structures are inconsistent (Galloway et immob,pot = dec i − FN,ij fij FC,i max ,0 (A6) al., 2004). Recognizing this important structural uncertainty, CNj CNi BNF we used a simple model where biological N2 fixation (F ) [NH ] immob,pot = immob,pot · 4 is modeled as a function of annual NPP (Cleveland et al., FNH ,ij FN,ij (A7) 4 [NH4] + [NO3] 1999). [NO ] − immob,pot = immob,pot · 3 Soil NO3 content is modified by external deposition in- FNO ,ij FN,ij (A8) dep 3 [NH4] + [NO3] puts (F ) and leaching losses (F leach): NO3 NO3  g 1  immob,pot = dec i − FP,ij fij FC,i max ,0 , (A9) [NO ] CPj CPi leach 3 F = · Qdis, (A14) NO3 W immob,pot immob,pot immob,pot immob,pot where F , F , F and F are − N NH4 NO3 P 2 + − where soil nitrate concentration ([NO3]: gN m ) divided − microbial N, NH4 , NO3 and POx immobilization rates. by soil water content (W: gH O m 2) results in the concen- + − 2 (NH4) and (NO3) are the free NH4 and NO3 pools, respec- tration of dissolved nitrate (DIN). The hydrologic discharge + −2 −1 −1 tively. We assume that microbes have no preference for NH4 (Qdis: gH2O m s ) applied to DIN (gN gH2O ) leads to − −2 −1 or NO3 (Eqs. A7, A8). If soil nutrients are limited, a limi- the leaching loss (gN m s ). tation factor will be applied to those potential soil decompo- Soil POx content is affected by external inputs from parent weather leach sition CNP fluxes (Eqs. A1–A9) to maintain the soil organic material weathering (F ) and leaching losses (FP ). matter CNP stoichiometry. Sorbed P (PS) could be further strongly occluded and become

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 357

Figure A1. MCMC chain. Blue line represents the MCMC samples that are used to infer our model posterior parameters. Two other replicated MCMC calibrations (with different random number seeds) were conducted (yellow and red lines), in order to check the convergence of MCMC calibration.

Figure A2. Gelman–Rubin convergence criterion (solid lines) calculated from three chains in Fig. A1. Baseline value is set to 1.1 (dashed lines). When the Gelman–Rubin criterion is smaller than or equal to 1.1, the chains are thought to converge.

www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 358 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Figure A3. Posterior model parameters (blue bars) fitted to Gaussian distribution (red line). unavailable for plant and microbial uptake. Parent material dep stock can be increased by atmospheric dust deposition (FP ) (Mahowald et al., 2008):

weather F = [PP ]· kweather (A15) [PO ] F leach = x · Q (A16) P W dis occl FP = [PS]· koccl, (A17) where parent material weathering (F weather) is calculated us- ing a weather rate (kweather) and parent material P content ([PP]). POx leaching loss is modeled with a similar approach to nitrate leaching (Eq. A16). The phosphorus occlusion rate is modeled as the product of a constant rate (koccl) and the sorbed P content ([PS]).

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 359

Appendix B: Derivation of VMAX

+ k k The enzyme substrate reaction is S + E ←→1 C −→ P + E, − k1 where the enzyme (E) and substrate (S) reaction is reversible and forms complex (C). The irreversible reaction releases product (P) and liberates enzyme (E). At steady state, the formation rate of the enzyme substrate complex is equal to the consumption rate:

+[ ][ ] = −[ ] + [ ] k1 S E k1 C k C . (B1) To simply the equation, we define an affinity parameter:

k− + k [S]·[E] K = 1 = . (B2) M + [ ] k1 C

By definition, the total enzymes [Etot] in the system is the sum of free enzymes [E] and enzymes that are bound with the substrate [C]:

[Etot] = [E] + [C]. (B3)

Substituting Eq. (B3) into (B2), we have

[S]· ([Etot] − [C]) KM = . (B4) [C] Collecting terms containing [C], we have

[C]· (KM + [S]) = [Etot]·[S]. (B5)

The production rate is d[P] = k ·[C]. (B6) dt Substituting Eq. (B5) into (B6), we have d[P] [S] = k ·[Etot]· . (B7) dt KM + [S] Comparing Eq. (B7) with the classic Michaelis–Menten equation, it is clear that the definition of maximum produc- tion rate is the product of the reaction rate and enzyme abun- dance in the system:

VMAX = k ·[Etot]. (B8)

www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 360 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Acknowledgements. This research was supported by the Direc- Dise, N. B. and Wright, R. F.: Nitrogen leaching from European tor, Office of Science, Office of Biological and Environmental forests in relation to nitrogen deposition, Forest Ecol. Manage., Research of the US Department of Energy under Contract no. 71, 153–161, 1995. DE-AC02-05CH11231 as part of the Regional and Global Climate Drake, J. E., Budynek, A., Hofmockel, K. S., Bernhardt, E. S., Modeling (RGCM) and ACME programs. Billings, S. A., Jackson, R. B., Johnsen, K. S., Lichter, J., Mc- Carthy, H. R., and McCormack, M. L.: Increase in the flux of car- Edited by: A. Rammig bon belowground stimulate nitrogen uptake and sustain the long- term enhancement of forest productivity under elevated CO2, Ecol. Lett., 14, 349–357, 2011. Drtil, M., Nemeth, P., and Bodik, I.: Kinetic constants of nitrifica- tion, Water Res., 27, 35–39, 1993. References Elser, J. J., Bracken, M. E. S., Cleland, E. E., Gruner, D. S., Harpole, W. S., Hillebrand, H., Ngai, J. T., Seabloom, E. W., Shurin, J. B., Anav, A., Friedlingstein, P., Kidston, M., Bopp, L., Ciais, P., Cox, P., and Smith, J. E.: Global analysis of nitrogen and phosphorus lim- Jones, C., Jung, M., Myneni, R., and Zhu, Z.: Evaluating the land itation of primary producers in freshwater, marine and terrestrial and ocean components of the global carbon cycle in the CMIP5 ecosystems, Ecol. Lett., 10, 1135–1142, 2007. Earth System Models, J. Climate, 26, 6801–6843, 2013. Fisher, J. B., Sitch, S., Malhi, Y., Fisher, R. A., Huntingford, C., Barrow, N. J.: The description of phosphate adsorption curves, J. and Tan, S. Y.: Carbon cost of plant nitrogen acquisition: A Soil Sci., 29, 447–462, 1978. mechanistic, globally applicable model of plant nitrogen up- Begum, H. H. and Islam, M. T.: Role of synthesis and exudation of take, retranslocation, and fixation, Global Biogeochem. Cy., 24, organic acids in phosphorus nutrition in plants in tropical soils, GB1014, doi:10.1029/2009GB003621, 2010. Biotechnology, 4, 333–340, 2005. Föhse, D., Claassen, N., and Jungk, A.: Phosphorus efficiency of Bonachela, J. A., Raghib, M., and Levin, S. A.: Dynamic model of plants: I. External and internal P requirement and P uptake effi- flexible phytoplankton nutrient uptake, P. Natl. Acad. Sci. USA, ciency of different plant species, Plant Soil, 110, 101–109, 1988. 108, 20633–20638, 2011. Friedlingstein, P., Cox, P., Betts, R., Bopp, L., Von Bloh, W., Bonan, G. B. and Cleve, K. V.: Soil temperature, nitrogen miner- Brovkin, V., Cadule, P., Doney, S., Eby, M., and Fung, alization, and carbon source-sink relationships in boreal forests, I.: Climate-carbon cycle feedback analysis: Results from the Can. J. Forest Res., 22, 629–639, 1992. C4MIP model intercomparison, J. Climate, 19, 3337–3353, Chauhan, B. S., Stewart, J. W. B., and Paul, E. A.: Effect of labile 2006. inorganic phosphate status and organic carbon additions on the Galloway, J. N., Dentener, F. J., Capone, D. G., Boyer, E. W., microbial uptake of phosphorus in soils, Can. J. Soil Sci., 61, Howarth, R. W., Seitzinger, S. P., Asner, G. P., Cleveland, C. C., 373–385, 1981. Green, P. A., and Holland, E. A.: Nitrogen cycles: past, present, Chaves, M. M., Maroco, J. P., and Pereira, J. S.: Understanding plant and future, Biogeochemistry, 70, 153–226, 2004. responses to drought – from genes to the whole plant, Funct. Hartwig, U. A.: The regulation of symbiotic N2 fixation: a concep- Plant Biol., 30, 239–264, 2003. tual model of N feedback from the ecosystem to the gene expres- Chen, M.: Kinetics of phosphorus absorption byCorynebacterium sion level, Perspect. Plant Ecol. Evol. Syst., 1, 92–120, 1998. bovis, Microb. Ecol., 1, 164–175, 1974. Hobbie, J. E. and Hobbie, E. A.: 15N in symbiotic fungi and plants Cogliatti, D. H. and Clarkson, D. T.: Physiological changes in, and estimates nitrogen and carbon flux rates in arctic tundra, , phosphate uptake by potato plants during development of, and re- 87, 816–822, 2006. covery from phosphate deficiency, Physiol. Plant., 58, 287–294, Hodge, A. and Fitter, A. H.: Substantial nitrogen acquisition by ar- 1983. buscular mycorrhizal fungi from organic material has implica- Colpaert, J. V., Van Tichelen, K. K., Van Assche, J. A., and Van tions for N cycling, P. Natl. Acad. Sci. USA, 107, 13754–13759, Laere, A.: Short-term phosphorus uptake rates in mycorrhizal 2010. and non-mycorrhizal roots of intact Pinus sylvestris seedlings, Hodge, A., Robinson, D., and Fitter, A.: Are microorganisms more New Phytol., 143, 589–597, 1999. effective than plants at competing for nitrogen?, Trends Plant Cleveland, C. C., Townsend, A. R., Schimel, D. S., Fisher, H., Sci., 5, 304–308, 2000a. Howarth, R. W., Hedin, L. O., Perakis, S. S., Latty, E. F., Von Hodge, A., Stewart, J., Robinson, D., Griffiths, B. S., and Fitter, Fischer, J. C., and Elseroad, A.: Global patterns of terrestrial A. H.: Competition between roots and soil micro-organisms for biological nitrogen (N2) fixation in natural ecosystems, Global nutrients from nitgrogen-rich patches of varying complexity, J. Biogeochem. Cy., 13, 623–645, 1999. Ecol., 88, 150–164, 2000b. Del Grosso, S. J., Parton, W. J., Mosier, A. R., Ojima, D. S., Kul- Houghton, R. A.: Revised estimates of the annual net flux of carbon mala, A. E., and Phongpan, S.: General model for N2O and to the atmosphere from changes in land use and land manage- N2 gas emissions from soils due to dentrification, Global Bio- ment 1850–2000, Tellus B, 55, 378–390, 2003. geochem. Cy., 14, 1045–1060, 2000. Hu, S., Chapin, F. S., Firestone, M. K., Field, C. B., and Chiariello, DeLuca, T. H., Zackrisson, O., Gundale, M. J., and Nilsson, M.- N. R.: Nitrogen limitation of microbial decomposition in a grass- C.: Ecosystem feedbacks and nitrogen fixation in boreal forests, land under elevated CO2, Nature, 409, 188–191, 2001. Science, 320, 1181–1181, 2008. Iversen, C. M., Hooker, T. D., Classen, A. T., and Norby, R. J.: Net DeLuca, T. H., Zackrisson, O., Nilsson, M.-C., and Sellstedt, A.: mineralization of N at deeper soil depths as a potential mecha- Quantifying nitrogen-fixation in feather moss carpets of boreal forests, Nature, 419, 917–920, 2002.

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 361

nism for sustained forest production under elevated [CO2], Glob. fication in soils, J. Geophys. Res.-Biogeo., 114, G04012, Change Biol., 17, 1130–1139, 2011. doi:10.1029/2008JG000878, 2009. Jackson, R. B., Mooney, H. A., and Schulze, E. D.: A global budget Maggi, F., Gu, C., Riley, W. J., Hornberger, G. M., Venterea, for fine root biomass, surface area, and nutrient contents, P. Natl. R. T., Xu, T., Spycher, N., Steefel, C., Miller, N. L., and Acad. Sci. USA, 94, 7362–7366, 1997. Oldenburg, C. M.: A mechanistic treatment of the domi- Jackson, R. B., Cook, C. W., Pippen, J. S., and Palmer, S. M.: In- nant soil nitrogen cycling processes: Model development, test- creased belowground biomass and soil CO2 fluxes after a decade ing, and application, J. Geophys. Res.-Biogeo., 113, G02016, of enrichment in a warm-temperate forest, Ecol- doi:10.1029/2007JG000578, 2008. ogy, 90, 3352–3366, 2009. Mahowald, N., Jickells, T. D., Baker, A. R., Artaxo, P., Benitez- Ji, D., Wang, L., Feng, J., Wu, Q., Cheng, H., Zhang, Q., Yang, Nelson, C. R., Bergametti, G., Bond, T. C., Chen, Y., Cohen, J., Dong, W., Dai, Y., Gong, D., Zhang, R.-H., Wang, X., Liu, D. D., and Herut, B.: Global distribution of atmospheric phos- J., Moore, J. C., Chen, D., and Zhou, M.: Description and ba- phorus sources, concentrations and deposition rates, and an- sic evaluation of Beijing Normal University Earth System Model thropogenic impacts, Global Biogeochem. Cy., 22, GB4026, (BNU-ESM) version 1, Geosci. Model Dev., 7, 2039–2064, doi:10.1029/2008GB003240, 2008. doi:10.5194/gmd-7-2039-2014, 2014. Marion, G. M., Miller, P. C., Kummerow, J., and Oechel, W. C.: Johnson, D. W.: Nitrogen retention in forest soils, J. Environ. Qual., Competition for nitrogen in a tussock tundra ecosystem, Plant 21, 1–12, 1992. Soil, 66, 317–327, 1982. Kaspari, M., Garcia, M. N., Harms, K. E., Santana, M., Wright, S. Marland, G., Boden, T. A., and Andres, R. J.: Global, Regional, and J., and Yavitt, J. B.: Multiple nutrients limit litterfall and decom- National Fossil Fuel CO2 Emissions, in Trends: A Compendium position in a tropical forest, Ecol. Lett., 11, 35–43, 2008. of Data on Global Change, Carbon Dioxide Information Anal- Kaye, J. P. and Hart, S. C.: Competition for nitrogen between plants ysis Center, Oak Ridge National Laboratory, US Department of and soil microorganisms, Trends Ecol. Evol., 12, 139–143, 1997. Energy, Oak Ridge, Tenn., USA, 2008 Korsaeth, A., Molstad, L., and Bakken, L. R.: Modelling the com- Matson, P., McDowell, W. H., Townsend, A. R., and Vitousek, P. petition for nitrogen between plants and microflora as a function M.: The globalization of N deposition: ecosystem consequences of soil heterogeneity, Soil Biol. Biochem., 33, 215–226, 2001. in tropical environments, Biogeochemistry, 46, 67–83, 1999. Koven, C. D., Riley, W. J., Subin, Z. M., Tang, J. Y., Torn, M. S., Matson, P., Lohse, K. A., and Hall, S. J.: The globalization of ni- Collins, W. D., Bonan, G. B., Lawrence, D. M., and Swenson, trogen deposition: consequences for terrestrial ecosystems, AM- S. C.: The effect of vertically resolved soil biogeochemistry and BIO: A Journal of the Human Environment, 31, 113–119, 2002. alternate soil C and N models on C dynamics of CLM4, Biogeo- McGroddy, M. E., Silver, W. L., De Oliveira, R. C., De Mello, W. sciences, 10, 7109–7131, doi:10.5194/bg-10-7109-2013, 2013. Z., and Keller, M.: Retention of phosphorus in highly weathered Kuzyakov, Y. and Xu, X.: Competition between roots and microor- soils under a lowland Amazonian forest ecosystem, J. Geophys. ganisms for nitrogen: mechanisms and ecological relevance, Res.-Biogeo., 113, GO4012, doi:10.1029/2008JG000756, 2008. New Phytol., 198, 656–669, 2013. McGuire, A. D., Melillo, J. M., Joyce, L. A., Kicklighter, D. W., Lambers, H., Mougel, C., Jaillard, B., and Hinsinger, P.: Plant- Grace, A. L., Moore, B., and Vorosmarty, C. J.: Interactions be- microbe-soil interactions in the rhizosphere: an evolutionary per- tween carbon and nitrogen dynamics in estimating net primary spective, Plant Soil, 321, 83–115, 2009. productivity for potential vegetation in North America, Global Le Quéré, C., Raupach, M. R., Canadell, J. G., and Marland, G.: Biogeochem. Cy., 6, 101–124, 1992. Trends in the sources and sinks of carbon dioxide, Nat. Geosci., Medvigy, D., Wofsy, S. C., Munger, J. W., Hollinger, D. Y., 2, 831–836, 2009. and Moorcroft, P. R.: Mechanistic scaling of ecosystem func- Le Quéré, C., Andres, R. J., Boden, T., Conway, T., Houghton, R. tion and dynamics in space and time: Ecosystem Demogra- A., House, J. I., Marland, G., Peters, G. P., van der Werf, G. R., phy model version 2, J. Geophys. Res.-Biogeo., 114, G01002, Ahlström, A., Andrew, R. M., Bopp, L., Canadell, J. G., Ciais, doi:10.1029/2008JG000812, 2009. P., Doney, S. C., Enright, C., Friedlingstein, P., Huntingford, C., Min, X., Siddiqi, M. Y., Guy, R. D., Glass, A. D. M., and Kro- Jain, A. K., Jourdain, C., Kato, E., Keeling, R. F., Klein Gold- nzucker, H. J.: A comparative kinetic analysis of nitrate and am- ewijk, K., Levis, S., Levy, P., Lomas, M., Poulter, B., Raupach, monium influx in two early-successional tree species of temper- M. R., Schwinger, J., Sitch, S., Stocker, B. D., Viovy, N., Zaehle, ate and boreal forest ecosystems, Plant Cell Environ., 23, 321– S., and Zeng, N.: The global carbon budget 1959–2011, Earth 328, 2000. Syst. Sci. Data, 5, 165–185, doi:10.5194/essd-5-165-2013, 2013. Moorcroft, P. R., Hurtt, G. C., and Pacala, S. W.: A method for LeBauer, D. S. and Treseder, K. K.: Nitrogen limitation of net scaling vegetation dynamics: the ecosystem demography model primary productivity in terrestrial ecosystems is globally dis- (ED), Ecol. Monogr., 71, 557–586, 2001. tributed, Ecology, 89, 371–379, 2008. Moorhead, D. L. and Sinsabaugh, R. L.: A theoretical model of lit- Li, C., Aber, J., Stange, F., Butterbach-Bahl, K., and Papen, H.: ter decay and microbial interaction, Ecol. Monogr., 76, 151–174, A process-oriented model of N2O and NO emissions from for- 2006. est soils: 1. Model development, J. Geophys. Res.-Atmos., 105, Mooshammer, M., Wanek, W., Zechmeister-Boltenstern, S., and 4369–4384, 2000. Richter, A.: Stoichiometric imbalances between terrestrial de- Maggi, F. and Riley, W. J.: Transient competitive complexa- composer communities and their resources: mechanisms and im- tion in biological kinetic isotope fractionation explains non- plications of microbial adaptations to their resources, Front. Mi- steady isotopic effects: Theory and application to denitri- crobiol., 5, 22, doi:10.3389/fmicb.2014.00022, 2014.

www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016 362 Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces

Murray, R. E., Parsons, L. L., and Smith, M. S.: Kinetics of nitrate Raynaud, X. and Leadley, P. W.: Soil characteristics play a key role utilization by mixed populations of denitrifying bacteria, Appl. in modeling nutrient competition in plant communities, Ecology, Environ. Microbiol., 55, 717–721, 1989. 85, 2200–2214, 2004. Nedwell, D. B.: Effect of low temperature on microbial growth: Raynaud, X., Lata, J.-C., and Leadley, P. W.: Soil lowered affinity for substrates limits growth at low temperature, and nutrient uptake by plants: a test using a coupled C: N model FEMS Microbiol. Ecol., 30, 101–111, 1999. of plant–microbial interactions, Plant Soil, 287, 95–116, 2006. Norby, R. J., Ledford, J., Reilly, C. D., Miller, N. E., and O’Neill, Reich, P. B. and Hobbie, S. E.: Decade-long soil nitrogen con- E. G.: Fine-root production dominates response of a deciduous straint on the CO2 fertilization of plant biomass, Nature Climate forest to atmospheric CO2 enrichment, P. Natl. Acad. Sci. USA, Change, 3, 278–282, 2013. 101, 9689–9693, 2004. Reich, P. B., Hobbie, S. E., Lee, T., Ellsworth, D. S., West, J. B., Norby, R. J., Warren, J. M., Iversen, C. M., Medlyn, B. E., and Tilman, D., Knops, J. M. H., Naeem, S., and Trost, J.: Nitro- McMurtrie, R. E.: CO2 enhancement of forest productivity con- gen limitation constrains sustainability of ecosystem response to strained by limited nitrogen availability, P. Natl. Acad. Sci. USA, CO2, Nature, 440, 922–925, 2006. 107, 19368–19373, 2010. Reynolds, H. L. and Pacala, S. W.: An analytical treatment of root- Nordin, A., Högberg, P., and Näsholm, T.: Soil nitrogen form and to-shoot ratio and plant competition for soil nutrient and light, plant nitrogen uptake along a boreal forest productivity gradient, Am. Nat., 141, 51–70, 1993. Oecologia, 129, 125–132, 2001. Ricciuto, D. M., Davis, K. J., and Keller, K.: A Bayesian calibration Olander, L. P. and Vitousek, P. M.: Biological and geochemical of a simple carbon cycle model: The role of observations in es- sinks for phosphorus in soil from a wet tropical forest, Ecosys- timating and reducing uncertainty, Global Biogeochem. Cy., 22, tems, 7, 404–419, 2004. GB2030, doi:10.1029/2006GB002908, 2008. Olander, L. P. and Vitousek, P. M.: Short-term controls over inor- Rillig, M. C., Allen, M. F., Klironomos, J. N., Chiariello, N. R., ganic phosphorus during soil and ecosystem development, Soil and Field, C. B.: Plant species-specific changes in root-inhabiting Biol. Biochem., 37, 651–659, 2005. fungi in a California annual grassland: responses to elevated CO2 Oren, R., Ellsworth, D. S., Johnsen, K. H., Phillips, N., Ewers, B. and nutrients, Oecologia, 113, 252–259, 1998. E., Maier, C., Schäfer, K. V. R., McCarthy, H., Hendrey, G., and Running, S. W. and Coughlan, J. C.: A general model of forest McNulty, S. G.: Soil fertility limits by forest ecosystem processes for regional applications I. Hydrologic bal- ecosystems in a CO2-enriched atmosphere, Nature, 411, 469– ance, canopy gas exchange and primary production processes, 472, 2001. Ecol. Model., 42, 125–154, 1988. Pappas, C., Fatichi, S., Leuzinger, S., Wolf, A., and Burlando, Running, S. W., Nemani, R. R., Peterson, D. L., Band, L. E., Potts, P.: Sensitivity analysis of a process-based : D. F., Pierce, L. L., and Spanner, M. A.: Mapping regional forest Pinpointing parameterization and structural issues, J. Geophys. evapotranspiration and photosynthesis by coupling satellite data Res.-Biogeo., 118, 505–528, 2013. with ecosystem simulation, Ecology, 70, 1090–1101, 1989. Parton, W. J., Stewart, J. W. B., and Cole, C. V.: Dynamics of C, Schimel, J. P. and Bennett, J.: Nitrogen mineralization: challenges N, P and S in grassland soils: a model, Biogeochemistry, 5, 109– of a changing paradigm, Ecology, 85, 591–602, 2004. 131, 1988. Schimel, J. P., Jackson, L. E., and Firestone, M. K.: Spatial and Parton, W. J., Scurlock, J. M. O., Ojima, D. S., Gilmanov, T. G., Sc- temporal effects on plant-microbial competition for inorganic ni- holes, R. J., Schimel, D. S., Kirchner, T., Menaut, J. C., Seastedt, trogen in a California annual grassland, Soil Biol. Biochem., 21, T., and Garcia Moya, E.: Observations and modeling of biomass 1059–1066, 1989. and soil organic matter dynamics for the grassland biome world- Scholze, M., Kaminski, T., Rayner, P., Knorr, W., and Giering, R.: wide, Global Biogeochem. Cy., 7, 785–809, 1993. Propagating uncertainty through prognostic carbon cycle data Parton, W. J., Holland, E. A., Del Grosso, S. J., Hartman, M. D., assimilation system simulations, J. Geophys. Res.-Atmos., 112, Martin, R. E., Mosier, A. R., Ojima, D. S., and Schimel, D. S.: D17305, doi:10.1029/2007JD008642, 2007. Generalized model for NOx and N2O emissions from soils, J. Shen, J., Yuan, L., Zhang, J., Li, H., Bai, Z., Chen, X., Zhang, W., Geophys. Res.-Atmos., 106, 17403–17419, 2001. and Zhang, F.: Phosphorus dynamics: from soil to plant, Plant Perakis, S. S. and Hedin, L. O.: Nitrogen loss from unpolluted South Physiol., 156, 997–1005, 2011. American forests mainly via dissolved organic compounds, Na- Silver, W. L., Thompson, A. W., McGroddy, M. E., Varner, R. ture, 415, 416–419, 2002. K., Robertson, J. R., J.D. Dias, H. S., Crill, P., and Keller, Phillips, R. P., Finzi, A. C., and Bernhardt, E. S.: Enhanced root M.: LBA-ECO TG-07 Long-Term Soil Gas Flux and Root exudation induces microbial feedbacks to N cycling in a pine Mortality, Tapajos National Forest, Data set, available at: forest under long-term CO2 fumigation, Ecol. Lett., 14, 187–194, http://daac.ornl.gov from Oak Ridge National Laboratory Dis- 2011. tributed Active Archive Center, Oak Ridge, Tennessee, USA, Potter, C. S., Randerson, J. T., Field, C. B., Matson, P. A., Vitousek, doi:10.3334/ORNLDAAC/1116. 2012. P. M., Mooney, H. A., and Klooster, S. A.: Terrestrial ecosystem Sokolov, A. P., Kicklighter, D. W., Melillo, J. M., Felzer, B. S., production: a process model based on global satellite and surface Schlosser, C. A., and Cronin, T. W.: Consequences of consid- data, Global Biogeochem. Cy., 7, 811–841, 1993. ering carbon-nitrogen interactions on the feedbacks between cli- Provides, G. F. W. I.: Climate Model Intercomparisons: Preparing mate and the terrestrial carbon cycle, J. Climate, 21, 3776–3796, for the Next Phase, Eos, 95, 77, 2014. 2008. Springate, D. A. and Kover, P. X.: Plant responses to elevated tem- peratures: a field study on phenological sensitivity and fitness re-

Biogeosciences, 13, 341–363, 2016 www.biogeosciences.net/13/341/2016/ Q. Zhu et al.: Multiple soil nutrient competition between plants, microbes, and mineral surfaces 363

sponses to simulated climate warming, Glob. Change Biol., 20, Waksman, S. A.: Principles of soil microbiology, Principles Soil 456–465, 2014. Microbiol, 2nd edition, Williams and Wilkins, Baltimore, Mary- Stocker, T. F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S. land, US., 1931. K., Boschung, J., Nauels, A., Xia, Y., Bex, V., and Midg- Walker, T. W. and Syers, J. K.: The fate of phosphorus during pedo- ley, P. M.: Climate Change 2013, The Physical Science Ba- genesis, Geoderma, 15, 1–19, 1976. sis. Working Group I Contribution to the Fifth Assessment Re- Wang, J. and Lars, B., R: Competition for nitrogen during mineral- port of the Intergovernmental Panel on Climate Change-Abstract ization of plant residues in soil: microbial response to C and N for decision-makersRep, Groupe d’experts intergouvernemental availability, Soil Biol. Biochem., 29, 163–170, 1997. sur l’evolution du climat/Intergovernmental Panel on Climate Wang, Y. P., Houlton, B. Z., and Field, C. B.: A model of biogeo- Change-IPCC, C/O World Meteorological Organization, 7bis chemical cycles of carbon, nitrogen, and phosphorus including Avenue de la Paix, CP 2300 CH-1211 Geneva 2, Switzerland, symbiotic nitrogen fixation and phosphatase production, Global 2013. Biogeochem. Cy., 21, GB1018, doi:10.1029/2006GB002797, Tang, J. Y. and Riley, W. J.: A total quasi-steady-state formulation 2007. of substrate uptake kinetics in complex networks and an example Wang, Y. P., Law, R. M., and Pak, B.: A global model of carbon, application to microbial litter decomposition, Biogeosciences, nitrogen and phosphorus cycles for the terrestrial biosphere, Bio- 10, 8329–8351, doi:10.5194/bg-10-8329-2013, 2013. geosciences, 7, 2261–2282, doi:10.5194/bg-7-2261-2010, 2010. Tang, J. Y. and Riley, W. J.: Weaker soil carbon-climate feedbacks Wieder, W. R., Cleveland, C. C., and Townsend, A. R.: Controls resulting from microbial and abiotic interactions, Nature Climate over leaf litter decomposition in wet tropical forests, Ecology, Change, 5, 56–60, 2014. 90, 3333–3341, 2009. Templer, P. H., Silver, W. L., Pett-Ridge, J., M. DeAngelis, K., and Woodmansee, R. G., Vallis, I., and Mott, J. J.: Grassland nitrogen, Firestone, M. K.: Plant and microbial controls on nitrogen reten- Ecological Bulletins (Sweden), 1981. tion and loss in a humid tropical forest, Ecology, 89, 3030–3040, Xu, X., Thornton, P. E., and Post, W. M.: A global analysis of soil 2008. microbial biomass carbon, nitrogen and phosphorus in terrestrial Thomas, R. Q., Bonan, G. B., and Goodale, C. L.: Insights ecosystems, Global Ecol. Biogeogr., 22, 737–749, 2013. into mechanisms governing forest carbon response to nitro- Yang, X., Thornton, P. E., Ricciuto, D. M., and Post, W. M.: gen deposition: a model-data comparison using observed re- The role of phosphorus dynamics in tropical forests – a mod- sponses to nitrogen addition, Biogeosciences, 10, 3869–3887, eling study using CLM-CNP, Biogeosciences, 11, 1667–1681, doi:10.5194/bg-10-3869-2013, 2013a. doi:10.5194/bg-11-1667-2014, 2014. Thomas, R. Q., Zaehle, S., Templer, P. H., and Goodale, C. L.: Zaehle, S. and Dalmonech, D.: Carbon–nitrogen interactions on Global patterns of nitrogen limitation: confronting two global land at global scales: current understanding in modelling climate biogeochemical models with observations, Glob. Change Biol., biosphere feed 19, 2986–2998, 2013b. Zaehle, S. and Friend, A. D.: Carbon and dynamics Thornton, P. E., Lamarque, J. F., Rosenbloom, N. A., and in the O-CN land surface model: 1. Model description, site-scale Mahowald, N. M.: Influence of carbon-nitrogen cycle cou- evaluation, and sensitivity to parameter estimates, Global Bio- pling on land model response to CO2 fertilization and geochem. Cy., 24, GB1005, doi:10.1029/2009GB003521, 2010. climate variability, Global Biogeochem. Cy., 21, GB4018, Zaehle, S., Friedlingstein, P., and Friend, A. D.: Terrestrial nitrogen doi:10.1029/2006GB002868, 2007. feedbacks may accelerate future climate change, Geophys. Res. Treseder, K. K. and Vitousek, P. M.: Effects of soil nutrient avail- Lett., 37, L01401, doi:10.1029/2009GL041345, 2010. ability on investment in acquisition of N and P in Hawaiian rain Zaehle, S., Medlyn, B. E., De Kauwe, M. G., Walker, A. P., Di- forests, Ecology, 82, 946–954, 2001. etze, M. C., Hickler, T., Luo, Y., Wang, Y. P., El-Masri, B., and Trumbore, S., Da Costa, E. S., Nepstad, D. C., Barbosa De Ca- Thornton, P.: Evaluation of 11 terrestrial carbon–nitrogen cycle margo, P., Martinelli, L. A., Ray, D., Restom, T., and Silver, W.: models against observations from two temperate Free-Air CO2 Dynamics of fine root carbon in Amazonian tropical ecosystems Enrichment studies, New Phytol., 202, 803–822, 2014. and the contribution of roots to , Glob. Change Zhang, Q., Wang, Y. P., Pitman, A. J., and Dai, Y. J.: Limi- Biol., 12, 217–229, 2006. tations of nitrogen and phosphorous on the terrestrial carbon Vitousek, P. M. and Farrington, H.: Nutrient limitation and soil de- uptake in the 20th century, Geophys. Res. Lett., 38, L22701, velopment: experimental test of a biogeochemical theory, Bio- doi:10.1029/2011GL049244, 2011. geochemistry, 37, 63–75, 1997. Zhu, Q. and Riley, W. J.: Improved modelling of soil nitrogen Vitousek, P. M. and Howarth, R. W.: Nitrogen limitation on land losses, Nature Climate Change, 5, 705–706, 2015. and in the sea: how can it occur?, Biogeochemistry, 13, 87–115, Zhu, Q. and Zhuang, Q.: Modeling the effects of organic nitrogen 1991. uptake by plants on the carbon cycling of boreal forest and tundra Vitousek, P. M. and Field, C. B.: Ecosystem constraints to symbi- ecosystems, Biogeosciences, 10, 7943–7955, doi:10.5194/bg- otic nitrogen fixers: a simple model and its implications, Biogeo- 10-7943-2013, 2013. chemistry, 46, 179–202, 1999. Zhu, Q. and Zhuang, Q.: Parameterization and sensitivity anal- Vitousek, P. M. and Sanford, R. L.: Nutrient cycling in moist tropi- ysis of a process-based terrestrial ecosystem model us- cal forest, Annu. Rev. Ecol. Syst., 17, 137–167, 1986. ing adjoint method, J. Adv. Model. Earth Syst., 315–331, Vitousek, P. M., Porder, S., Houlton, B. Z., and Chadwick, O. A.: doi:10.1002/2013MS000241, 2014. Terrestrial phosphorus limitation: mechanisms, implications, and nitrogen-phosphorus interactions, Ecol. Appl., 20, 5–15, 2010.

www.biogeosciences.net/13/341/2016/ Biogeosciences, 13, 341–363, 2016