Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/ doi:10.5194/gmd-9-927-2016 Author(s) 2016. CC Attribution 3.0 License.
Guoping Tang1, Fengming Yuan1, Gautam Bisht1,2, Glenn E. Hammond3,4, Peter C. Lichtner5, Jitendra Kumar1, Richard T. Mills6, Xiaofeng Xu1,7, Ben Andre2,8, Forrest M. Hoffman1, Scott L. Painter1, and Peter E. Thornton1
1Oak Ridge National Laboratory, Oak Ridge, Tennessee, USA
2Lawrence Berkeley National Laboratory, Berkeley, California, USA
3Pacic Northwest National Laboratory, Richland, Washington, USA
4Sandia National Laboratories, Albuquerque, New Mexico, USA
5OFM ResearchSouthwest, Santa Fe, New Mexico, USA
6Intel Corporation, Hillsboro, Oregon, USA
7San Diego State University, San Diego, California, USA
8National Center for Atmospheric Research, Boulder, Colorado, USA
Correspondence to: P. E. Thornton ([email protected]) and S. L. Painter ([email protected])
Received: 18 November 2015 Published in Geosci. Model Dev. Discuss.: 17 December 2015 Revised: 13 February 2016 Accepted: 17 February 2016 Published: 4 March 2016
Abstract. We explore coupling to a congurable subsur-face reactive transport code as a exible and extensible approach to biogeochemistry in land surface models. A reaction network with the Community Land Model carbon nitrogen (CLM-CN) decomposition, nitrication, denitrication, and plant uptake is used as an example. We implement the reactions in the open-source PFLOTRAN (massively parallel subsurface ow and reactive transport) code and couple it with the CLM. To make the rate formulae designed for use in explicit time stepping in CLMs compatible with the implicit time stepping used in PFLOTRAN, the Monod substrate rate-limiting function with a residual concentration is used to represent the limitation of nitrogen availability on plant uptake and immobilization. We demonstrate that CLMPFLOTRAN predictions (without invoking PFLOTRAN transport) are consistent with CLM4.5 for Arctic, temperate, and tropical sites.
Switching from explicit to implicit method increases rigor but introduces numerical challenges. Care needs to be taken to use scaling, clipping, or log transformation to avoid negative concentrations during the Newton iterations. With a tight relative update tolerance (STOL) to avoid false convergence, an accurate solution can be achieved with about 50 % more
computing time than CLM in point mode site simulations using either the scaling or clipping methods. The log transformation method takes 60100 % more computing time than CLM. The computing time increases slightly for clipping and scaling; it increases substantially for log transformation for half saturation decrease from 103 to 109 molm3, which normally results in decreasing nitrogen concentrations. The frequent occurrence of very low concentrations (e.g. below nanomolar) can increase the computing time for clipping or scaling by about 20 %, double for log transformation. Overall, the log transformation method is accurate and robust, and the clipping and scaling methods are efcient. When the reaction network is highly nonlinear or the half saturation or residual concentration is very low, the allowable time-step cuts may need to be increased for robustness for the log transformation method, or STOL may need to be tightened for the clipping and scaling methods to avoid false convergence.
As some biogeochemical processes (e.g., methane and nitrous oxide reactions) involve very low half saturation and thresholds, this work provides insights for addressing non-physical negativity issues and facilitates the representation of a mechanistic biogeochemical description in Earth system models to reduce climate prediction uncertainty.
Published by Copernicus Publications on behalf of the European Geosciences Union.
Addressing numerical challenges in introducing a reactive transport code into a land surface model: a biogeochemical modeling proof-of-concept with CLMPFLOTRAN 1.0
928 G. Tang et al.: CLMPFLOTRAN biogeochemistry
1 Introduction
Land surface (terrestrial ecosystem) models (LSMs) calculate the uxes of energy, water, and greenhouse gases across the landatmosphere interface for the atmospheric general circulation models for climate simulation and weather forecasting (Sellers et al., 1997). Evolving from the rst generation bucket, second generation biophysical, and third generation physiological models (Seneviratne et al., 2010), current LSMs such as the Community Land Model (CLM) implement comprehensive thermal, hydrological, and biogeochemical processes (Oleson et al., 2013). The importance of accurate representation of soil biogeochemistry in LSMs is suggested by the conrmation that the increase of carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) in the atmosphere since preindustrial time is the main driving cause of climate change, and interdependent water, carbon and nitrogen cycles in terrestrial ecosystems are very sensitive to climate changes (IPCC, 2013). In addition to 250
soil biogeochemical models developed in the past 80 years
(Manzoni and Porporato, 2009), increasingly mechanistic models continue to be developed to increase the delity of process representation for improving climate prediction (e.g., Riley et al., 2014).
As LSMs usually hardcode the soil biogeochemistry reaction network (pools/species, reactions, rate formulae), substantial effort is often required to modify the source code for testing alternative models and incorporating new process understanding. Fang et al. (2013) demonstrated the use of a reaction-based approach to facilitate implementation of the Community Land Model carbonnitrogen (CLM-CN) and CENTURY models and incorporation of the phosphorus cycle. Tang et al. (2013) solved the advection diffusion equation in CLM using operator splitting. In contrast, TOUGH-REACT, a reactive transport modeling (RTM) code, was used to develop multiphase mechanistic carbon and nitrogen models with many speciation and microbial reactions (Maggi et al., 2008; Gu and Riley, 2010; Riley et al., 2014), but it has not been coupled to an LSM. PHREEQC was coupled with DayCent to describe soil and stream water equilibrium chemistry (Hartman et al., 2007). Coupling a RTM code with CLM will facilitate testing of increasingly mechanistic biogeochemical models in LSMs.
An essential aspect of LSMs is to simulate competition for nutrients (e.g., mineral nitrogen, phosphorus) among plants and microbes. In CLMs, plant and immobilization nitrogen demands are calculated independent of soil mineral nitrogen.The limitation of nitrogen availability on plant uptake and immobilization is simulated by a demand-based competition: demands are downregulated by soil nitrogen concentration (Oleson et al., 2013; Thornton and Rosenbloom, 2005). This avoids negative concentrations and does not introduce mass balance errors (Tang and Riley, 2016) as CLM uses explicit time stepping.
RTMs often account for limitation of reactant availability on reaction rates for each individual reaction to allow mechanistic representations and exibility in adding reactions. For maximum robustness, some RTM codes support the use of fully implicit time stepping, usually employing a variant of the NewtonRaphson method. Negative concentration can be introduced during NewtonRaphson iterations, which is not physical and can cause numerical instability and errors (Shampine et al., 2005). In our work with CLMs, this is expected to worsen when we implement microbial reactions for methane and nitrous oxide consumption and production as the threshold, and half saturation are at or below nanomolar (109 M) level (Conrad, 1996). The redox potential (Eh) needs to be decreased to 0.35 V (oxygen concen
tration < 1022 M; Hungate, 1975) for methanogens to grow (Jarrell, 1985).
Three methods are used to avoid negative concentration in RTM codes. One is to use the logarithm concentration as the primary variable (Bethke, 2007; Hammond, 2003; Parkhurst and Appelo, 1999). The other two either scale back the update vector (Bethke, 2007; Hammond, 2003) or clip the concentrations for the species that are going negative (Yeh et al., 2004; White and McGrail, 2005; Sonnenthal et al., 2014) in each iteration. Except that log transformation is more computationally demanding (Hammond, 2003), how these methods for enforcing non-negativity affect computational accuracy and efciency is rarely discussed.
As LSMs need to run under various conditions at the global scale for simulation duration of centuries, it is necessary to resolve accuracy and efciency issues to use RTM codes for LSMs. The objective of this work is to explore some of the implementation issues associated with using RTM codes in LSMs, with the ultimate goal being accurate, efcient, robust, and congurable representations of sub-surface biogeochemical reactions in CLM. To this end, we develop an alternative implementation of an existing CLM biogeochemical reaction network using PFLOTRAN (massively parallel subsurface ow and reactive transport) (Lichtner et al., 2015; Hammond et al., 2014); couple that model to CLM (we refer to the coupled model as CLMPFLOTRAN); test the implementation at Arctic, temperate, and tropical sites; and examine the implication of using scaling, clipping, and log transformation for avoiding negative concentrations.Although we focus on a number of carbonnitrogen reactions implemented in PFLOTRAN and CLM, the critical numerical issue of avoiding negative concentrations has broader relevance as biogeochemical representations in LSMs become more mechanistic.
2 Methods
Among the many reactions in LSMs are the soil biogeochemical reactions for carbon and nitrogen cycles, in particular the organic matter decomposition, nitrication, denitrica-
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 929
tion, plant nitrogen uptake, and methane production and oxidation. The kinetics are usually described by a rst-order rate modied by response functions for environmental variables (temperature, moisture, pH, etc.) (Bonan et al., 2012; Boyer et al., 2006; Schmidt et al., 2011). In this work, we use a network of the CLM-CN decomposition (Bonan et al., 2012;Oleson et al., 2013; Thornton and Rosenbloom, 2005), nitrication, denitrication (Dickinson et al., 2002; Parton et al., 2001, 1996), and plant nitrogen uptake reactions (Fig. 1) as an example. The reactions and rate formulae are detailed in Appendix A.
2.1 CLMPFLOTRAN biogeochemistry coupling
In CLMPFLOTRAN, CLM can instruct PFLOTRAN to solve the partial differential equations for energy (including freezing and thawing), water ow, and reaction and transport in the surface and subsurface. This work focuses on the PFLOTRAN biogeochemistry, with the CLM solving the energy and water ow equations and handling the solute transport (mixing, advection, diffusion, and leaching). Here, we focus on how reactions are implemented and thus only use PFLOTRAN in batch mode (i.e. without transport). However, PFLOTRANs advection and diffusion capabilities are operational in the CLMPFLOTRAN coupling described here.In each CLM time step, the CLM provides production rates for Lit1C, Lit1N, Lit2C, Lit2N, Lit3C, Lit3N for litter fall;CWDC and CWDN for coarse woody debris production, nitrogen deposition and xation, and plant nitrogen demand; and species liquid water content, matrix potential, and temperature for PFLOTRAN. PFLOTRAN solves ordinary differential equations for the kinetic reactions and mass action equations for equilibrium reactions and provides the nal concentrations back to CLM.
The reactions and rates are implemented using the reaction sandbox concept in PFLOTRAN (Lichtner et al., 2015).For each reaction, we specify a rate and a derivative of the rate with respect to any components in the rate formula, given concentrations, temperature, moisture content, and other environmental variables (see reaction_sandbox_example.F90 in potran-dev source code for details). PFLOTRAN accumulates these rates and derivatives into a residual vector and a Jacobian matrix, and the global equation is discretized in time using the backward Euler method; the resulting system of algebraic equations is solved using the NewtonRaphson method (Appendix B). A Krylov subspace method is usually employed to solve the Jacobian systems arising during the NewtonRaphson iterations, but, as the problems are relatively small in this study, we solve them directly via LU (lower upper) factorization.
Unlike the explicit time stepping in CLM, in which only reaction rates need to be calculated, implicit time stepping requires evaluating derivatives. While PFLOTRAN provides an option to calculate derivatives numerically via nite-
differencing, we use analytical expressions for efciency and accuracy.
Many reactions can be specied in an input le, providing exibility in adding various reactions with user-dened rate formulae. As typical rate formulae consist of rst-order, Monod, and inhibition terms, a general rate formula with a exible number of terms and typical moisture, temperature, and pH response functions is coded in PFLOTRAN. Most of the biogeochemical reactions can be specied in an input le, with a exible number of reactions, species, rate terms, and various response functions without source code modication. Code modication is necessary only when different rate formulae or response functions are introduced. In contrast, the pools and reactions are traditionally hardcoded in CLM. Consequently, any change of the pools, reactions, or rate formula may require source code modication. Therefore, the more general approach used by PFLOTRAN facilitates implementation of increasingly mechanistic reactions and tests of various representations with less code modication.
2.2 Mechanistic representation of rate-limiting processes
To use RTMs in LSMs, we need to make reaction networks designed for use in explicit time-stepping LSMs compatible with implicit time-stepping RTMs. The limitation of reactant availability on reaction rate is well represented by the rst-order rate (Eqs. A1, A2, A5, and A7): the rate decreases to zero as the concentration decreases to zero. A residual concentration [CNu]r is often added to represent a threshold be
low which a reaction stops, for example, to the decomposition rate (Eq. A1) as
d[CNu]dt = kdfT fw([CNu] [CNu]r). (1)
Here CNu is the upstream pool with 1 : u as the CN ratio in mole; [ ] denotes concentration; and kd, fT , and fw are the rate coefcient, and temperature and moisture response functions, respectively. When CNu goes below [CNu]r in an iter
ation, Eq. (1) implies a hypothetical reverse reaction to bring it back to [CNu]r. The residual concentration can be set to
zero to nullify the effect.
For the litter decomposition reactions (Appendix Reactions AR7AR9) that immobilize N, the nitrication Reaction (AR11) associated with decomposition to produce nitrous oxide, and the plant nitrogen uptake Reactions (AR13) and (AR14), the rate formulae do not account for the limitation of the reaction rate by nitrogen availability. Mechanistically, a nitrogen-limiting function needs to be added. For example, using the widely used Monod substrate limitation function (Fennell and Gossett, 1998), Eq. (1) becomes
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
930 G. Tang et al.: CLMPFLOTRAN biogeochemistry
(a) C cyclE
Plant
CWD = 2.74 y
C transformaCon REspiraCon
N minEralizaCon
1.00
N immobilizaCon/takE
NitricaCon
DEnitricaCon
N dEposiCon/xaCon
0.76
0.24
Lit1 = 20 h
Lit2 = 14 d
Lit3 = 71 d
0.39
0.55
0.29
0.28
0.46
0.55
SOM1 = 14 d
CN = 12
SOM2 = 71 d
CN = 12
SOM3 = 2 y
CN = 10
SOM4 = 27.4 y
CN = 10
(a) N cyclE
N2O
NH4+
NO3
N2
Figure 1. The reaction network for the carbon (a) and nitrogen (b) cycles implemented in this work. The carbon cycle is modied from Thornton and Rosenbloom (2005) and Bonan et al. (2012). is the turnover time, and CN is the CN ratio in g C over g N.
d[CNu]dt = kdfT fw([CNu] [CNu]r) [
N] [N]r [N] [N]r + km
Rn = (Rp Ra) [
NO3]
[NO3] + km
= Rp
, (2)
with half saturation km and a mineral nitrogen residual concentration [N]r. In the case of [N][N]r = km, the rate-
limiting function is equal to 0.5. For [N] km, Eq. (2) ap
proximates zero order with respect to [N]. For [N] km,
Eq. (2) approximates rst order with respect to [N]. The
derivative of the Monod term, km([N] + km)2, increases to
about k1m as the concentration decreases to below km. This represents a steep transition when km is small. The half saturation is expected to be greater than the residual concentration. When both are zero, the rate is not limited by the substrate availability.
To separate mineral nitrogen into ammonium (NH+4) and nitrate (NO3), it is necessary to partition the demands between ammonium and nitrate for plant uptake and immobilization. If we simulate the ammonium limitation on plant uptake with
Ra = Rp [
NH+4]
[NH+4] + km
km
[NH+4] + km
[NO3]
[NO3] + km
. (4)
In this equation Rp, Ra, and Rn are the plant uptake rates for nitrogen, ammonium (Reaction AR13), and nitrate (Reaction AR14). Rp is calculated in the CLM and input to
PFLOTRAN as a constant during each CLM time step. Equation (4) essentially assumes an inhibition of ammonium on nitrate uptake, which is consistent with the observation that plant nitrate uptake rate remains low until ammonium concentrations drop below a threshold (Eltrop and Marschner, 1996). However, the form of the rate expression will differ for different plants (Pfautsch et al., 2009; Warren and Adams, 2007; Nordin et al., 2001; Falkengren-Grerup, 1995; Gherardi et al., 2013), which will require different representations in future developments.
CLM uses a demand-based competition approach (Appendix A, Sect. A5) to represent the limitation of available nitrogen on plant uptake and immobilization. It is similar to the Monod function except that it introduces a discontinuity during the transition between the zero and rst-order rate.Implementation of the demand-based competition in a RTM involves separating the supply and consumption rates for each species in each reaction and limiting the consumption
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
, (3)
the plant nitrate uptake can be represented by
G. Tang et al.: CLMPFLOTRAN biogeochemistry 931
rates if supply is less than demand after contributions from all of the reactions are accumulated. It involves not only the rate terms for the residual but also the derivative terms for the Jacobian. The complexity increases quickly when more species need to be downregulated (e.g., ammonium, nitrate, and organic nitrogen) and there are transformation processes among these species. It becomes challenging to separate, track, and downregulate consumption and production rates for an indenite number of species, and calculation of the Jacobian becomes convoluted. In contrast, use of the Monod function with a residual concentration for individual reactions is easier to implement and allows for more exibility in adding reactions.
2.3 Scaling, clipping, and log transformation for avoiding negative concentration
Negative components of the concentration update ( ck+1,p+1 for iteration p from time-step k to k+1 in a NewtonRaphson
iteration) can result in negative concentration in some entries of ck+1,p (Eq. B6), which is nonphysical. One approach to avoid negative concentration is to scale back the update with a scaling factor ( ) (Bethke, 2007; Hammond, 2003) such that
ck+1,p+1 = ck+1,p + ck+1,p+1 > 0, (5)
where
= min
i=0,m[
1, ck+1,p(i)/ ck+1,p+1(i)] (6)
for negative ck+1,p+1(i) with m as the number of species times the number of numerical grid cells and as a factor with default value of 0.99. A second approach, used by RTM codes STOMP, HYDROGEOCHEM 5.0, and TOUGH-REACT, is clipping (i.e., for any ck+1,p+1(i) ck+1,p(i),
ck+1,p+1(i) = ck+1,p(i) [epsilon1] with [epsilon1] as a small number, e.g.,
1020). This limits the minimum concentration that can be modeled.
A third approach, log transformation, also ensures a positive solution (Bethke, 2007; Hammond, 2003; Parkhurst and Appelo, 1999). It is widely used in geochemical codes for describing highly variable concentrations for primary species such as H+ or O2 that can vary over many orders of magnitude as pH or redox state changes without the need to use variable switching. Instead of solving Eq. (B3) for ck+1 using
Eqs. (B4)(B6), this method solves for (lnck+1) (Hammond, 2003) with
Jln(i,j) =
@f (i)
@ ln(c(j)) = c(j)
3 Tests, results, and discussions
The NewtonRaphson method and scaling, clipping, and log transformation are widely used and extensively tested for RTMs, but not for coupled LSMRTM applications. The CLM describes biogeochemical dynamics within daily cycles for simulation durations of hundreds of years; the nitrogen concentration can be very low (mM to nM) while the carbon concentrations can be very high (e.g., thousands molm3 carbon in an organic layer); the concentrations and dynamics can vary dramatically in different locations around the globe. It is not surprising that the complex biogeochemical dynamics in a wide range of temporal and spatial scales in CLM poses numerical challenges for the RTM methods. Our simulations reveal some numerical issues (errors, divergence, and small time-step sizes) that were not widely reported. We identify the issues from coupled CLMPFLOTRAN simulations and reproduce them in simple test problems. We examine remedies in the simple test problems and test them in the coupled simulations.
For tests 13, we start with plant ammonium uptake to examine the numerical solution for Monod function, and then add nitrication and denitrication incrementally to assess the implications of adding reactions. For test 4, we check the implementation of mineralization and immobilization in the decomposition reactions. Third, we compare the nitrogen demand partition into ammonium and nitrate between CLM and PFLOTRAN. With coupled CLMPFLOTRAN spin-up simulations for Arctic, temperate, and tropical sites, we assess the application of scaling, clipping, and log transformation to achieve accurate, efcient, and robust simulations. Spreadsheets, PFLOTRAN input les, and additional materials are provided as Supplement, and archived at Tang et al.(2016).
Our implementation of CLM soil biogeochemistry introduces mainly two parameters: half saturation (km) and residual concentration. A wide range of km values was reported for ammonium, nitrate, and organic nitrogen for microbes and plants. The median, mean, and standard deviations range from 10 to 100, 50 to 500, and 10 to 200 M, respectively (Kuzyakov and Xu, 2013). Reported residual concentrations are limited likely because of the detection limits of the analytical methods and are considered to be zero (e.g., Hgh-Jensen et al., 1997). The detection limits are usually at the micromolar level, while up to the nanomolar level was reported (Nollet and De Gelder, 2013). In Ecosys, the km is 0.40 and 0.35 gNm3, and the residual concentration is 0.0125 and 0.03 gNm3 (Grant, 2013) for ammonium and nitrate for microbes, respectively. We start with km =
106 M or molm3, and residual concentration = 1015 M
or molm3 for plants and microbes. To further investigate the nonphysical solution negativity for the current study and for future application for other reactants (e.g., H2 and O2)
where the concentrations can be much lower, we examine km from 103 to 109 in our test problems. The km is expected to be different for different plants, microbes, and for ammo-
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
@f (i)
@c(j) = c(j)J(i,j), (7)
lnck+1,p+1 = J1lnf (ck+1,p), (8)
and
ck+1,p+1 = ck+1,p exp[ ln(ck+1,p+1)]. (9)
932 G. Tang et al.: CLMPFLOTRAN biogeochemistry
nium and nitrate. We do not differentiate them in this work as we focus on numerical issues.
3.1 Simple tests
3.1.1 Plant nitrogen uptake, nitrication, and denitrication
It was observed that plants can decrease nitrogen concentration to below the detection limit in hours (Kamer et al., 2001). In CLM, the total plant nitrogen demand is calculated based on photosynthesized carbon allocated for new growth and the C : N stoichiometry for new growth allocation, and the plant nitrogen demand from the soil (Rp) is equal to the total nitrogen demand minus retranslocated nitrogen stored in the plants (Oleson et al., 2013). The demand is provided as an input to PFLOTRAN. We use the Monod function to represent the limitation of nitrogen availability on uptake. Once the plant nitrogen uptake is calculated in PFLOTRAN, it is returned to CLM, which allocates the uptake among, leaf, live stem, dead stem, etc., and associated storage pools (Oleson et al., 2013). We examine the numerical solutions for the Monod equation at rst. Incrementally, we add rst-order reactions (e.g., nitrication, denitrication, and plant nitrate uptake) to look into the numerical issues in increasingly complex reaction networks.
Plant ammonium uptake (test 1)
We consider the plant ammonium uptake Reaction (AR13) with a rate Ra
d[NH+4]dt = Ra [
NH+4]
[NH+4] + km
. (10)
Discretizing it in time using the backward Euler method for a time-step size [Delta1]t, a solution is
[NH+4]k+1 = 0.5[bracketleftBig][
NH+4]k km Ra[Delta1]t[notdef]
[radicalBig]
[NH+4]k km Ra[Delta1]t
Even though clipping avoids convergence to the negative solution, the ammonium consumption is clipped, but the PlantA (Reaction AR13) production is not clipped (spreadsheet case5), violating the reaction stoichiometry. This results in mass balance errors for explicit time stepping (Tang and Riley, 2016). For implicit time stepping, additional iterations can resolve this violation to avoid mass balance error. However, if a nonreactive species is added with a concentration of 1000 molm3 (e.g., soil organic matter pool4 SOM4 in organic layer), the relative update snormrel =
[notdef][notdef] ck+1,p[notdef][notdef]2/[notdef][notdef]ck+1,p[notdef][notdef]2 decreases to 9.3 [notdef] 109 in the iter
ation (spreadsheet case5a). With a relative update tolerance (STOL; Eq. B9) of 108, the iteration would be deemed converged. This false convergence may produce mass balance error. Satisfying the relative update tolerance criteria does not guarantee that the residual equations are satised (Lichtner et al., 2015). For this reason, we need a tight STOL to avoid this false convergence so that additional iterations can be taken to solve the residual equation accurately.
In contrast to clipping, scaling applies the same scaling factor to limit both ammonium consumption and PlantA production following the stoichiometry of the reaction (spreadsheet case6). However, if we add a production reaction that is independent of plant ammonium uptake, say nitrate deposition, which can come from CLM as a constant rate in a time step rather than an internally balanced reaction, scaling reduces the plant ammonium uptake to account for the availability of ammonium as we intend, but the nitrate deposition rate is also reduced by the same scaling factor even though it is not limited by the availability of either ammonium or nitrate (spreadsheet case7). Like clipping, this unintended consequence can be resolved in the subsequent iterations, and a loose STOL may lead to false convergence and mass balance errors.
Small to zero concentration for ammonium and PlantA has no impact on the iterations for the clipping or scaling methods in this test. In contrast, a small initial PlantA concentration can cause challenges for the log transformation method even though PlantA is only a product. When it is zero, the Jacobian matrix is singular because zero is multiplied to the column corresponding to PlantA (Eq. 7). An initial PlantA concentration of 109 can result in underow of the exponential function (spreadsheet caseA, as a 64-bit real number (double precision), is precise to 15 signicant digits and has a range of e709 to e709 in IEEE 754 oating-point representation; Lemmon and Schafer, 2005). Clipping the update (say to be between 5 and 5) is needed to prevent numerical
overow or underow. Like the cases with clipping without log transformation, clipping violates reaction stoichiometry in the clipping iteration, and this violation needs to be resolved in subsequent iterations (spreadsheet caseB).
This simple test for the Monod function indicates that(1) NewtonRaphson iterations may converge to a negative concentration; (2) scaling, clipping, and log transformation can be used to avoid convergence to negative concentration;
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
2 +4km[NH+4]k[bracketrightbigg]
. (11)
Ignoring the negative root, [NH+4]k+1 0. Adding a resid
ual concentration by replacing [NH+4] with [NH+4][NH+4]r, [NH+4] [NH+4]r. Namely, the representation of plant am
monium uptake with the Monod function mathematically ensures [NH+4]k+1 [NH+4]r.
We use a spreadsheet to examine the NewtonRaphson iteration process for solving Eq. (10) and the application of clipping, scaling, and log transformation (test1.xlsx in the Supplement). When an overshoot gets the concentration closer to the negative than the positive root (Eq. 11), the iterations converge to the nonphysical negative semianalytical solution (spreadsheet case3). This can be avoided by using clipping, scaling, or log transformation (spreadsheet case4, case6, case8).
G. Tang et al.: CLMPFLOTRAN biogeochemistry 933
(3) small or zero concentration makes the Jacobian matrix stiff or singular when log transformation is used, and clipping is needed to guard against overow or underow of the exponential function; (4) clipping limits the consumption, but not the corresponding production, violating reaction stoichiometry in the iteration; (5) production reactions with external sources are inhibited in the iterations when scaling is applied, which is unintended; (6) additional iterations can resolve issues in (4) or (5); and (7) loose update tolerance convergence criteria may cause false convergence and result in mass balance errors for clipping and scaling.
Plant ammonium uptake and nitrication (test 2)
Adding a nitrication Reaction (AR10) with a rst-order rate to the plant ammonium uptake Reaction (AR13), Eq. (10) becomes
d[NH+4]dt = Ra [
NH+4]
[NH+4] + km
knitr[NH+4] = Rat Rnitr. (12)
A semianalytical solution similar to Eq. (11) can be derived for Eq. (12). With Jat = dRat/d[NH+4] =
Rakm([NH+4]+km)2, and Jnitr = dRnitr/d[NH+4] = knitr, the
matrix equation Eq. (B5) becomes
2
4
1/[Delta1]t + Jat + Jnitr 0 0
Jat 1/[Delta1]t 0
Jnitr 0 1/[Delta1]t
The scaling factor ( ) is a function of not only the update but also the concentration (Eq. 6). If a negative update is produced for a zero concentration, the scaling factor is zero, decreasing the scaled update to zero. The iteration converges without any change to the concentrations, numerically stopping all of the reactions in the time step unless STOL is negative. We add the denitrication reaction with Rnitr =
106 s1 to test1.xlsx case6 (Supplement) to create test2.xlsx (Supplement) to demonstrate that a small enough initial concentration relative to the negative update may numerically inhibit all of the reactions as well. An update of 6.6[notdef]106 M
is produced for nitrate (spreadsheet scale1). When the initial nitrate concentration [NO3]0 is not too small, say 106 M, the solution converges to the semianalytical solution in six iterations. When [NO3]0 is decreased to 109 M, the relative update snormrel is 9.2[notdef]1010. If STOL = 109, the solution
is deemed converged as Eq. (B9) is met, but not to the positive semianalytical solution. The ammonium uptake and nitrication reactions are numerically inhibited because the small scaling factor and a high concentration of a nonreactive species decreases the update to below STOL to reach false convergence. If we tighten STOL to 1030, the iterations continue, with decreasing nitrate concentration, , and snormrel by 2 orders of magnitude (1 as default = 0.99)
in each iteration, until snormrel reaches 1030 (spreadsheet scale2). Unless STOL is sufciently small, or MAXIT (the maximum number of iterations before stopping the current iterations and cutting the time step) is small (Appendix B), false convergence is likely to occur for the scaling method.The impact of numerical consumption on clipping and log transformation is much less dramatic than the scaling method as the latter applies the same scaling factor to the whole update vector following stoichiometric relationships of the reactions to maintain mass balance, and the limiting concentration decreases by (1 ) times in each iteration, with the
possibility of resulting in less than STOL relative update in MAXIT iterations.
In summary, this test problem demonstrates that (1) a negative update can be produced even for products during a NewtonRaphson iteration; and (2) when a negative update is produced for a very low concentration, a very small scaling factor may numerically inhibit all of the reactions due to false convergence even with very tight STOL.
Plant uptake, nitrication, and denitrication (test 3)
The matrix and update equations with plant nitrate uptake and denitrication added to test 2 are available in Appendix C. In addition to nitrate and PlantA, PlantN (Reaction AR14) and the denitrication product nitrogen gas may have negative updates. In addition to the off-diagonal terms due to the derivative of plant uptake with respect to ammonium concentration, the derivative of plant uptake with respect to nitrate concentration is added in the Jacobian matrix for PlantN (Eq. C1). As a result, a negative update for both ammonium
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
3
5
0
@
[NH+4]k+1,1
[PlantA]k+1,1
[NO3]k+1,1
1
1
A, (13)
for the rst iteration. As Rat + Rnitr 0, the ammonium up
date is negative even when the ammonium concentration is not very low. The off-diagonal terms for PlantA and nitrate in the Jacobian matrix bring the negative ammonium update into the updates for PlantA and nitrate even though there is no reaction that consumes them. Specically,
[PlantA]k+1,1[Delta1]t =
1 [Delta1]t
=
0
@
Rat + Rnitr
Rat
Rnitr
+Jnitr
1 [Delta1]t
+Jat + Jnitr
Rat
Jat
1 [Delta1]t
+Jat + Jnitr
Rnitr, (14)
[NO3]k+1,1
[Delta1]t =
Jnitr
1 [Delta1]t
+Jat + Jnitr
Rat
1 [Delta1]t
+Jat
1 [Delta1]t
Rnitr. (15)
Depending on the rates (Rat, Rnitr), derivatives (Jat, Jnitr), and time-step size [Delta1]t, the update can be negative for PlantA and nitrate, producing a zero-order numerical consumption, in which the limitation of availability is not explicitly represented. This has implications for the scaling method.
+ +Jat + Jnitr
934 G. Tang et al.: CLMPFLOTRAN biogeochemistry
0.20
(a)
0.05
(b)
0.2
0.04
0.04
Lit1 (mol m3 )
0.15
0.1
0.02
0.03
0.0 0 1 2
SOM1 (mol m3 )
0.00 0 1 2
0.10
0.02
0.05
km =106 km =109 km =1012
km =106 km =109 km =1012
0.01
0.00
0.00
102
10 5
-
(c)
(d)
102
104
104
+ (M)
105
1010
Rate (M d1 )
1 0 1 2
06
NH 4
1010
106
1 0 1 2
015
Immobilization Mineralization
1015
108
0 50 100 150 200Time (d)
0 50 100 150 200Time (d)
because of the faster immobilization than mineralization rate (Fig. 2c and d), Lit1 decomposition rate slows down to the level such that the immobilization rate is less than the mineralization rate. Namely, SOM1 decomposition controls Lit1 decomposition through limitation of mineralization on immobilization. As the immobilization rate decreases with decreasing Lit1, ammonium concentration rebounds after Lit1 is depleted. For km values of 106, 109, and 1012 M, Lit1 and SOM1 dynamics are similar except for slight differences in the early transit periods, but the ammonium values are decreased to 108, 1011, and 1014 M, respectively.
Smaller km values result in lower ammonium concentrations, which have implications for the clipping, scaling, and log transformation methods as discussed in tests 13.
3.1.3 Nitrogen demand partitioning between ammonium and nitrate
For comparison with CLM, we examine the uptake rate as a function of demands and available concentrations fpi =
(Ra + Rn)/Rp as implemented in Eqs. (3) and (4). As an
example, we consider uptake Rp = 109 Ms1 from a so
lution with various [NH+4] and [NO3] for a 0.5 h time step.
With CLM, fpi = 1 when [NH+4] + [NO3] Rp[Delta1]t; other
wise, it decreases with decreasing [NH+4] + [NO3] (Fig. 3).
The new representation (Eqs. 3 and 4) is generally similar, with fpi = 1 or 0 when [NH+4] or [NO3] or km.
For the intermediate concentrations, fpi in the new scheme is less than or equal to that in CLM because NH+4 inhibits NO3 uptake. The difference decreases with decreas-
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
Figure 2. Inuence of half saturation km on decomposition that involves both nitrogen immobilization and mineralization. Smaller half saturation can result in lower nitrogen concentration (c) but does not substantially impact the calculated concentrations other than ammonium (a, b).
and nitrate contributes to negative PlantN update through the two nonzero off-diagonal terms. Therefore, the likelihood for a negative update to PlantN is greater than PlantA as the former is inuenced by more rates and derivatives. We add plant nitrate uptake and denitrication into test2.xlsx (Supplement) and assess the implications of increased reactions and complexity in test3.xlsx (Supplement). In addition to nitrate, this introduces a negative update for nitrogen gas in the rst iteration (spreadsheet scale1). As the iterations resolve the balance between nitrite production from nitrication, and consumption due to plant uptake and denitrication, update to PlantN becomes negative and eventually leads to false convergence. The time-step size needs to be decreased from 1800 to 15 s to resolve the false convergence (spreadsheet scale2). In contrast, the added reactions have less impact on the clipping and log transformation methods.
3.1.2 Nitrogen immobilization and mineralization during decomposition (test 4)
We examine another part of the reaction network: decomposition, nitrogen immobilization, and mineralization (Fig. 1).We consider a case of decomposing 0.2 M Lit1C + 0.005 M
Lit1N to produce SOM1 with initial 4 M ammonium using the reactions (AR2) and (AR7) in the CLM-CN reaction network (Fig. 1). We use PFLOTRAN with a water-saturated grid cell with porosity of 0.25. At the beginning, Lit1 decreases and SOM1 increases sharply because the rate coefcient for Lit1 is 16 times that for SOM1 (Fig. 2a and b). As ammonium concentration decreases by orders of magnitude
G. Tang et al.: CLMPFLOTRAN biogeochemistry 935
Figure 3. fpi =
used because they are numerically more challenging as the simulations start far away from equilibrium. In these site simulations, PFLOTRAN uses the same 10 layer grid for the3.8 m one-dimensional column as CLM. The simulation durations are 1000, 600, and 600 years for the Arctic, temperate, and tropical sites, respectively. In the base case, km =
106 molm3 and residual concentration is 1015 molm3.
To assess the sensitivity of various preference levels for ammonium and nitrate uptake, and downregulation levels, we examine km = 103 to 109 molm3. We evaluate how scal
ing, clipping, and log transformation for avoiding negative concentrations inuence accuracy and efciency. The simulations are conducted using the ORNL Institutional Cluster (OIC Phase5, an SGI Altix with dual 8-core AMD Opterons per compute node) with CLMPFLOTRAN (as well as third-party libraries MPICH, PETSc, NetCDF, HDF5, etc.) compiled with gfortran 4.8.1 with the -O1 optimization level. Due to the small size of the simulations, our tests use only a single CPU core.
3.2.1 Site descriptions
The US-Brw site (71.35 N, 156.62 W) is located near
Barrow, Alaska. The mean annual temperature, precipitation, and snowfall are 12 C, 11 cm, and 69 cm, respec
tively (19712000) (Lara et al., 2012). The landscape is poorly drained polygonized tundra. The maximum thaw depth ranges from 30 to 40 cm, and the snow-free period is
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
Ntake
Ndemand = max
(bd) in a 0.5 h time step with an
uptake rate of 109 Ms1. fpi for the latter representation is less than or equal to that for the rst one. The difference decreases with decreasing half saturation km.
ing km, apparently disappearing at km = 1010. Various lev
els of preferences of ammonium over nitrate uptake were ob-served for plants (Pfautsch et al., 2009; Warren and Adams, 2007; Nordin et al., 2001; Falkengren-Grerup, 1995; Gherardi et al., 2013), which is similar to microbial uptake of inorganic and organic nitrogen species (Fouilland et al., 2007; Kirchman, 1994; Kirchman and Wheeler, 1998; Middelburg and Nieuwenhuize, 2000; Veuger et al., 2004). CLM implies a strong preference for ammonium over nitrate. For example, if ammonium is abundant, nitrate will not be taken by plants. The new scheme allows the level of preference to be adjusted by varying km; more realistic representations can be implemented relatively easily.
3.2 CLMPFLOTRAN simulations
We test the implementation by running CLMPFLOTRAN simulations for Arctic (US-Brw), temperate (US-WBW), and tropical (BR-Cax) AmeriFlux sites. The CLMPFLOTRAN simulations are run in the mode in which PFLOTRAN only handles subsurface chemistry (decomposition, nitrication, denitrication, plant nitrogen uptake). For comparison with CLM, (1) depth and O2 availability impact on decomposition, (2) cryoturbation, (3) SOM transport, and (4) nitrogen leaching are ignored by setting (1) decomp_depth_efolding to 106 m, o_scalar to 1, (2) cryoturb_diffusion, (3) som_diffusion, and (4) sf_no3 and sf_sminn to 0 (Oleson et al., 2013). Spin-up simulations are
1
1, [NH+4]+[NO3] Ndemand
[parenrightbigg]
[parenrightbigg]
[NO3] km+[NO3]
(a) vs. =
[NH+4] km+[NH+4] + [parenleftbigg]
[NH+4] km+[NH+4]
936 G. Tang et al.: CLMPFLOTRAN biogeochemistry
Figure 4. Calculated LAI and nitrogen distribution among vegetation, litter, SOM, NH+
4 , and NO
3 pools in spin-up simulations for the
US-Brw site.
variable in length but generally begins in early June and lasts until early September (Hinkel and Nelson, 2003). The area is composed of several different representative wetmoist coastal sedge tundra types, including wet sedges, grasses, moss, and assorted lichens. The leaf area index (LAI) is
1.1 (AmeriFlux data).
The US-WBW site (35.96 N, 84.29 W) is located in the Walker Branch Watershed in Oak Ridge, Tennessee (Hanson and Wullschleger, 2003). The climate is typical of the humid southern Appalachian region. The mean annual precipitation is 139 cm, and the mean median temperature is 14.5 C.
The soil is primarily Ultisols that developed in humid climates in the temperate zone on old or highly weathered material under forest. The temperate deciduous broadleaf forest was regenerated from agricultural land 50 years ago. LAI is
6.2 (Hanson et al., 2004).
The BR-Cax site (1.72 N, 51.46 W) is located in the
eastern Amazon tropical rainforest. The mean annual rainfall is between 200 and 250 cm, with a pronounced dry season between June and November. The soil is a yellow Oxisol (Brazilian classication Latossolo Amarelo) with a thick stony laterite layer at 34 m depth (da Costa et al., 2010). The vegetation is evergreen broadleaf forest. The LAI is
46 (Powell et al., 2013).
3.2.2 CLMPFLOTRAN site simulation results
The site climate data from 1998 to 2006, 2002 to 2010, and 2001 to 2006 are used to drive the spin-up simulation for the Arctic (US-Brw), temperate (US-WBW), and tropical (BRCax) sites, respectively. This introduces a multi-year cycle in addition to the annual cycle (Figs. 46). Overall, CLMPFLOTRAN is close to CLM4.5 in predicting LAI and nitrogen distribution among vegetation, litter, SOM (soil organic matter), ammonium and nitrate pools for the Arctic (Fig. 4), temperate (Fig. 5), and tropical (Fig. 6) sites. CLM4.5 does reach equilibrium earlier than CLMPFLOTRAN. The maximum differences occur during the transient periods (200 400 years for the Arctic, and 5070 years for the temperate and tropical sites) for SOMN (soil organic matter nitrogen), ammonium, and nitrate. This is not surprising as(1) the nitrogen demand competition scheme implemented in CLMPFLOTRAN is different from that in CLM4.5 (Fig. 3),(2) the former solves the reaction network simultaneously while the latter does so sequentially (resolve the plant up-take and decomposition rst, then nitrication, then denitrication), and (3) the carbon nitrogen cycle is very sensitive to the nitrogen competition representation. Close to steady state, both CLM4.5 and CLMPFLOTRAN overpredict the LAI at the Arctic and temperate sites, and underpredict soil organic matter accumulation, which will be resolved in future work.
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 937
Figure 5. Calculated LAI and nitrogen distribution among vegetation, litter, SOM, NH+
4 , and NO
3 pools in spin-up simulations for US-
WBW site.
The Arctic site shows a distinct summer growing season (Fig. 4): LAI and VEGN (vegetation nitrogen) jump up at the beginning, then level off, and drop down at the end of the growing season when LITN (litter nitrogen) jumps up due to litter fall. Ammonium and nitrate concentrations drop to very low levels at the beginning of the growing season and accumulate at other times. In addition to a longer growing season than the Arctic site, the temperate site shows more litter fall by the end of the growing season, as it is a temperate deciduous forest, which introduces immobilization demand that further lowers ammonium and nitrate concentrations (Fig. 5e inset). The seasonality is much less apparent in the tropical site than in the Arctic and temperate sites. LAI, VEGN, LITN, and SOMN accumulate with less seasonal variation to reach equilibrium.
The higher km of 103 molm3 results in lower immobilization, higher accumulation of LITN, and higher ammonium and nitrate concentrations than km of 106 molm3 for the tropical site during the spin-up (Fig. 6). This is not surprising becasue a higher km of 103 molm3 numerically poses a stricter limitation on the extent that plants and microbes can take from soils. The range of km values (106, and 109 molm3) generally has limited impact on the overall calculations except that the nitrogen concentrations drop lower with lower km values (e.g., inset in Figs. 4e and f and 5e). The lack of sensitivity is because these very low concentrations do not make up a mass of nitrogen that is sig-
nicant enough to inuence the carbon and nitrogen cycle.
However, as a small km means low concentrations (test 4), and weak downregulation and steep transition between zero order and rst order, it has implications on accuracy and efciency of the numerical solutions.
3.2.3 Accuracy and efciency
Numerical errors introduced due to false convergence in clipping, scaling, or log transformation are captured in CLM when it checks carbon and nitrogen mass balance for every time step for each column, and reports 108 gm2 errors
(to reduce log le size as the simulation durations are hundreds of years and the time-step size is half an hour). When log transformation is used, mass balance errors are not reported for the Arctic, temperate, and tropical sites with km values 103, 106, and 109 molm3. The computing time for CLMPFLOTRAN is about 60100 % more than that of
CLM (Table 1). This is not unreasonably high as the implicit method involves solving a Jacobian system for each NewtonRaphson step (Eq. B5), and log transformation converts the linear problem into a nonlinear one. The computational cost increases substantially with decreasing half saturation, which is expected as a smaller half saturation requires smaller time-step sizes to march through steeper transition between the zero- and rst-order rate in Monod function. Overall, log transformation is accurate, robust, and reasonably efcient.
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
938 G. Tang et al.: CLMPFLOTRAN biogeochemistry
Table 1. Wall time for CLMPFLOTRAN relative to CLM for spin-up simulation on OIC (ORNL Institutional Cluster Phase5).
Site Clipping Scaling Log transformation
km 103 106 109 103 106 109 103 106 109
US-Brw 1.28 1.30 1.30 1.29 1.29 1.32 1.45 1.49 1.72 US-WBW 1.45 1.47 1.47 1.45 1.45 1.47 1.64 1.68 1.89 BR-Cax 1.43 1.49 1.55 1.44 1.48 1.52 1.62 1.66 1.99
CLM wall time is 29.3, 17.7, and 17.1 h for the Arctic, temperate, and tropical sites for a simulation duration of 1000, 600, and 600 years. km
is the half saturation (mol m3).
Mass balance errors are reported for km values of 106, and 109 but not for 103 molm3 when clipping is applied. With km = 103 molm3, the plant uptake and immo
bilization are inhibited at relatively high concentration so that nitrogen concentrations are high. With km decreasing from 106 to 109 molm3, nitrogen concentrations are lowered to a much lower level (Figs. 46, similar to Fig. 2c), increasing the likelihood of overshoot. Mass balance errors are reported when the relative update is below STOL, preventing further iterations from resolving the violation of reaction stoichiometry introduced by clipping. The frequency of mass balance errors decreases with increasing km, and decreasing
STOL. Tightening STOL from 108 to 1012, the reported greater than 108 gm3 mass balance errors are eliminated.
The computing time is about 50 % more than CLM, which is more efcient than log transformation (Table 1), particularly for km = 109 molm3. Tightening STOL only slightly
increases the computing time. Because clipping often occurs at very low concentrations, the reported mass balance errors are usually small ( 108 gNm2 to 107 gNm2), and
do not have substantial impact on the overall simulation results.
The results for scaling is similar to clipping: mass balance errors are reported for km values of 106 and 109 but not for 103 molm3; tightening STOL to 1012 eliminates these errors; it takes about 50 % more computing time than CLM.
To examine the inuence of low concentrations on the accuracy and efciency of the scaling method, we conduct numerical experiments in which we reset the nitrous oxide concentration produced from decomposition (Reaction AR11, rate Eq. A3) to 1012, 1010, or 108 molm3 in each CLM half-hour time step for the tropical site for the rst year. This can be used to calculate the nitrous oxide production rate from decomposition and feed back to CLM without saving the concentration for the previous time step. Overall, nitrogen is abundant in the rst half year, and then becomes limiting in the last 5 months (Fig. 6e and f, inset). We look into the daily ammonium cycles as an example during the nitrogen-limiting period (day 250 to 260, Fig. 7a). Every day the ammonium concentration increases with time due to deposition, but drops when the plant nitrogen demand shots up. With a reset concentration of 108 molm3, the minimum nitrous oxide concentration for the 10 layers is 108 molm3, and
ammonium concentrations show two peaks followed by two drops due to the two plant uptake peaks every day. Decreasing the reset concentration to 1010 molm3, the minimum concentration drops to 1012, 1014, and 1016 molm3, corresponding to 1, 2, and 3 scaling iterations with overshoot for nitrous oxide. These result in numerical inhibition of nitrogen rebound every day. It worsens with further decrease of the reset concentration to 1012 molm3. This introduces mass balance errors as reported in CLM because the false convergence numerically inhibits all of the reactions including nitrogen deposition and litter input from CLM to PFLOTRAN. Unlike clipping, these false convergences introduce excessive mass balance errors because of the inhibition of productions specied from CLM. If all of the reactions are internally balanced, false convergence does not result in mass balance errors.
The frequent negative update to nitrous oxide is produced because the rate for the nitrication Reaction (AR11) is parameterized as a fraction of the net nitrogen mineralization rate to reect the relationship between labile carbon content and nitrous oxide production (Parton et al., 1996). A Monod function is added to describe the limitation of ammonium concentration on nitrication. Calculation of the net mineralization rate involves all of the decomposition reactions, and the litter decomposition reactions bring in ammonium and nitrate limitation, and ammonium inhibition on nitrate immobilization. As a result, the off-diagonal terms for nitrous oxide in the Jacobian matrix corresponding to Lit1C, Lit1N, Lit2C, Lit2N, Lit3C, Lit3N, SOM1, SOM2, SOM3, SOM4, ammonium, and nitrate are nonzero. Negative updates to all of these species contribute to negative updates to nitrous oxide. Similar to tests 23, a negative update to a low nitrous oxide concentration can decrease snormrel to below STOL, resulting in false convergence and mass balance errors. While the empirical parameterization of nitrication rate as a function of net mineralization rate is conceptually convenient, it increases the complexity of the reaction network and computational cost due to the reduced sparsity of the Jacobian matrix. While we use nitrous oxide as an example here, similar results can be obtained for PlantN, PlantA, and nitrogen gas produced from denitrication, etc. Theoretically, the numerical inhibition of all reactions can be caused by negative updates to very low concentrations for any species.
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 939
Figure 6. Calculated LAI and nitrogen distribution among vegetation, litter, SOM, NH+
4 , and NO
3 pools in spin-up simulations for BR-Cax
site.
10
(a)
9
Demand (10 gN m s)
NH (gN m)
10
(a)
9
Demand (10 gN m s)
6
10
10
NH (gN m)
6
3
10
10
Plant nitrogen demand
3
0
10 (b)
Plant nitrogen demand
NO-N (mol m)
0
10
10 (b)
10
NO-N (mol m)
STOL = 1010
STOL = 1030
STOL = 1050
10
10
10
10
ci = 1012
ci = 1010
ci = 108
10 250 252 254 256 258 260Day in the first year
Figure 8. Decreasing STOL can decrease and eliminate the numerical inhibition in the case of ci = 10
10 molm3 in Fig. 7. N2ON
concentration in the y axis in (b) is the minimum of the 10 soil layers.
The numerical errors can be decreased and eliminated by decreasing STOL. Similar to the tests 23 a small STOL can result in small snormrel, then a very small STOL is needed.
For the case of reset concentration 1010 for the 1 year tropical site simulation, the numerical inhibition decreases with decreasing STOL and varnishes for the observed time win-
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
250 252 254 256 258 260 Day in the first year
Figure 7. Resetting nitrous oxide concentration to 108, 1010, and 1012 molm3 in every CLM 0.5 h time-step results in no inhibition to increasing inhibition of reactions when the scaling method is used with STOL = 10
8. N2ON concentration in the
y axis in (b) is the minimum of the 10 soil layers. Numerical experiments are conducted for the tropical site for the rst year with km = 10
6 molm3. See inset in Fig. 6e for ammonium concentration in the rst year with daily data points.
940 G. Tang et al.: CLMPFLOTRAN biogeochemistry
dow when STOL = 1050 (Fig. 8), indicating the need for
very small, zero, or even negative STOL to avoid false convergence. The impact of resetting nitrous oxide concentration on clipping and log transformation is less dramatic. Nevertheless, the computing time increases about 10 % for clipping, and doubles for the log transformation.
4 Summary and conclusions
Global land surface models have traditionally represented subsurface soil biogeochemical processes using precongured reaction networks. This hardcoded approach makes it necessary to revise source code to test alternative models or to incorporate improved process understanding. We couple PFLOTRAN with CLM to facilitate testing of alternative models and incorporation of new understanding. We implement CLM-CN decomposition cascade, nitrication, denitrication, and plant nitrogen uptake reactions in CLMPFLOTRAN. We illustrate that with implicit time stepping using the NewtonRaphson method, the concentration can become negative during the iterations even for species that have no consumption, which need to be prevented by intervening in the NewtonRaphson iteration procedure.
Simply stopping the iteration with negative concentration and returning to the time-stepping subroutine to cut time-step size can avoid negative concentration, but may result in small time-step sizes and high computational cost. Clipping, scaling, and log transformation can all prevent negative concentration and reduce computational cost but at the risk of accuracy. Our results reveal implications when the relative update tolerance (STOL) is used as one of the convergence criteria. While use of STOL improves efciency in many situations, satisfying STOL does not guarantee satisfying the residual equation, and therefore it may introduce false convergence. Clipping reduces the consumption but not the production in some reactions, violating reaction stoichiometry.Subsequent iterations are required to resolve this violation.A tight STOL is needed to avoid false convergence and prevent mass balance errors. While the scaling method reduces the whole update vector following the stoichiometry of the reactions to maintain mass balance, a small scaling factor caused by a negative update to a small concentration may diminish the update and result in false convergence, numerically inhibiting all reactions, which is not intended for productions with external sources (e.g., nitrogen deposition from CLM to PFLOTRAN). For accuracy and efciency, a very tight STOL is needed when the concentration can be very low. Log transformation is accurate and robust, but requires more computing time. The computational cost increases with decreasing concentrations, most substantially for log transformation.
These computational issues arise because we switch from the explicit methods to the implicit methods for soil biogeo-chemistry. We use small half saturation (e.g., 109), resid-
ual concentration (e.g., 1015, slightly above machine zero), and large initial time-step size (e.g., 0.5 h) to exemplify the causes of the accuracy, efciency, and robustness issues. For the zero-order rate formulae in CLM, the limitation of reactant (nitrogen in this work) availability needs to be explicitly represented for robustness and exibility. With mechanistic representations, reaction stops or reverses when the rate-limiting reactants decrease to a threshold, or when the reaction becomes thermodynamically unfavorable, nullifying the need for half saturation and residual concentration. Before our representations are sufciently mechanistic, a small residual concentration (say 1015) serves as a safeguard to avoid failure for the log transformation method and unnecessary efciency and accuracy loss for the clipping and scaling methods unless a smaller residual concentration is dictated by physical and chemical conditions.
For reactions with very low half saturation and residual concentrations, e.g., redox reactions involving O2, H2, and
CH4, STOL can be set to zero to avoid false convergence, at the risk of potential increased computational cost. Increasing STOL (say to 1012) decreases computing time at the risk of potential accuracy loss. To cover a wide range of many orders of magnitude concentrations in soil biogeochemistry and for accuracy and robustness, the modelers can start with the log transformation method. If it is desirable to reduce the computational time, STOL can be relaxed, and clipping can be used without log transformation. The scaling method is another option but with strict STOL requirement. As the accuracy is checked and logged in CLM for carbon and nitrogen mass balance, the modelers can assess the tradeoff between efciency and accuracy in CLMPFLOTRAN and select their optimum.
Our CLMPFLOTRAN spin-up simulations at Arctic, temperate, and tropical sites produce results similar to CLM4.5, and indicate that accurate and robust solutions can be achieved with clipping, scaling, or log transformation. The computing time is 50 to 100 % more than CLM4.5 for a range of half saturation values from 103 to 109 and a residual concentration of 1015 for nitrogen. As physical half saturation ranges from 105 to 106 M for nitrogen, and the detection limits are often above 109 M, our results indicate that accurate, efcient, and robust solutions for current CLM soil biogeochemistry can be achieved using CLMPFLOTRAN.We thus demonstrate the feasibility of using an open-source, general-purpose reactive transport code with CLM, enabling signicantly more complicated and more mechanistic biogeochemical reaction systems.
An alternative to our approach of coupling LSMs with reactive transport codes is to code the solution to the advection, diffusion, and reaction equations directly in the LSM.This has been done using explicit time stepping and operator splitting to simulate the transport and transformation of carbon, nitrogen, and other species in CLM (Tang et al., 2013). An advantage of our approach of using a community RTM to solve the advectiondispersionreaction system is
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 941
that the signicant advances that the RTM community has made in the past several decades can be leveraged to better represent the geochemical processes (e.g., pH, pE) in a systematic, exible, and numerically reliable way. Given that a wide range of conditions may be encountered in any one global LSM simulation, it is particularly important to have robust solution methods such as fully implicit coupling of the advectiondispersionreaction equations (Yeh and Tripathi, 1989; Zheng and Bennett, 2002; Steefel et al., 2015).As a next step, we hope CLMPFLOTRAN will facilitate investigation of the role of the redox sensitive microbial reactions for methane production and consumption, and nitrication and denitrication reactions in ecological responses to climate change.
Code availability
PFLOTRAN is open-source software. It is distributed under the terms of the GNU Lesser General Public License as published by the Free Software Foundation either version 2.1 of the License, or any later version. It is available at https://bitbucket.org/pflotran
Web End =https://bitbucket.org/potran . CLM PFLOTRAN is under development and will be made available according to the NGEE-Arctic Data Management Policies and Plans (http://ngee-arctic.ornl.gov/content/ngee-arctic-data-management-policies-and-plans
Web End =http://ngee-arctic.ornl.gov/content/ http://ngee-arctic.ornl.gov/content/ngee-arctic-data-management-policies-and-plans
Web End =ngee-arctic-data-management-policies-and-plans http://ngee-arctic.ornl.gov/content/ngee-arctic-data-management-policies-and-plans
Web End = ). Before it becomes publicly available, please contact the corresponding authors to obtain access to the code.
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
942 G. Tang et al.: CLMPFLOTRAN biogeochemistry
Appendix A: CLM biogeochemical reactions and rates
A1 CLM-CN decomposition
The CLM-CN decomposition cascade consists of three litter pools with variable CN ratios, four soil organic matter (SOM) pools with constant CN ratios, and seven reactions (Bonan et al., 2012; Oleson et al., 2013; Thornton and Rosen-bloom, 2005). The reaction can be described by
CNu ! (1 f )CNd + f CO2 + nN, (AR1) where CNu and CNd are the upstream and downstream pools (molecular formula, for 1 mol upstream and downstream pool, there is u and d mol N), N is either NH+4 or NO3, f is the respiration fraction, and n = u (1 f )d. The rate is d[CNu]
dt = kdfT fw[CNu], (A1)
where kd is the rate coefcient, and fT and fw are the temperature and moisture response functions. With a constant CN ratio, the decomposition reactions for the four SOM pools are
SOM1 ! 0.72SOM2 + 0.28CO2 + 0.02N, (AR2)
SOM2 ! 0.54SOM3 + 0.46CO2 + 0.025143N, (AR3)
SOM3 ! 0.45SOM4 + 0.55CO2 + 0.047143N, (AR4) and
SOM4 ! CO2 + 0.085714N. (AR5)
The exact stoichiometric coefcients are calculated in the code using values for respiration factor, CN ratio, and molecular weight specied in the input le.
CLM4.5 has an option to separate N into NH+4 and NO3.
The N mineralization product is NH+4.
As the CN ratio is variable for the three litter pools, litter N pools need to be tracked such that Reaction (AR1) becomes
LitC + uLitN ! (1 f )CNd + f CO2 + nN, (AR6) where u = [LitN]/[LitC]. The three litter decomposition re
actions are
Lit1C + u1Lit1N ! 0.41SOM1 + 0.39CO2
+(u1 0.029286)N, (AR7)
Lit2C + u2Lit2N ! 0.45SOM2 + 0.55CO2
+(u2 0.032143)N, (AR8) and
Lit3C + u3Lit3N ! 0.71SOM3 + 0.29CO2
+(u3 0.060857)N. (AR9)
As the CN ratio of the litter pools is generally high, u1, u2, and u3 are usually small, and n in these reactions (e.g., n1 = u1 0.029286 for Lit1) is normally negative. Namely,
these reactions consume (immobilize) N, which can be NH+4, NO3, or both.
A2 Nitrication
The nitrication reaction to produce NO3 is
NH+4 + [notdef][notdef][notdef] ! NO3 + [notdef][notdef][notdef] (AR10) with [notdef][notdef][notdef] for additional reactants and products to balance the
reaction. The rate is (Dickinson et al., 2002)
d[NH+4]dt =
d[NO3]dt = knfT fw[NH+4]. (A2)
The nitrication reaction to produce N2O is
NH+4 + [notdef][notdef][notdef] ! 0.5N2O + [notdef][notdef][notdef], (AR11)
with one component related to decomposition as
d[NH+4]dt = 2
d[N2O]dt = fnmfT fwfpHmax(Rnm,0) (A3)
with fnm as a fraction (Parton et al., 1996), and Rnm as the net N mineralization rate,
Rnm =Xi
niRi, (A4)
where Ri denotes the rate of Reactions (AR2)(AR5) and (AR7)(AR9). The second component is (Parton et al., 1996)
d[NH+4]dt = 2
d[N2O]dt =
kN2OfT fwfpH(1 e0.0105[NH
+
4 ]). (A5)
Ignoring the high-order terms and moving the unit conversion factor into kN2O, it can be simplied as a rst-order rate as
d[NH+4]dt = 2
d[N2O]dt = kN2OfT fwfpH[NH+4]. (A6)
A3 Denitrication
The denitrication reaction is
NO3 + [notdef][notdef][notdef] ! 0.5N2 + [notdef][notdef][notdef] (AR12)
with rate (Dickinson et al., 2002)
d[NO3]dt = 2
d[N2]dt = kdenifT fwfpH[NO3]. (A7)
A4 Plant nitrogen uptake
The plant nitrogen uptake reaction can be written as
NH+4 + [notdef][notdef][notdef] ! PlantA + [notdef][notdef][notdef] (AR13) and
NO3 + [notdef][notdef][notdef] ! PlantN + [notdef][notdef][notdef]. (AR14)
The rate is specied by CLM (plant nitrogen demand) and assumed to be constant in each half-hour time step.
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 943
A5 Demand-based competition and demand distribution between ammonium and nitrate
Denote Rd, p and Rd, i as the potential plant, immobilization, nitrication, and denitrication demand (rate); Ra, tot =
Rd, p + Rd, i as the total NH+4 demand; and Rn, tot as the
total NO3 demand. CLM uses a demand-based competition approach to split the available sources in proportion to the demand rates to meet the demands (Oleson et al., 2013; Thornton and Rosenbloom, 2005). Specically, for each time step, if Ra, tot[Delta1]t [NH+4], the uptakes are equal
to potential demands, and Rn, tot = 0; otherwise, the uptakes
for NH+4 are [NH+4]Rd, p/Ra, tot[Delta1]t and [NH+4]Rd, i/Ra, tot[Delta1]t
for plants and immobilization; Rn, tot = Ra, tot [NH+4]/[Delta1]t.
If Rn, tot[Delta1]t < [NO3], all of the remaining demand Rn, tot is
met with available NO3. Otherwise, available NO3 is split to meet the remaining plant, immobilization, and denitrication demands in proportion to their rates.
Appendix B: Implicit time stepping and NewtonRaphson iteration
Ignoring equilibrium reactions and transport for simplicity of discussion in this work, PFLOTRAN solves the ordinary differential equation,
dc/dt = R(c) + F, (B1)
where c is the concentration vector, R is the kinetic reaction rate, and F is the uxes (e.g., nitrogen deposition). Discretizing Eq. (B1) in time using the backward Euler method,
(ck+1 ck)/[Delta1]t = R(ck+1) + F k. (B2)
Solving the equation using the NewtonRaphson method, we denote the residual as
f (ck+1,p) = (ck+1,p ck)/[Delta1]t R(ck+1,p) F k, (B3)
and Jacobian as
J =
@f (ck+1,p)
@ck+1,p
Specically,
[notdef][notdef]f (ck+1,p+1)[notdef][notdef]2 < ATOL, (B7)
[notdef][notdef]f (ck+1,p+1)[notdef][notdef]2
[notdef][notdef]f (ck+1,0)[notdef][notdef]2
< RTOL, (B8)
or
[notdef][notdef] ck+1,p+1[notdef][notdef]2
[notdef][notdef]ck+1,p+1[notdef][notdef]2
< STOL. (B9)
If none of these tolerances are met in MAXIT iterations or MAXF function evaluations, the iteration is considered to diverge, and PFLOTRAN decreases the time-step size for MAX_CUT times. The default values in PFLOTRAN are ATOL = 1050, RTOL = 108, STOL = 108,
MAXIT = 50, MAXF = 104, and MAX_CUT = 16.
Appendix C: Matrix equation for test 3
Adding to test 2 a plant NO3 uptake Reaction (AR14) with rate Rnt = Rp
[NH+4] [NH+4]+km
[NO3] [NO3]+km
,
Jnt, n =
dRnt d[NO3] =
Rp [
NH+4] [NH+4]+km
km ([NO3]+km)
2 , and Jnt, a =
dRnt d[NH+4] =
dRn d[NH+4]
km ([NH+4]+km)
[NO3]
2 [NO3]+km
, and a denitri
cation Reaction (AR12) with rate Rdeni = kdeni[NO3], and Jdeni =
dRdeni d[NO3] =
kdeni, the matrix equation (Eq. B5)
becomes
2
6
6
6
6
4
1 [Delta1]t
+Jat + Jnitr 0 0 0 0
Jat 1[Delta1]t 0 0 0
Jnitr + Jnt, a 0
1 [Delta1]t
+Jnt + Jdeni 0 0
Jnt, a 0 Jnt, n 1/[Delta1]t 0
0 0 0.5Jdeni 0 1/[Delta1]t
3
7
7
7
7
5
0
B
B
B
B
@
[NH+4]k+1,1
[PlantA]k+1,1
[NO3]k+1,1
[PlantN]k+1,1
[N2]k+1,1
1
C
C
C
C
A
=
0
B
B
B
B
@
Rat + Rnitr
Rat
Rnitr + Rnt + Rdeni
Rnt
0.5Rdeni
1
C
C
C
C
A
. (C1)
, (B4)
where p is the iteration counter, the update is
ck+1,p+1 = J1f (ck+1,p), (B5)
and the iteration equation is
ck+1,p+1 = ck+1,p + ck+1,p+1. (B6)
The iteration continues until either the residual f (ck+1,p+1) or the update ck+1,p+1 is less than a specied tolerance.
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
944 G. Tang et al.: CLMPFLOTRAN biogeochemistry
The Supplement related to this article is available online at http://dx.doi.org/10.5194/gmd-9-927-2016-supplement
Web End =doi:10.5194/gmd-9-927-2016-supplement .
Author contributions. G. Bisht, B. Andre, R. T. Mills, J. Kumar, and F. M. Hoffman. developed the CLMPFLOTRAN framework that this work is built upon. F. Yuan, G. Tang, G. Bisht, and X. Xu added biogeochemistry to the CLMPFLOTRAN interface. F. Yuan proposed the nitrication and denitrication reactions and rate formulae. G. Tang, F. Yuan, and X. Xu implemented the CLM soil biogeochemistry in PFLOTRAN under guidance of G. E. Hammond,P. C. Lichtner, S. L. Painter, and P. E. Thornton. G. Tang prepared the manuscript with contributions from all co-authors. G. Tang,F. Yuan, G. Bisht, and G. E. Hammond contributed equally to the work.
Acknowledgements. Thanks to Nathaniel O. Collier at ORNL for many discussions that contributed signicantly to this work. Thanks to Kathie Tallant and Kathy Jones at ORNL for editing service. This research was funded by the U.S. Department of Energy, Ofce of Sciences, Biological and Environmental Research, Terrestrial Ecosystem Sciences and Subsurface Biogeochemical Research Program, and is a product of the Next-Generation Ecosystem Experiments in the Arctic (NGEE-Arctic) project. Development of CLMPFLOTRAN was partially supported by the ORNL Laboratory Directed Research and Development (LDRD) program. ORNL is managed by UT-Battelle, LLC, for the U.S. Department of Energy under contract DE-AC05-00OR22725.
Disclaimer. This manuscript has been authored by UT-Battelle, LLC under contract no. DE-AC05-00OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan
Web End =http://energy.gov/downloads/doe-public-access-plan ).
Edited by: C. Sierra
References
Bethke, C. M.: Geochemical and Biogeochemical Reaction Modeling, Cambridge University Press, 2007.
Bonan, G. B., Hartman, M. D., Parton, W. J., and Wieder, W. R.: Evaluating litter decomposition in earth system models with long-term litterbag experiments: an example using the Community Land Model version 4 (CLM4), Glob. Change Biol., 19, 957974, doi:http://dx.doi.org/10.1111/gcb.12031
Web End =10.1111/gcb.12031 http://dx.doi.org/10.1111/gcb.12031
Web End = , 2012.
Boyer, E. W., Alexander, R. B., Parton, W. J., Li, C., Butterbach-Bahl, K., Donner, S. D., Skaggs, R. W., and Grosso, S. J. D.:
Modeling denitrication in terrestrial and aquatic ecosystems at regional scales, Ecol. Appl., 16, 21232142, doi:10.1890/1051-0761(2006)016[2123:MDITAA]2.0.CO;2 , 2006.
Conrad, R.: Soil microorganisms as controllers of atmospheric trace gases (H2, CO, CH4, OCS, N2O, and NO), Microbi.
Rev., 60, 609640, available at: http://mmbr.asm.org/content/60/4/609.full.pdf
Web End =http://mmbr.asm.org/content/60/ http://mmbr.asm.org/content/60/4/609.full.pdf
Web End =4/609.full.pdf (last access: 14 December 2015), 8987358[pmid], 1996.da Costa, A. C. L., Galbraith, D., Almeida, S., Portela, B. T. T., da Costa, M., de Athaydes Silva Junior, J., Braga, A. P., de Gonalves, P. H. L., de Oliveira, A. A. R., Fisher, R., Phillips, O. L., Metcalfe, D. B., Levy, P., and Meir, P.: Effect of 7 yr of experimental drought on vegetation dynamics and biomass storage of an eastern Amazonian rainforest, New Phytol., 187, 579591, doi:http://dx.doi.org/10.1111/j.1469-8137.2010.03309.x
Web End =10.1111/j.1469-8137.2010.03309.x http://dx.doi.org/10.1111/j.1469-8137.2010.03309.x
Web End = , 2010.
Dickinson, R. E., Berry, J. A., Bonan, G. B., Collatz, G. J., Field, C. B., Fung, I. Y., Goulden, M., Hoff-mann, W. A., Jackson, R. B., Myneni, R., Sellers, P. J., and Shaikh, M.: Nitrogen controls on climate model evapotranspiration, J. Climate, 15, 278295, doi:http://dx.doi.org/10.1175/1520-0442(2002)015<0278:NCOCME>2.0.CO;2
Web End =10.1175/1520- http://dx.doi.org/10.1175/1520-0442(2002)015<0278:NCOCME>2.0.CO;2
Web End =0442(2002)015<0278:NCOCME>2.0.CO;2 , 2002.
Eltrop, L. and Marschner, H.: Growth and mineral nutrition of non-mycorrhizal and mycorrhizal Norway spruce (Picea abies) seedlings grown in semi-hydroponic sand culture, New Phytol., 133, 469478, doi:http://dx.doi.org/10.1111/j.1469-8137.1996.tb01914.x
Web End =10.1111/j.1469-8137.1996.tb01914.x http://dx.doi.org/10.1111/j.1469-8137.1996.tb01914.x
Web End = , 1996. Falkengren-Grerup, U.: Interspecies differences in the preference of ammonium and nitrate in vascular plants, Oecologia, 102, 305 311, doi:http://dx.doi.org/10.1007/BF00329797
Web End =10.1007/BF00329797 http://dx.doi.org/10.1007/BF00329797
Web End = , 1995.
Fang, Y., Huang, M., Liu, C., Li, H., and Leung, L. R.: A generic biogeochemical module for Earth system models: Next Generation BioGeoChemical Module (NGBGC), version 1.0, Geosci. Model Dev., 6, 19771988, doi:http://dx.doi.org/10.5194/gmd-6-1977-2013
Web End =10.5194/gmd-6-1977- http://dx.doi.org/10.5194/gmd-6-1977-2013
Web End =2013 , 2013.
Fennell, D. E. and Gossett, J. M.: Modeling the production of and competition for hydrogen in a dechlorinating culture, Environ. Sci. Technol., 32, 24502460, doi:http://dx.doi.org/10.1021/es980136l
Web End =10.1021/es980136l http://dx.doi.org/10.1021/es980136l
Web End = , 1998. Fouilland, E., Gosselin, M., Rivkin, R. B., Vasseur, C., and Mostajir, B.: Nitrogen uptake by heterotrophic bacteria and phytoplankton in Arctic surface waters, J. Plankton Res., 29, 369376, doi:http://dx.doi.org/10.1093/plankt/fbm022
Web End =10.1093/plankt/fbm022 http://dx.doi.org/10.1093/plankt/fbm022
Web End = , 2007.
Gherardi, L. A., Sala, O. E., and Yahdjian, L.: Preference for different inorganic nitrogen forms among plant functional types and species of the Patagonian steppe, Oecologia, 173, 10751081, doi:http://dx.doi.org/10.1007/s00442-013-2687-7
Web End =10.1007/s00442-013-2687-7 http://dx.doi.org/10.1007/s00442-013-2687-7
Web End = , 2013.
Grant, R. F.: Modelling changes in nitrogen cycling to sustain increases in forest productivity under elevated atmospheric CO2 and contrasting site conditions, Biogeosciences, 10, 77037721, doi:http://dx.doi.org/10.5194/bg-10-7703-2013
Web End =10.5194/bg-10-7703-2013 http://dx.doi.org/10.5194/bg-10-7703-2013
Web End = , 2013.
Gu, C. and Riley, W. J.: Combined effects of short term rainfall patterns and soil texture on soil nitrogen cycling a modeling analysis, J. Contam. Hydrol., 112, 141154, doi:http://dx.doi.org/10.1016/j.jconhyd.2009.12.003
Web End =10.1016/j.jconhyd.2009.12.003 http://dx.doi.org/10.1016/j.jconhyd.2009.12.003
Web End = , 2010.
Hammond, G. E.: Innovative Methods for Solving Multicomponent Biogeochemical groundwater Transport on Supercomputers, Thesis, University of Illinois, Urbana-Champaign, 2003. Hammond, G. E., Lichtner, P. C., and Mills, R. T.: Evaluating the performance of parallel subsurface simulators: an illustrative
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
G. Tang et al.: CLMPFLOTRAN biogeochemistry 945
example with PFLOTRAN, Water Resour. Res., 50, 208228, doi:http://dx.doi.org/10.1002/2012WR013483
Web End =10.1002/2012WR013483 http://dx.doi.org/10.1002/2012WR013483
Web End = , 2014.
Hanson, P. and Wullschleger, S.: North American Temperate Deciduous Forest Responses to Changing Precipitation Regimes, Springer Verlag, 2003.
Hanson, P. J., Amthor, J. S., Wullschleger, S. D., Wilson, K. B., Grant, R. F., Hartley, A., Hui, D., Hunt, J. E. R., Johnson, D. W., Kimball, J. S., King, A. W., Luo, Y., McNulty, S. G., Sun, G., Thornton, P. E., Wang, S., Williams, M., Baldocchi, D. D., and Cushman, R. M.: Oak forest carbon and water simulations: model intercomparisons and evaluations against independent data, Ecol. Monogr., 74, 443489, doi:http://dx.doi.org/10.1890/03-4049
Web End =10.1890/03-4049 http://dx.doi.org/10.1890/03-4049
Web End = , 2004.
Hartman, M. D., Baron, J. S., and Ojima, D. S.: Application of a coupled ecosystem-chemical equilibrium model, DayCent-Chem, to stream and soil chemistry in a Rocky Mountain watershed, Ecol. Model., 200, 493510, doi:http://dx.doi.org/10.1016/j.ecolmodel.2006.09.001
Web End =10.1016/j.ecolmodel.2006.09.001 http://dx.doi.org/10.1016/j.ecolmodel.2006.09.001
Web End = , 2007.
Hinkel, K. M. and Nelson, F. E.: Spatial and temporal patterns of active layer thickness at Circumpolar Active Layer Monitoring (CALM) sites in northern Alaska, 19952000, J. Geophys. Res.-Atmos., 108, 8168, doi:http://dx.doi.org/10.1029/2001JD000927
Web End =10.1029/2001JD000927 http://dx.doi.org/10.1029/2001JD000927
Web End = , 2003.Hgh-Jensen, H., Wollenweber, B., and Schjoerring, J. K.: Kinetics of nitrate and ammonium absorption and accompanying H+ uxes in roots of Lolium perenne L., and N2-xing Trifolium repens L., Plant Cell Environ., 20, 11841192, doi:http://dx.doi.org/10.1046/j.1365-3040.1997.d01-145.x
Web End =10.1046/j.1365-3040.1997.d01-145.x http://dx.doi.org/10.1046/j.1365-3040.1997.d01-145.x
Web End = , 1997.
Hungate, R.: The rumen microbial ecosystem, Annu. Rev. Ecol. Syst., 6, 3966, doi:http://dx.doi.org/10.1146/annurev.es.06.110175.000351
Web End =10.1146/annurev.es.06.110175.000351 http://dx.doi.org/10.1146/annurev.es.06.110175.000351
Web End = , 1975.
IPCC: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge University Press, Cambridge, UK and New York, NY, USA, doi:http://dx.doi.org/10.1017/CBO9781107415324
Web End =10.1017/CBO9781107415324 http://dx.doi.org/10.1017/CBO9781107415324
Web End = , 2013.
Jarrell, K. F.: Extreme oxygen sensitivity in methanogenic archae-bacteria, BioScience, 35, 298302, doi:http://dx.doi.org/10.2307/1309929
Web End =10.2307/1309929 http://dx.doi.org/10.2307/1309929
Web End = , 1985.Kamer, K., Kennison, R. L., and Fong, P.: Rates of inorganic nitrogen uptake by the estuarine green macroalgae Enteromorpha intestinalis and Ulva expansa, 2003, 130141, 2001.Kirchman, D. L.: The uptake of inorganic nutrients by heterotrophic bacteria, Microb. Ecol., 28, 255271, doi:http://dx.doi.org/10.1007/BF00166816
Web End =10.1007/BF00166816 http://dx.doi.org/10.1007/BF00166816
Web End = , 1994.
Kirchman, D. L. and Wheeler, P. A.: Uptake of ammonium and nitrate by heterotrophic bacteria and phytoplankton in the sub-Arctic Pacic, Deep-Sea Res. Pt. I, 45, 347365, doi:http://dx.doi.org/10.1016/S0967-0637(97)00075-7
Web End =10.1016/S0967-0637(97)00075-7 http://dx.doi.org/10.1016/S0967-0637(97)00075-7
Web End = , 1998.
Kuzyakov, Y. and Xu, X.: Competition between roots and microorganisms for nitrogen: mechanisms and ecological relevance, New Phytol., 198, 656669, doi:http://dx.doi.org/10.1111/nph.12235
Web End =10.1111/nph.12235 http://dx.doi.org/10.1111/nph.12235
Web End = , 2013.Lara, M. J., Villarreal, S., Johnson, D. R., Hollister, R. D., Webber, P. J., and Tweedie, C. E.: Estimated change in tundra ecosystem function near Barrow, Alaska between 1972 and 2010, Environ. Res. Lett., 7, 015507, doi:http://dx.doi.org/10.1088/1748-9326/7/1/015507
Web End =10.1088/1748-9326/7/1/015507 http://dx.doi.org/10.1088/1748-9326/7/1/015507
Web End = , 2012.
Lemmon, D. R. and Schafer, J. L.: Developing Statistical Software in Fortran 95, Statistics and Computing, Springer, 2005.Lichtner, P. C., Hammond, G. E., Lu, C., Karra, S., Bisht, G., Andre, B., Mills, R. T., and Jitu, K.: PFLOTRAN User Manual:
a Massively Parallel Reactive Flow and Transport Model for Describing Surface and Subsurface Processes, Report, 2015. Maggi, F., Gu, C., Riley, W. J., Hornberger, G. M., Venterea, R. T., Xu, T., Spycher, N., Steefel, C., Miller, N. L., and Oldenburg, C. M.: A mechanistic treatment of the dominant soil nitrogen cycling processes: model development, testing, and application, J. Geophys. Res.-Biogeo., 113, G02016, doi:http://dx.doi.org/10.1029/2007JG000578
Web End =10.1029/2007JG000578 http://dx.doi.org/10.1029/2007JG000578
Web End = , 2008.
Manzoni, S. and Porporato, A.: Soil carbon and nitrogen mineralization: theory and models across scales, Soil Biol. Biochem., 41, 13551379, doi:http://dx.doi.org/10.1016/j.soilbio.2009.02.031
Web End =10.1016/j.soilbio.2009.02.031 http://dx.doi.org/10.1016/j.soilbio.2009.02.031
Web End = , 2009. Middelburg, J. J. and Nieuwenhuize, J.: Nitrogen uptake by heterotrophic bacteria and phytoplankton in the nitrate-rich Thames Estuary, Mar. Ecol.-Prog. Ser., 203, 1321, doi:http://dx.doi.org/10.3354/meps203013
Web End =10.3354/meps203013 http://dx.doi.org/10.3354/meps203013
Web End = , 2000.
Nollet, L. M. L. and De Gelder, L. S. P.: Handbook of Water Analysis, 3rd Edn., CRC Press, 2013.
Nordin, A., Hgberg, P., and Nsholm, T.: Soil nitrogen form and plant nitrogen uptake along a boreal forest productivity gradient, Oecologia, 129, 125132, doi:http://dx.doi.org/10.1007/s004420100698
Web End =10.1007/s004420100698 http://dx.doi.org/10.1007/s004420100698
Web End = , 2001. Oleson, K., Lawrence, D., Bonan, G., Levis, S., Swenson, S.,
Thornton, P., Bozbiyik, A., Fisher, R., Heald, C., Kluzek, E., Lamarque, J.-F., Lawrence, P., Lipscomb, W., Muszala, S., and Sacks, W.: Technical Description of Version 4.5 of the Community Land Model (CLM), NCAR/TN-503+STR, NCAR Technical Note, NCAR, doi:http://dx.doi.org/10.5065/D6RR1W7M
Web End =10.5065/D6RR1W7M http://dx.doi.org/10.5065/D6RR1W7M
Web End = , 2013.
Parkhurst, D. L. and Appelo, C.: Users Guide to PHREEQC (Version 2): a Computer Program for Speciation, Batch-Reaction, One-Dimensional Transport, and Inverse Geochemical Calculations, Water-Resources Investigations 99-4259, USGS, 1999. Parton, W. J., Mosier, A. R., Ojima, D. S., Valentine, D. W.,
Schimel, D. S., Weier, K., and Kulmala, A. E.: Generalized model for N2 and N2O production from nitrication and denitrication, Global Biogeochem. Cy., 10, 401412, doi:http://dx.doi.org/10.1029/96GB01455
Web End =10.1029/96GB01455 http://dx.doi.org/10.1029/96GB01455
Web End = , 1996.
Parton, W. J., Holland, E. A., Del Grosso, S. J., Hartman, M. D., Martin, R. E., Mosier, A. R., Ojima, D. S., and Schimel, D. S.: Generalized model for NOx and N2O emissions from soils, J. Geophys. Res.-Atmos., 106, 1740317419, doi:http://dx.doi.org/10.1029/2001JD900101
Web End =10.1029/2001JD900101 http://dx.doi.org/10.1029/2001JD900101
Web End = , 2001.
Pfautsch, S., Rennenberg, H., Bell, T. L., and Adams, M. A.: Nitrogen uptake by Eucalyptus regnans and Acacia spp. preferences, resource overlap and energetic costs, Tree Physiol., 29, 389399, doi:http://dx.doi.org/10.1093/treephys/tpn033
Web End =10.1093/treephys/tpn033 http://dx.doi.org/10.1093/treephys/tpn033
Web End = , 2009.
Powell, T. L., Galbraith, D. R., Christoffersen, B. O., Harper, A., Imbuzeiro, H. M. A., Rowland, L., Almeida, S., Brando, P. M., da Costa, A. C. L., Costa, M. H., Levine, N. M., Malhi, Y., Saleska, S. R., Sotta, E., Williams, M., Meir, P., and Moor-croft, P. R.: Confronting model predictions of carbon uxes with measurements of Amazon forests subjected to experimental drought, New Phytol., 200, 350365, doi:http://dx.doi.org/10.1111/nph.12390
Web End =10.1111/nph.12390 http://dx.doi.org/10.1111/nph.12390
Web End = , 2013.
Riley, W. J., Maggi, F., Kleber, M., Torn, M. S., Tang, J. Y., Dwivedi, D., and Guerry, N.: Long residence times of rapidly decomposable soil organic matter: application of a multi-phase, multi-component, and vertically resolved model (BAMS1) to soil carbon dynamics, Geosci. Model Dev., 7, 13351355, doi:http://dx.doi.org/10.5194/gmd-7-1335-2014
Web End =10.5194/gmd-7-1335-2014 http://dx.doi.org/10.5194/gmd-7-1335-2014
Web End = , 2014.
www.geosci-model-dev.net/9/927/2016/ Geosci. Model Dev., 9, 927946, 2016
946 G. Tang et al.: CLMPFLOTRAN biogeochemistry
Schmidt, M. W. I., Torn, M. S., Abiven, S., Dittmar, T., Guggenberger, G., Janssens, I. A., Kleber, M., Kogel-Knabner, I., Lehmann, J., Manning, D. A. C., Nannipieri, P., Rasse, D. P., Weiner, S., and Trumbore, S. E.: Persistence of soil organic matter as an ecosystem property, Nature, 478, 4956, doi:http://dx.doi.org/10.1038/nature10386
Web End =10.1038/nature10386 http://dx.doi.org/10.1038/nature10386
Web End = , 2011.
Sellers, P. J., Dickinson, R. E., Randall, D. A., Betts, A. K.,
Hall, F. G., Berry, J. A., Collatz, G. J., Denning, A. S., Mooney, H. A., Nobre, C. A., Sato, N., Field, C. B., and Henderson-Sellers, A.: Modeling the exchanges of energy, water, and carbon between continents and the atmosphere, Science, 275, 502509, doi:http://dx.doi.org/10.1126/science.275.5299.502
Web End =10.1126/science.275.5299.502 http://dx.doi.org/10.1126/science.275.5299.502
Web End = , 1997.Seneviratne, S. I., Corti, T., Davin, E. L., Hirschi, M.,
Jaeger, E. B., Lehner, I., Orlowsky, B., and Teuling, A. J.: Investigating soil moistureclimate interactions in a changing climate: a review, Earth-Sci. Rev., 99, 125161, doi:http://dx.doi.org/10.1016/j.earscirev.2010.02.004
Web End =10.1016/j.earscirev.2010.02.004 http://dx.doi.org/10.1016/j.earscirev.2010.02.004
Web End = , 2010.
Shampine, L. F., Thompson, S., Kierzenka, J. A., and Byrne, G. D.: Non-negative solutions of ODEs, Appl. Math. Comput., 170, 556569, doi:http://dx.doi.org/10.1016/j.amc.2004.12.011
Web End =10.1016/j.amc.2004.12.011 http://dx.doi.org/10.1016/j.amc.2004.12.011
Web End = , 2005.
Sonnenthal, E. L., Spycher, N., Xu, T., Zheng, L., Miller, N. L., and Pruess, K.: TOUGHREACT V3. 0-OMP Reference Manual: a Parallel Simulation Program for Non-Isothermal Multiphase Geochemical Reactive Transport, LBNL, Report, available at: http://esd1.lbl.gov/files/research/projects/tough/documentation/TOUGHREACT_V3-OMP_RefManual.pdf
Web End =http://esd1.lbl.gov/les/research/projects/tough/ http://esd1.lbl.gov/files/research/projects/tough/documentation/TOUGHREACT_V3-OMP_RefManual.pdf
Web End =documentation/TOUGHREACT_V3-OMP_RefManual.pdf (last access: 29 February 2016), 2014.
Steefel, C. I., Appelo, C. A. J., Arora, B., Jacques, D., Kalbacher, T., Kolditz, O., Lagneau, V., Lichtner, P. C., Mayer, K. U., Meeussen, J. C. L., Molins, S., Moulton, D., Shao, H., imunek, J., Spycher, N., Yabusaki, S. B., and Yeh, G. T.: Reactive transport codes for subsurface environmental simulation, Comput. Geosci., 19, 445478, doi:http://dx.doi.org/10.1007/s10596-014-9443-x
Web End =10.1007/s10596-014-9443-x http://dx.doi.org/10.1007/s10596-014-9443-x
Web End = , 2015.
Tang, G., Yuan, F., Bisht, G., Hammond, G. E., Lichtner, P. C., Kumar, J., Mills, R. T., Xu, X., Andre, B., Hoffman, F. M., Painter, S. L., and Thornton, P. E.: Addressing numerical challenges in introducing a reactive transport code into a land surface model: A biogeochemical modeling proof-of-concept with CLM-PFLOTRAN 1.0: Modeling Archive, Next Generation Ecosystem Experiments Arctic Data Collection, Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, Oak Ridge, Tennessee, USA, doi:http://dx.doi.org/10.5440/1239799
Web End =10.5440/1239799 http://dx.doi.org/10.5440/1239799
Web End = , 2016.
Tang, J. Y. and Riley, W. J.: Technical Note: A generic lawof-the-minimum ux limiter for simulating substrate limitation in biogeochemical models, Biogeosciences, 13, 723735, doi:http://dx.doi.org/10.5194/bg-13-723-2016
Web End =10.5194/bg-13-723-2016 http://dx.doi.org/10.5194/bg-13-723-2016
Web End = , 2016.
Tang, J. Y., Riley, W. J., Koven, C. D., and Subin, Z. M.: CLM4-BeTR, a generic biogeochemical transport and reaction module for CLM4: model development, evaluation, and application, Geosci. Model Dev., 6, 127140, doi:http://dx.doi.org/10.5194/gmd-6-127-2013
Web End =10.5194/gmd-6-127-2013 http://dx.doi.org/10.5194/gmd-6-127-2013
Web End = , 2013.
Thornton, P. E. and Rosenbloom, N. A.: Ecosystem model spin-up: estimating steady state conditions in a coupled terrestrial carbon and nitrogen cycle model, Ecol. Model., 189, 2548, doi:http://dx.doi.org/10.1016/j.ecolmodel.2005.04.008
Web End =10.1016/j.ecolmodel.2005.04.008 http://dx.doi.org/10.1016/j.ecolmodel.2005.04.008
Web End = , 2005.
Veuger, B., Middelburg, J. J., Boschker, H. T. S., Nieuwenhuize, J., van Rijswijk, P., Rochelle-Newall, E. J., and Navarro, N.: Microbial uptake of dissolved organic and inorganic nitrogen in Randers Fjord, Estuar. Coast. Shelf S., 61, 507515, doi:http://dx.doi.org/10.1016/j.ecss.2004.06.014
Web End =10.1016/j.ecss.2004.06.014 http://dx.doi.org/10.1016/j.ecss.2004.06.014
Web End = , 2004.
Warren, C. R. and Adams, P. R.: Uptake of nitrate, ammonium and glycine by plants of Tasmanian wet eucalypt forests, Tree Physiol., 27, 413419, doi:http://dx.doi.org/10.1093/treephys/27.3.413
Web End =10.1093/treephys/27.3.413 http://dx.doi.org/10.1093/treephys/27.3.413
Web End = , 2007.
White, M. D. and McGrail, B. P.: Stomp (Subsurface Transport Over Multiple Phase) Version 1.0 Addendum: Eckechem Equilibrium-Conservationkinetic Equation Chemistry and Reactive Transport, Report, 2005.
Yeh, G. T. and Tripathi, V. S.: A critical evaluation of recent developments in hydrogeochemical transport models of reactive multi-chemical components, Water Resour. Res., 25, 93108, doi:http://dx.doi.org/10.1029/WR025i001p00093
Web End =10.1029/WR025i001p00093 http://dx.doi.org/10.1029/WR025i001p00093
Web End = , 1989.
Yeh, G. T., Sun, J., Jardine, P. M., Burgos, W. D., Fang, Y., Li, M. H., and Siegel, M. D.: HYDROGEOCHEM 5.0: a Coupled Model of Fluid Flow, Thermal Transport, and HYDROGEO-CHEMical Transport through Saturated-Unsaturated Media: Version 5.0, Report, 2004.
Zheng, C. and Bennett, G. D.: Applied Contaminant TransportModeling, 2nd Edn., Wiley-Interscience, 2002.
Geosci. Model Dev., 9, 927946, 2016 www.geosci-model-dev.net/9/927/2016/
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Copyright Copernicus GmbH 2016
Abstract
We explore coupling to a configurable subsurface reactive transport code as a flexible and extensible approach to biogeochemistry in land surface models. A reaction network with the Community Land Model carbon-nitrogen (CLM-CN) decomposition, nitrification, denitrification, and plant uptake is used as an example. We implement the reactions in the open-source PFLOTRAN (massively parallel subsurface flow and reactive transport) code and couple it with the CLM. To make the rate formulae designed for use in explicit time stepping in CLMs compatible with the implicit time stepping used in PFLOTRAN, the Monod substrate rate-limiting function with a residual concentration is used to represent the limitation of nitrogen availability on plant uptake and immobilization. We demonstrate that CLM-PFLOTRAN predictions (without invoking PFLOTRAN transport) are consistent with CLM4.5 for Arctic, temperate, and tropical sites.<br><br>Switching from explicit to implicit method increases rigor but introduces numerical challenges. Care needs to be taken to use scaling, clipping, or log transformation to avoid negative concentrations during the Newton iterations. With a tight relative update tolerance (STOL) to avoid false convergence, an accurate solution can be achieved with about 50% more computing time than CLM in point mode site simulations using either the scaling or clipping methods. The log transformation method takes 60-100% more computing time than CLM. The computing time increases slightly for clipping and scaling; it increases substantially for log transformation for half saturation decrease from 10<sup>-3</sup> to 10<sup>-9</sup>mol m<sup>-3</sup>, which normally results in decreasing nitrogen concentrations. The frequent occurrence of very low concentrations (e.g. below nanomolar) can increase the computing time for clipping or scaling by about 20%, double for log transformation. Overall, the log transformation method is accurate and robust, and the clipping and scaling methods are efficient. When the reaction network is highly nonlinear or the half saturation or residual concentration is very low, the allowable time-step cuts may need to be increased for robustness for the log transformation method, or STOL may need to be tightened for the clipping and scaling methods to avoid false convergence.<br><br>As some biogeochemical processes (e.g., methane and nitrous oxide reactions) involve very low half saturation and thresholds, this work provides insights for addressing nonphysical negativity issues and facilitates the representation of a mechanistic biogeochemical description in Earth system models to reduce climate prediction uncertainty.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer