86
Purdue University Purdue e-Pubs Open Access eses eses and Dissertations 4-2016 Quantifying the global N2O emissions from natural ecosystems using a mechanistically-based biogeochemistry model Tong Yu Purdue University Follow this and additional works at: hps://docs.lib.purdue.edu/open_access_theses Part of the Biogeochemistry Commons , and the Environmental Sciences Commons is document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Recommended Citation Yu, Tong, "Quantifying the global N2O emissions from natural ecosystems using a mechanistically-based biogeochemistry model" (2016). Open Access eses. 829. hps://docs.lib.purdue.edu/open_access_theses/829

Quantifying the global N2O emissions from natural

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Quantifying the global N2O emissions from natural

Purdue UniversityPurdue e-Pubs

Open Access Theses Theses and Dissertations

4-2016

Quantifying the global N2O emissions fromnatural ecosystems using a mechanistically-basedbiogeochemistry modelTong YuPurdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_theses

Part of the Biogeochemistry Commons, and the Environmental Sciences Commons

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] foradditional information.

Recommended CitationYu, Tong, "Quantifying the global N2O emissions from natural ecosystems using a mechanistically-based biogeochemistry model"(2016). Open Access Theses. 829.https://docs.lib.purdue.edu/open_access_theses/829

Page 2: Quantifying the global N2O emissions from natural

Graduate School Form30 Updated

PURDUE UNIVERSITYGRADUATE SCHOOL

Thesis/Dissertation Acceptance

This is to certify that the thesis/dissertation preparedBy Entitled

For the degree of

Is approved by the final examining committee:

To the best of my knowledge and as understood by the student in the Thesis/Dissertation Agreement, Publication Delay, and Certification Disclaimer (Graduate School Form 32), this thesis/dissertation adheres to the provisions of Purdue University’s “Policy of Integrity in Research” and the use of copyright material.

Approved by Major Professor(s):

Approved by:Head of the Departmental Graduate Program Date

Page 3: Quantifying the global N2O emissions from natural
Page 4: Quantifying the global N2O emissions from natural

QUANTIFYING THE GLOBAL N2O EMISSIONS FROM NATURAL

ECOSYSTEMS USING A MECHANISTICALLY-BASED BIOGEOCHEMISTRY

MODEL

A Thesis

Submitted to the Faculty

of

Purdue University

by

Tong Yu

In Partial Fulfillment of the

Requirements for the Degree

of

Master of Science

May 2016

Purdue University

West Lafayette, Indiana

Page 5: Quantifying the global N2O emissions from natural

ii

ACKNOWLEDGEMENTS

Firstly, I wish to express my sincere thanks to my advisor Dr. Qianlai Zhuang, for

his continuous encouragement, patience, and immense knowledge, helping me

overcome many difficult situations to finish the thesis. His guidance and advices

widened my horizon on research in various aspects. I also want to give my deep

gratitude to my committee members: Drs. Melba Crawford, Greg Michalski, and

Tonglin Zhang. Thank them for their insightful questions and constructive comments

in different periods of my research. Their advices were thought-provoking, holding me

meet a higher research standard. I am grateful for them to comment on thesis, concern

my progress, and make me a better scientist.

I am also indebted to the members of the Ecosystems& Biogeochemical Dynamics

Laboratory, for their stimulating discussions and technical assistance. Without their

support, I cannot finish the thesis research.

I also want to take this opportunity to thank all the faculty members, for their help

and support of the department of earth, atmospheric and planetary sciences. Especially,

I would like to thank the ITaP staff who maintained the computers in our lab and the

Community Clusters. With their help, I did not need to worry about losing data,

insufficient quota or software problems.

Page 6: Quantifying the global N2O emissions from natural

iii

TABLE OF CONTENTS

Page

LIST OF TABLES ................................................................................................... v

LIST OF FIGURES ................................................................................................ vi

CHAPTER 1. INTRODUCTION ............................................................................ 1

CHAPTER 2. MATERIALS AND METHODS ..................................................... 9

2.1 Overview ....................................................................................................... 9

2.2 Study sites and data source ......................................................................... 10

2.3 Model description ....................................................................................... 10

2.4 Uncertainty analysis and parameterization ................................................. 18

CHAPTER 3 RESULTS AND DISCUSSION ...................................................... 22

3.1 Results .............................................................................................................. 22

3.1.1 Parameters ............................................................................................. 22

3.1.2 Comparison between observation and model simulation of N2O emission

........................................................................................................................ 24

3.1.3 Sensitivity to forcing data and parameters ............................................ 25

3.1.5 Seasonal and inter-annual variations of N2O emissions ........................ 34

3.2 Discussion ........................................................................................................ 48

3.2.1 Environmental factors to N2O emission ................................................ 48

3.2.2 Comparison with previous studies......................................................... 55

3.2.3 Model limitation .................................................................................... 60

CHAPTER 4 CONCLUSION AND FUTURE DIRECTIONS............................. 64

Page 7: Quantifying the global N2O emissions from natural

iv

Page

REFERENCES ...................................................................................................... 66

Page 8: Quantifying the global N2O emissions from natural

v

LIST OF TABLES

Table Page

Table 1. Site information of observational data used for model parameterization 22

Table 2. Key parameters for different vegetation types ......................................... 23

Table 3. Present-day global budget of N2O emissions .......................................... 44

Table 4 . N2O emissions from previous studies ..................................................... 60

Page 9: Quantifying the global N2O emissions from natural

vi

LIST OF FIGURES

Figure Page

Figure 1. Main processed for nitrogen cycle in natural ecosystems ................................... 5

Figure 2. Sensitivity to key climate data: comparison of N2O emission . ........................ 27

Figure 3. Sensitivity analysis for NO3- and NH4

+. ............................................................ 28

Figure 4. Sensitivity of key parameters: comparisons of N2O flux [mg N/(m2 day)]. ...... 32

Figure 5. Global N2O emissions [mg N/(day m2)] ............................................................ 36

Figure 6. N2O emissions during 1900-2000 [mg N/(m2 day)] .......................................... 39

Figure 7. Estimated global N2O emissions from natural ecosystem soils for 1900-2000

(TgN/month). .................................................................................................................... 40

Figure 8. Comparison of monthly N2O emissions with (blue line) and without (orange line)

considering the effects of soil carbon limitation. .............................................................. 42

Figure 9. Nitrous oxide emissions in IPCC report (Tg N2O, Source EDGAR 4.0) .......... 43

Figure 10. Average oil moisture data for July of 2000 (mm/month) ................................ 45

Figure 11. Trends for global precipitation (land only) (mm/year) .................................... 46

Figure 12. (a) NH4 deposition from 1900 to 2000 (Zhuang et al., 2013); (b) N2O emissions

due to NH4 deposition ....................................................................................................... 47

Figure 13. Upper panel: anomalies of precipitation and temperature; Lower panel ......... 53

Figure 14. N2O emissions for different latitudes (% of the total emissions) ................ 57

Figure 15. Comparison for of N2O emission in previous studies (%) .............................. 58

Page 10: Quantifying the global N2O emissions from natural

vii

ABSTRACT

Yu, Tong. M.S., Purdue University, May 2016. Quantifying the Global N2O Emissions

from Natural Ecosystems Using a Mechanistically-based Biogeochemistry Model.

Major Professor: Qianlai Zhuang

Nitrous oxide (N2O) has a great influence on atmospheric chemistry and climate.

It is not only a major greenhouse gas, but also one of the largest ozone-depleting

substances emitted from the biosphere. Process-based biogeochemistry modeling

constrained by site-level observations is a feasible approach to quantifying its emissions

at large temporal and spatial scales. This study developed a process-based

biogeochemistry model of N2O emissions based on an extant biogeochemistry model,

the Terrestrial Ecosystem Model (TEM). The model development includes: 1)

incorporating the effects of physical conditions on both nitrification and denitrification

and 2) implementing principles of the stoichiometry of carbon and nitrogen dynamics

in soils. To simulate the global-scale emission, model parameterizations were first

conducted using observation data at 11 sites. The model was then used to quantify N2O

emissions at these sites for the period during 1980-2010, driven with data of climate,

soils, vegetation, and topography. Finally, the parameterized model simulated global

soil N2O emissions for the 20th century. The model sensitivity to parameters and major

input data was also conducted. Parameters rpctp, bpctp, Kntf, k and pn2 are influential in

model simulations and the variation of rpctp, bpctp (parameters related to precipitation

Page 11: Quantifying the global N2O emissions from natural

viii

and soil moisture) and Kntf (rate for nitrification) are large. Considering the carbon

limitation, a sensitivity analysis suggested that the N2O emissions can change 10-50%

depending on sites by varying climate inputs by 20%. Without considering carbon

limitation in the model, N2O emissions are increased by 10-30%. These sensitivities

vary over season and region. Simulations indicate that there is a slightly decreasing

trend form 1900 to 2000 for global N2O emissions. The emissions in 1900 are 1.92 Tg

N2O month-1 in summer (July) and 1.46 Tg N2O month-1 in winter (February) , while

in 2000 are 1.30 Tg N2O month-1 in summer (July) and 1.07 Tg N2O month-1 in winter

(January) respectively. Among all the 11 vegetation types, boreal forests contribute to

the most emission because they occupy a large area (45oN north). Considering different

latitudes, tropical areas (23.5°N~ 23.5°S) emit more than half of the global emission,

because of their warm temperature and moist soils. My future study will improve the

representation of microbial dynamics including microbial traits and their effects on

N2O emissions in the model. The model will then be used to analyze how global N2O

emissions from natural ecosystem soils will change during the 21st century.

Page 12: Quantifying the global N2O emissions from natural

1

CHAPTER 1. INTRODUCTION

Nitrous oxide (N2O) plays an important role in atmospheric climate and chemistry,

along with other greenhouse gases (e.g. CO2 and CH4). It is a natural source of NOx

(NO+NO2) in the stratosphere and NOx are key chemical compounds to reduce ozone

in the atmosphere (Portmann et al., 2012). N2O contributes significantly to the non-CO2

greenhouse effect as the third important greenhouse gas behind CO2 and CH4.

Approximately 5% of global warming potentials are due to N2O in greenhouse effect

assessment (Rodhe 1990). The warming potentials of unit volume of N2O is about 300

times of unit volume of CO2.

The major sink for N2O in the atmosphere is in the stratosphere where it is converted

into NOx, which causes the depletion of ozone through chain reactions (Equation 1.

1-1.4) (Nevison and Holland 1997). N2O is very stable in troposphere with no

photochemical sinks and has a lifetime of ~114 years (Brown et al. 2013). In the

stratosphere, about 5% N2O reacts with O(1D) and the produce reducing the ozone

concentration via chain reactions (Warneck et al.1999):

(1.1)

(1.2)

(1.3)

3 22O O O (1.4)

1

2 2N O O D NO

3 2 2NO O NO O

2 2NO O NO O

Page 13: Quantifying the global N2O emissions from natural

2

In the net reaction (1.4), a mole of N2O consumes 2 moles of O3. A better estimate

of N2O emission or observation will improve the quantification of its influence on

ozone depletion via these chain reactions.

Sources of N2O are found on land and in the ocean, and are both natural and

anthropogenic. Approximately, 16 TgN/yr N2O was emitted globally in the 1990s,

resulting in concentration of 310 ppb by volume, and have increased by 0.2% to 0.3%

per year (Machida et al. 1995). Compared to ocean, land contributed most of the

emissions (Thompson et al. 2014). Among the global emissions, 30- 50% of emissions

were a result of human activities, while almost 60% of the total emissions are through

microbial production (Reay, et al. 2012). However land surface emission estimates still

have large uncertainties, especially for daily fluxes.

To date, many studies have suggested that cultivated land plays an important role

in emissions of N2O, which is mostly due to fertilizer implementation in cropland. In a

traditional view, the highest terrestrial N2O emissions are observed from croplands and

are mainly associated with agronomic management of irrigation and fertilization

(Kimble et al. 1998; Xing 1998; Yan et al. 2003; Yuesi and Yinghong 2003). Meanwhile,

in natural environment, soils have acted as a sink of organic carbon (Eswaran et al.

1993; Lal 2004) and organic nitrogen (Nadelhoffer et al. 1999) for thousands to millions

of years, especially the land-use type with weak decomposition. The storage of organic

carbon and nitrogen supplies a plenty of substrates for nitrification and denitrification.

Although agronomic activities are important source of N2O, natural sources cannot be

ignored. At the global scale, other ecosystem typesmay also be essential sources of N2O.

The release of N2O is highly dependent on physical and biological conditions of soils,

Page 14: Quantifying the global N2O emissions from natural

3

such as soil temperature, soil moisture, soil acidity, and the activity of related microbes.

Temperature leads to the seasonal variation of N2O emissions, which is also a reason

for the regional difference (Smith et al. 1998). Soil moisture determines the rates of gas

destruction via oxidation. High water content leads to low oxygen content, which is

associated with reducing environment; whereas less water leads to an oxidizing

environment. Soil acidity controls the chemical reactions involving H+ or OH-, both of

which is influence the activity of enzyme. In natural ecosystems, tropical soils emit the

largest amount of N2O among various ecosystems because of suitable temperature and

adequate precipitation for whole year(Kroeze et al. 1999, Werner et al. 2007). In

contrast, N2O emissions originating from areas of high latitudes are viewed as an

insignificant source (Martikainen et al. 1993, Potter et al. 1996) due to slow

mineralization rate under low temperature, humid condition and low atmospheric

deposition of nitrogen (Dentener et al. 2006).

Processes of nitrogen cycle consist of ammonification, immobilization,

decomposition, nitrification, denitrification, N2 fixation, and others (Figure 1). Three

sub-cycles are important to nitrogen balance in ecosystems: elemental, autotrophic and

heterotrophic cycles (Ayres 1996). The elemental cycle is related to the variation of

oxidation state. Autotrophic cycle is driven by nitrogen uptake by plants. The

heterotrophic cycle is linked to the decomposition process, driven by the need of

heterotrophic organisms for carbon, which supports microbial living and propagating.

When considering this cycle, a microbial biomass pool is necessary. We need to

consider three nitrogen pools: Atmosphere (NH4 +NO3-), vegetation N, and soil N,

Page 15: Quantifying the global N2O emissions from natural

4

including soil organic nitrogen (microbial biomass N and litterfall) and inorganic

nitrogen (nitrate and ammonia).

For a proper simulation of nitrogen cycle, it is necessary to incorporate key

processes and N pools into N biogeochemical models. Here I ignore the dinitrogen (N2)

pool and only consider the deposition of NH4, considering the availability of data in

global scale. There are two processes in producing N2O including denitrification by

anaerobic bacteria and nitrification by aerobic bacteria (Robertson and Kuenen 1990).

Plants mainly uptake NO3- and NH4

+ from soil available mineral nitrogen pool, and

some of them can also absorb small organic molecules containing N (Jones and Hodge,

1999).

Denitrification is conducted by anaerobic bacteria, in which nitrate is used as

electron acceptors by microbes and decomposed organic matters as electron donors. In

this process, NO3- is reduced into NO, N2O and N2 step by step:

3 2

3 2 2

3 2 2

4 3 2

2 10 8 5

2 12 10 6

NO H e NO H O

NO H e N O H O

NO H e N H O

(1.5)

Temperature, soil moisture, acidity, and available soil carbon influence the rate of the

whole process (Bremner and Shaw 1958).

Nitrification is the process of biological oxidation of ammonia with oxygen into

nitrite and nitrate by aerobic bacteria (1.6):

(1.6)

4 2 2 2

2 2 3

2 3 2 2 4

2 2

NH O NO H O H

NO O NO

Page 16: Quantifying the global N2O emissions from natural

5

The first step is the rate limiting step for nitrification. In natural ecosystem, each step

is conducted by species of ammonia oxidizing bacteria and archaea. The processes are

more complicated and limited by many factors in reality.

Figure 1. Main processed for nitrogen cycle in natural ecosystems: 1-ammonification; 2-

immonilization; deposition ;3-autotrophic nitrification ; 4-plant uptake; 5-nitrate immobilization;

6-denitrification ; 7-dissimilatory nitrate reduction to ammonium; 8-decomposition; 9-N2

fixation; 10-deposition

4 10

Page 17: Quantifying the global N2O emissions from natural

6

Nitrification can be initiated when large organic amine containing molecules are

decomposed into small ammonia or ammonium molecules. Ammonium is then

oxidized by the enzyme mono-oxygenase into NH2OH. Then NH2OH can be oxidized

into NO, N2O and NO3-. Regarding the nitrification pathway, only a small amount of

denitrified N is lost as N2O, from 0.5% (Engel and Priesack 1993) (Expert-N) to 2%

(Parton et al. 1996). Although NH3 is assumed to be substrate in the second reaction,

NH4+ is the major reactant when pH is less than 8.

(1.7)

This decomposition sequence shows that the stoichiometry of C, N, P and S in the

environment is important for the biogeochemical cycling of nitrogen, which thus may

be highly influenced by the carbon exchange between soils, vegetation and the

atmosphere. This is due to the relatively stable carbon: nitrogen: phosphorus: sulfur

ratios in microbial biomass (Kirkby et al. 2011). Studies indicate that there are

approximately consistent C:N:P ratios for soils and soil microbes (e.g., Cleveland and

Liptzin 2007, Griffiths et al., 2012). For soil, C: N is about 12 and C: P varies in a wider

range from 150-300. For microbes, C: N varies from 7 to 8.6 and C: P is from 30 to 35

in most cases. Soils have complex structure and high species diversity (Rösch et al.

2002, Young and Crawford 2004), but the principle of stoichiometry can be applied to

most ecosystems. Stoichiometry rests on the basic law, indicating that the ratio of C and

N in organic matters will maintain within a certain range in organisms and tissues. If

there is a large amount of C release from Arctic soils, the N release should follow this

release. For example, a peats circle with low C: N allows extensive nitrogen

2 2 2 3 2

3 2 2 2

2 2 2

2

2 2

2 4 4

NH CONH H O NH CO

NH O H e NH OH H O

NH OH N O e H H O

Page 18: Quantifying the global N2O emissions from natural

7

mineralization, nitrification, and N2O emissions (Susloto et al. 2009). The high bulk

density and intermediate water content in the surface of peatlands provides both aerobic

and anaerobic microenvironments (Repo, et al. 2009), leading to a good environment

for both nitrification and denitrification. To study the nitrogen balance in all kinds of

natural ecosystems, the nitrogen cycling should be coupled with carbon cycling

following the C and N stoichiometry.

Although terrestrial budgets of N2O have been developed, the global N2O emission

budget remains uncertain for the 20th century. It is possible to measure N2O emissions

from soils directly, but they have high spatial and temporal variatbility, so it is difficult

to extrapolate N2O measurements to an entire region. It is debatable to use a few site-

level measurements to generalize the regional emissions with a simple approach that

simply multiply site level fluxes by area. In order to overcome this limitation I

propose to use a process-based biogeochemistry model to quantify N2O emissions at

the global scale. The model is couples soil thermal and moisture physical models and

uses remote sensed data of land cover, soil temperature, and soil moisture as drivers of

N2O emissions. To date, several process-based models to quantify N2O fluxes exist: a

version of Terrestrial Ecosystem Model (TEM) for biofuel crop ecosystems (Ag-TEM)

(Qin et al. 2013a, b), the Community Land Model, Carbon and Nitrogen cycles (CLM-

CN)(Saikawa et al. 2013), the daily Century (DayCent) model (William et al. 1998,

Delgrosso et al. 2005), and the Denitrification/ Decomposition (DNDC) Model (Li

1992). Estimating Carbon in Organic Soils - Sequestration and Emissions (ECOSSE)

Model (Smith et al. 2010a & 2010b), a whole-farm model of C and N flows and of farm

production (FASSET)(Chatskikh et al. 2005), and a water and nitrogen management

Page 19: Quantifying the global N2O emissions from natural

8

model (WNMM) (Li et al. 2005) are also availableFurther, although most of the models

do well in predicting N2O emissions in tropical and temperate areas, these models are

not suitable for high latitude regions because they often do not consider the effects of

permafrost thawing and complex hydrological dynamics on biogeochemistry in the

region. My approach to quantifying N2O emissions was based on an extant

biogeochemistry model of C and N that has been coupled with a soil thermal model and

hydrological model (Zhuang et al., 2001, 2003, 2004), aimed to solve the problems

mentioned above.

Page 20: Quantifying the global N2O emissions from natural

9

CHAPTER 2. MATERIALS AND METHODS

2.1 Overview

The performance of different N2O emission models was found to be variable

(Saikawa et al. 2013, William et al. 1998, Delgrosso et al. 2005, Chatskikh et al. 2005).

Most process-based models treat nitrification and denitrification as simple chemical

reactions instead of considering the effects of microbial nitrifiers and denitrifiers and

the coupling of C and N. The Denitrification/ Decomposition (DNDC) Model is the

only model that uses denitrifier biomass as a predictor of N2O emission rate. Microbes

are the linkage of ecology and biogeochemistry, which link C and N cycling in soils

and are responsible for nitrification and denitrification, thus their inclusion is N2O

models is essential. Also, microbes need carbon, which is a main element in their body

and supplies energy. The key goals of trait-based ecosystem models are to consider how

the diversity of organisms is formed and shaped by the environment, including

interactions with other organisms, and how various traits affect the ecosystem processes

and biogeochemical cycling. These models organize life according to ‘functional traits’,

properties of a group of organisms that are essential to their success and functioning in

a particular environment (Allison 2012). Trait-based models give a more detailed

description of biological process, but usually need more data to evaluate all the

processes in comparison with traditional biogeochemical models. Here I also

incorporated the microbial traits to a certain

Page 21: Quantifying the global N2O emissions from natural

10

degree in addition to considering stoichiometry of C ad N of the ecosystems into N

cycle modeling.

2.2 Study sites and data source

Eleven sites were selected globally in order to represent 11 vegetation types (Table

1). Observational data were all obtained from previous publications. The observational

data include all representative ecosystem types including tundra, forests, and grasslands

with varieties of climate condition during different time periods. The emission data

were used as the monthly average over the measurement period, ranging from several

days to several years. Data of soil density were also collected from the publication if

given, or the average density for this soil type was applied. Climate data including air

temperature, water vapor pressure, precipitation, and cloudiness is obtained from the

Climate Research Unit (CRU)(Mitchell 2005).

2.3 Model description

Previous efforts in modeling carbon cycle and part of nitrogen cycle are based on

TEM modeling framework (Zhuang et al. 2004, 2006). The Terrestrial Ecosystem

Model (TEM) is a global-scale biogeochemical model for estimating C and N cycling

in terrestrial ecosystems. Recent developments of TEM were made towards fully

accounting for N cycling in managed ecosystems (Qin et al. 2013, Qin et al. 2013).

The algorithms for modeling C and N transformation were inherited from the original

TEM. In my new development, I assume that nitrogen in soil mainly comes from NH4

from atmospheric deposition and litter fall. The dynamics of state variables, such as

amount of carbon in vegetation and soil, amount of nitrogen in vegetation and soil, are

Page 22: Quantifying the global N2O emissions from natural

11

also expressed in a similar way as TEM. TEM uses five differential equations for the

predicting temporal changes in carbon and nitrogen:

(2.1)

(2.2)

(2.3)

(2.4)

(2.5)

Where Nv, Ns, Nav, are the nitrogen pool of vegetation, organic nitrogen in soil and

available nitrogen (g N m-2). Cv and Cs are the vegetation and soil carbon pools (g C m-

2). NUPTAKEt, NETNMINt, NLOSTt and NINPUTt are the nitrogen uptake by

vegetation, net soil mineralization of organic nitrogen, nitrogen lost from ecosystem,

and nitrogen (NH4+) input from atmosphere (g N m-2 day -1), respectively. LNt and Lct

are the nitrogen and carbon in litter fall (g N m-2 day -1). GPP is the gross primary

production. RAt and RHt are autotrophic respiration and heterotrophic respiration (g N

m-2 day -1) respectively. NETNMIN is modeled depending on RHt, which is soil carbon

decomposition. This means carbon cycle will impact on the nitrogen cycle in the model.

Specially, the amount of mineralized N is the substrate for nitrification and

denitrification.

Nitrogen deposits on the earth surface and converted into organic amine by

nitrogen-fixing microorganisms. Organic amines are decomposed into NH4+ which can

be nitrified into NO3- and both (NH4

+, NO3-) are available for plant uptake, and the latter

vt Nt

dNNUPTAKE L

dt

SNt t

dNL NETNMIN

dt

avt t t t

dNNINPUT NETNMIN NLOST NUPTAKE

dt

vt At Ct

dCGPP R L

dt

sCt Ht

dCL R

dt

Page 23: Quantifying the global N2O emissions from natural

12

can be denitrified. Most gaseous N emission from the soil is in form of N2, NO and N2O

are produced via this coupled nitrification and denitrification.

The change of ammonium and nitrate with time are calculated in the model using:

4t t nitf t ammonium

dNHNINPUT NETMIN N NUPTAKE r

dt (2.6)

(2.7)

Where NETNMINt is the mineralization rate in soil, NUPTAKEt is the nitrogen

taken by plants. Nnitf and Ndenit is the rate for nitrification and denitrification. rammonium

and rnitrate are the percentage of ammonium and nitrate in soils (2.8 and 2.9). Where NH4

is the content of ammonia N and NO3 is the content of nitrate N in the soil.

(2.8)

(2.9)

Ammonium N enters the soil in three ways: from mineralization of soil organic N,

deposition from the atmosphere and from fertilizers. In natural environment,

mineralization is viewed as the main source while deposition also contributes a small

portion, and fertilizer is nil. Ammonium N is nitrified by a first order process, which is

dependent on several environment factors, such as soil temperature, water content, and

pH:

(2.10)

3nitf denit t nitrate

dNON N NUPTAKE r

dt

4

4 3

ammonium

NHr

NH NO

3

4 3

nitrate

NOr

NH NO

4

4

4

1 2 3

max

1 exp 1NH

n NH ntf

NH

NN N k f T f W f pH

N C

Page 24: Quantifying the global N2O emissions from natural

13

Where Nn is the quantity of nitrified N formed at one time step (g N m-2day-1).

NNH4is the quantity of ammonium present in the soil layer. kntf is the rate constant

for nitrification and is a function of temperature, but for convenience is set as 0.6

(Bradbury et al. 1993). Because kntf is an essential rate for nitrification, I adjusted it

for a better simulation during parameterization. f(T), f(W), f(pH) are functions of

temperature, water content and pH in soil, respectively. They are the main factors

affecting nitrification rate.

According to the model of Kettunen (2003), the temperature rate is calculated:

(2.11)

Where Tair is the air temperature (oC) and the result is as unit less.

Below field capacity, the effect of soil water content follows the relationship given

by Bradbury, et al. (1993) and Smith et al (2001). The water modifier, f(w) (mm/layer),

is calculated as:

, when (2.12)

, when

(mm/layer) is the amount of water held between field capacity and the

permanent wilting point. (mm/layer) is the calculated soil water content, which can

be simulated by Brooks-Corey model or Van-Genuchten model. These are two models

calculating suction head of water with soil water content. is the soil moisture deficit

1 106

18.3

47.9( )

1 exp airT

f T

0

2

1( ) 1

f c i

f i

ff W

f c i

2( ) 1f W f c i

f

c

i

Page 25: Quantifying the global N2O emissions from natural

14

at the permanent wilting point (-1bar). is the soil water rate modifier at permanent

wilting point and saturation.

In order to avoid the complex parameters in above hydrological model, a linear

relationship between soil water content and precipitation (P) is applied to estimate the

effect of soil moisture, which is expressed as:

(2.13)

Where, r (mm) and b are parameters, which need to be adjusted according to

different soil types in the process of parameterization. The initial value r is set to 100,

b as 0 according to the approximate relation between precipitation and soil moisture.

f2(w) is the water filled pore space (%), which is determined by water content in total

pore space.

With the participation of H+ in nitrification, pH plays an important role in affecting

NH4+ and NH3. When the pH is less than 8, which is normal for most soil, NH4

+ is the

major form. In the range of 8~9, the ratio of NH3/ NH4+ is sensitive to the concentration

of H+. Nitrifiers are also sensitive to pH. At 68°F, the nitrifiers are the most efficient at

the pH of 8~8.5. When the pH reaches 6, the nitrification rate is only about 12% of the

maxima. And the nitrification rate is about 38% of the maxima with the pH of 10. pH

may hinder the reaction if not enough or too much alkalinity is present (Sawyer et

al.,1974). The relationship between nitrification rate and pH shows a linear response:

(2.14)

0f

2

Pf W b

r

min3 3 min 3 min

max min

( ) ( ) (1 ( ) )pH pH

f pH f pH f pHpH pH

Page 26: Quantifying the global N2O emissions from natural

15

is set to 0.2. pHmin and pHmax are site-specific based on vegetation and soil

types, but for the global simulation these values were set to 2 and 5.

When the pH of soil is no less than 5, the equation can be simplified (Parton, et al.

1996):

(2.15)

H+ is produced during the process of nitrification. However, soil can be viewed as

a huge buffer. Meanwhile, many reactions take place in the soil. Some of them

produce H+, and some of them consume it. The variation of H+ in these reactions will

lead to slightly change of pH. For a simple calculation, pH of soil is treated as a

constant.

Among all the nitrified nitrogen, only a small portion of them are released as gas

into the atmosphere. These gases are N2, NO and N2O. The N2O emission caused by

nitrification is estimated as:

𝑁𝑁,𝑁2𝑂 = [𝑁𝑔𝑎𝑠 × (1 − 𝑁𝑛𝑜,𝑛2)] × 𝑁𝑛 (2.16)

Where Ngas is set to 0.02, because during nitrification, only 2% of N is assumed to

be lost as gas. Nno is a proportion of NO as the released gas, which is set to 0.4 (Davison

et al., 2000). And Nn here is the same as in equation 2.10.

Water content determines the flux of oxygen into soil. Denitrification for each layer

in the model during each time step is calculated using (Hénault and Germon 2000):

(2.17)

Where Nd is the actual denitrification rate (g N m-2 layer-1day-1). Dp is the potential

denitrification rate (g N m-2 layer-1) that is a function of soil texture and soil biota, which

min( )f pH

1

3

tan 3.14 0.45 5( ) 0.56

3.14

pHf pH

1 3 2 3 4( ) ( ) ( ) ( )d pN D g NO g W g B g T

Page 27: Quantifying the global N2O emissions from natural

16

are both site specific parameters (Hénault and Germon 2000). Dp is a site specific

parameter in the model and it was set as 0.2 g N m-2 day-1. So within a land use type, I

can use certain values for calculation. g(NO3), g(W), g(B), and g(T) are denitrification

response factors to soil nitrate content, water-filled pore space (Linn and Doran 1984),

biological activity and soil temperature, respectively.

g(NO3) is a denitrification factor:

(2.18)

where NNO3 is the amount of NO3- in the soil. Nd50 is the soil nitrate content at which

denitrification is 50% of full potential (g N m-2). Nd50 is 16.5 g N m-2 based on the soil

nitrate content when denitrification is 50% of its full potential in experiments (Hénault

and Germon 2000).

g(W) is the denitrification response factor to soil water content (Grundmann and

Rolston, 1987):

, when water-filled pore space ≥0.62 (2.19)

g2(W)=0, ,when water-filled pore space <0.62

θc and θf are the same as in Equation (2.12).

g(T) is the temperature factor, and the temperature (T) is in centigrade (°C). The

total rate is highly sensitive to the variation of temperature as it is an exponential

function.

, if T<60 (2.20)

3

3

1 3

50

( )NO

NO d

Ng NO

N N

1.74

2

0.62

( )0.38

c

fg W

c

f

c

f

( 22.5)/10

4( ) 2 Tg T

Page 28: Quantifying the global N2O emissions from natural

17

g4(T)=0, if T≥60

Based on the relationship developed by Bradbury et al. (1993):

(2.21)

Considering the limitation of organic carbon in soils, I improve the existing

equations based on stoichiometry between carbon and nitrogen. Carbon sources provide

energy and affect the microbial productivity and activity. When there is equilibrium

between input and output, there should be a low community respiration per unit

microbial biomass, which is measured by CO2. In this equation, I use the respiration

CCO2 indicating the activity of microbes and consumption of organic carbon. Organic

carbon is important here because it acts as the electron donor in denitrification. This

process will consume carbon. The developed N2O model was integrated to the TEM

modeling framework and soil thermal and hydrological models (Zhuang et al. 2001,

Zhuang et al. 2002, Zhuang et al. 2004).

The N lost by denitrification is then divided into N2O and other products (N2 and

NO). The calculation for the amount of N2O gas produced by denitrification (Nd) is

given as:

(2.22)

Where h(W) (%) is defined :

(2.23)

The proportion of (N2+NO)/(N2O+N2+NO) (DN2) at field capacity is determined by

water content in soil (Parton et al. 1996). It was set as 0.5 originally, and adjusted for

23 c( ) +0.06COg B k C

2, 1 2 3[1 ( ) ( )]d N O dN h W h NO N

21( ) cN

f

h W D

Page 29: Quantifying the global N2O emissions from natural

18

different vegetation types in parameterization. θc and θf are the same as in Equation

(2.12).

Where h2(NO3) (%) is calculated as:

3

3 3

2 3( ) 1NO

NO NO

Nh NO

N p

(2.24)

Where pNO3 is the soil NO3- content in the soil at which N is release in equal amount

of N2O and other gases (Bell et al. 2011). It was set at 20 g N m-2.

2.4 Uncertainty analysis and parameterization

The difference between observations and model simulation may be caused by model

itself because it does not fully represent the ecosystem. It is also possible to be led by

the extreme weather condition or abnormal climate, such as tornado, flood or drought,

which have not been considered in the model simulations. The observational data will

also have uncertainties and errors due to instrument errors or measurements under

extreme climate conditions (e.g., flooding, strong wind). Besides, the sudden variability

of climate at the temporal resolution less than our resolution is also a possible cause.

Because a model is a simplification of the studied ecosystems, there will always be

differences between simulation and observation data. In addition to the imprecision in

observation, uncertainties in models mainly come from three sources: the error in the

model structure, the uncertainty from hydrological or some other time-series inputs,

and the parameters associated with processes of the ecosystem. For the part of input

data error, I need to quantify the confidence of the simulation using observation data.

Because of the limited amount of observational data, the confidence is estimated by

calculating the total error, which is the root mean squared error (RMSE) between

Page 30: Quantifying the global N2O emissions from natural

19

simulations and observations. Key parameters are also adjusted to have the least RMSE,

the average of the distance between observational data and simulations:

𝑅𝑀𝑆𝐸 = √𝑀𝑆𝐸 = √1

𝑛∑ (𝑥𝑠𝑖𝑚(𝑖) − 𝑥𝑜𝑏𝑠(𝑖))

2𝑛𝑖=1 (2.25)

Parameterization is applied to optimize the processes of physiological processes in

the model. Concerning nitrogen cycling in TEM, the initial value of each parameter was

chosen by conducting literature review. All parameters in equations are achieved based

on data and knowledge from experiment or field research. But when applying the

equations for a large area, parameters obtained from field experiment at a small scale

may be no longer valid. In the process of parameterization, the initial value was set as

the mean, ranging approximately from half to double of the mean. For a better

simulation, 5 key parameters were selected and adjusted within a certain range, which

are r and b in the relationship between precipitation and soil moisture, Kntf as the rate

constant for nitrification, k as the potential denitrification rate, and pn2 as the proportion

of N2 among emission gases. It is assumed that there is no correlation between these

parameters and all parameters are assumed to be uniform distributed. I choose them for

the following two reasons. First, they are crucial as rate factors, which means a small

variation of them may lead to a large result difference. Another important reason is that

no certain values for these parameters are available based on lab or field experiments.

The parameters are set as mean value for each of 11 representative vegetation types and

the parameter uncertainty is also obtained for each vegetation type.

Simulation uncertainties can also be resulted from the uncertain mode structure. It

may be led by the lack of information, or the improper simplification of some influential

processes. This part of error is not estimated because it cannot be solved by

Page 31: Quantifying the global N2O emissions from natural

20

parameterization. It is possible to quantify this uncertainty by varying the complexity

of the model (e.g., He et al., 2014).

To test how the simulations will change with physical conditions of environment, I

calculate their correlations and covariance between modeled emissions and temperature

and precipitation. The significance is determined using t-test:

𝜌𝑛2𝑜,𝑇 = 𝐶𝑜𝑟𝑟(𝑁2𝑂, 𝑇) =𝑐𝑜𝑣(𝑁2𝑂,𝑇)

𝜎𝑁2𝑂𝜎𝑇=

𝐸[(𝑁2𝑂−𝜇𝑛2𝑜)(𝑇−𝜇𝑇)]

𝜎𝑁2𝑂𝜎𝑇 (2.26)

𝜌𝑛2𝑜,𝑃 = 𝐶𝑜𝑟𝑟(𝑁2𝑂, 𝑃) =𝑐𝑜𝑣(𝑁2𝑂,𝑃)

𝜎𝑁2𝑂𝜎𝑃=

𝐸[(𝑁2𝑂−𝜇𝑁2𝑂)(𝑇−𝜇𝑃)]

𝜎𝑁2𝑂𝜎𝑃 (2.27)

is the correlation coefficient between nitrous oxide emission and physical

conditions. indicates high correlations and indicates low correlations,

Cov is the covariance, is the standard deviation. is the mean of each variable.

E[ ] is the expectation. T and P are anomalies of temperature and precipitation. The

anomalies are the differences between each value and the average of certain month or

year.

The positive covariance means a similar trend for emission and conditions, and

negative covariance means an opposite trend. The values of covariance indicate the

tendency of N2O emission in changing of the main physical conditions.

1 0

Page 32: Quantifying the global N2O emissions from natural

22

CHAPTER 3 RESULTS AND DISCUSSION

3.1 Results

3.1.1 Parameters

Table 1 displays the set of parameters for 11 land use types. It shows great variations

for rpctp, bpctp and Kntf, but similarities for k and pn2. rpctp and bpctp, which determine the

relationship between precipitation and soil moisture (water filled pore space) as shown

in equation (2.13). The variations of these two parameters are led by hydrological

situation for different vegetation types, or soil types. Soil moisture is not only

influenced by the intensity of rainfalls, but also by soil texture, consistent and other

characters. rpctp is large for grassland and forest, which means more precipitation is

needed for increasing soil moisture. Tropical forests have the largest r, as a result of the

most complicated rooting system compared to other types. Areas with dry climate or in

high latitudes usually have a smaller rpctp. Kntf is related to temperature. The other two

parameters are less influential compared to the first three, because the empirical values

are universal for all vegetation types. k and pn2 are similar for all vegetation types,

which doesn’t mean they are not influential for nitrogen gas emissions. k relates the

amount of soil organic carbon to nitrogen reactions, showing the effect of soil microbes

in the process of denitrification. The content of organic carbon in soils plays a similar

role in the nitrogen cycle for all ecosystem types, the differences among ecosystem

types are only due to the amount of carbon. The last

Page 33: Quantifying the global N2O emissions from natural

22

Table 1. Site information of observational data used for model parameterization

Site

Vegetation type (Longitude,

latitude) Location

Observation

time period

Comment

1 Grassland -106.0 41.0 , 6/92-10/93 Glacier lakes Experiment

Ecosystem

Wyoming, US

2 Tropical forest -67.0 18.0,

Caribbean

11/92-2/95 Mosier and Delgado (1997)

3 Temperate

coniferous forests

9.5 51.5,

Germany

6/91-10/91 Schulte‐Bisping (2003)

4 Temperate

deciduous forests

5.5 51.5,

Netherland

5/97-3/98 Mammarella et al (2010)

5 Xeric shrub lands -109.0 41.0,

Colorado, US

5/97-3/98 Matson et al (1991)

6 Tundra 63.0 67.0,

Russia

6/07-10/07 Mammarella et al (2010)

7 Boreal forest 30 61.5,

Russia

6/00-8/03 Maljanen et al (2006)

8 Wet tundra -111.5 65.0,

Canada

2005/06 Buckeridge et al (2010)

9 Temperate evergreen

broad leaf

120.0 30.5,

China

6/08-6/09 Liu et al (2011)

10 Mediterranean shrub

lands

-118.0 34.0,

California, US

7/89-12/89 Anderson and Poth (1989)

11 Xeric woodland -104.5 41.0,

Colorado, US

1/95-1/96 Central Plains Experimental

Range, LTER Colorado

Page 34: Quantifying the global N2O emissions from natural

23

Table 2. Key parameters for different vegetation types

Vegetation Type

rpctp

(mm/month)1

bpctp

(mm/mm)2

Kntf

(g/g)3

K

(per unit)4

pn2

(g/kg)5

Grassland 150 0.4 0.2 0.026 0.5

Tropical forest 190 0 1 0.026 0.5

Temperate coniferous forests 170 0.5 0.6 0.026 0.5

Xeric shrublands 50 0.5 0.3 0.026 0.5

Tundra 90 0 1 0.026 0.5

Boreal forest 50 0.4 0.4 0.020 0.1

Tundra 70 0.5 0.2 0.020 0.1

Wet tundra 50 2 0.9 0.026 0.5

Xeric woodlands 50 0.4 0.1 0.020 0.1

Temperate evergreen forest 70 0.6 0.1 0.020 0.1

Mediterranean shrublands 130 0 0.7 0.020 0.5

1 2

3 Rate for nitrification 4 Effect of organic carbon for denitrification 5 Determine the ratio of N2 and N2O for denitrification

__

precipitationsoilmoisture b pctp

r pctp

Page 35: Quantifying the global N2O emissions from natural

24

parameter, pn2 is a factor influencing the ratio between N2 and N2O in the process of

denitrification, which is dependent on soil types, but has a similar effect on nitrogen

cycling for all vegetation types.

3.1.2 Comparison between observation and model simulation of N2O emission

After parameterization, I obtained an optimized set of parameters for each

ecosystem type. The model performs better in temperate areas for deciduous forests,

evergreen forests, grasslands or shrubland. But the model did not do well for high

latitudes ecosystems like tundra, wet tundra, and boreal forests. For some sites, there

is a high variation for observational data, but for others, with similar conditions, there

are no much variations. Thus, model simulations do not match all observed trends.

Simulations for tropical forests did not show much seasonal variations, while

observations vary significantly. Seasonal variations are mostly due to temperature

change, while precipitation impacts less. In tropical areas, the variation of emissions is

due to soil water content to some extent.

The discrepancy between simulations and observations are due mainly to two

reasons. The most important one is the experimental error of observational data.

Because of the low flux of nitrous oxide, its measurement is still difficult. The

uncertainty of measurement can even be several times of the observational mean data,

especially for data measured in the 1980s. The second reason is that the small amount

of observation data are available, which does not constrain the model parameters well.

Page 36: Quantifying the global N2O emissions from natural

25

3.1.3 Sensitivity to forcing data and parameters

To test the model sensitivity to major forcing data, I changed the forcing data by

20%, including monthly air temperature, precipitation, water vapor pressure and

cloudiness (Figure 2). When all above are changed by 20%, no matter increase or

decrease, N2O emissions show large differences from 10% to 50%. Figure 2(a-d) shows

that decreases in air temperature, precipitation, water vapor pressure and cloudiness led

to more N2O emissions with larger fluctuations. While higher temperature, more

precipitation, higher water vapor pressure and cloudiness give smaller differences

between the highest and lowest emissions. The variation of N2O depends on reactants

and reaction conditions. The comparison of NO3- and NH4

+ emissions indicates that the

differences of N2O are largely contributed to the amount of reactants (Figure 3). NO3

turns into N2O through the process of denitrification and NH4 turns into N2O via the

process of nitrification. Reaction conditions are also key factors in the case of the same

amount of reactants. The variation of forcing data doesn’t lead to a linear change of

N2O emissions. If changing the forcing data one by one, from Figure 2(b) and 3(d), 20%

higher air temperature and more water vapor pressure have similar effects on N2O

emissions.

Parameters rpctp, bpctp and Kntf, k and pn2 are also changed to test the model

sensitivity. Each parameter value is altered by 5%, the results are shown in Figure 4.

During the parameterization, k and pn2 are almost the same for all vegetation types.

But when changed by 5%, they also have noticeable effects on N2O emissions (1%-

2%). When the parameters are changed by 1%, 5%, 10%, and 20%, respectively, the

difference of N2O emissions is larger as the change of parameter is larger. The degree

of output change is positively correlated with the degree of parameter change.

Page 37: Quantifying the global N2O emissions from natural

26

(a) Air temperature

(b) Precipitation

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

Page 38: Quantifying the global N2O emissions from natural

27

(c) Cloud

(d) Water vapor pressure

Figure 2. Sensitivity to key climate data: comparison of N2O emission [mg N/(m2 day)]: (a) Comparison

of N2O with original data set (blue line), 20% higher air temperature (orange line); (b) Comparison of

N2O with original data set (blue line), and 20% more precipitation (orange line); (c) Comparison of N2O

with original data set (blue line), and 20% more clouds (orange line); (d) Comparison of N2O with

original data set (blue line), and 20% more water vapor pressure (orange line).

0

0.5

1

1.5

2

2.5

Jan

-80

No

v-8

0

Sep

-81

Jul-

82

May

-83

Mar

-84

Jan

-85

No

v-8

5

Sep

-86

Jul-

87

May

-88

Mar

-89

Jan

-90

No

v-9

0

Sep

-91

Jul-

92

May

-93

Mar

-94

Jan

-95

No

v-9

5

Sep

-96

Jul-

97

May

-98

Mar

-99

Jan

-00

No

v-0

0

mg

N2O

/(m

2d

ay)

0

0.5

1

1.5

2

2.5

Jan

-80

No

v-8

0

Sep

-81

Jul-

82

May

-83

Mar

-84

Jan

-85

No

v-8

5

Sep

-86

Jul-

87

May

-88

Mar

-89

Jan

-90

No

v-9

0

Sep

-91

Jul-

92

May

-93

Mar

-94

Jan

-95

No

v-9

5

Sep

-96

Jul-

97

May

-98

Mar

-99

Jan

-00

No

v-0

0

mg

N2O

/(m

2 d

ay)

Page 39: Quantifying the global N2O emissions from natural

28

(a)

(b)

Figure 3. Sensitivity analysis for NO3- and NH4

+: (a) Comparison of NO3 when increasing forcing data

of monthly air temperature, water vapor pressure, cloudiness, and precipitation by ±20%.; (b)

Comparison of NH4+when decreasing forcing data of monthly air temperature, water vapor pressure,

cloudiness and precipitation by 20%. Diamonds are for original model estimates with original forcing

data, squares for increasing all data by 20%, triangles for decreasing all data by 20%.

0

500

1000

1500

2000

2500

1 9 9 0 1 9 9 1 1 9 9 2 1 9 9 3 1 9 9 4 1 9 9 5 1 9 9 6 1 9 9 7 1 9 9 8 1 9 9 9 2 0 0 0

(mg

N)/

(Day

m2 )

origin -20 20

0

100

200

300

400

500

600

700

800

900

1 9 9 0 1 9 9 1 1 9 9 2 1 9 9 3 1 9 9 4 1 9 9 5 1 9 9 6 1 9 9 7 1 9 9 8 1 9 9 9 2 0 0 0

(mg

N)/

(Day

m2 )

origin 20 -20

+20%

+20

-20%

-20

Page 40: Quantifying the global N2O emissions from natural

29

3.1.4 Spatial variation of global N2O emissions

Global simulations show a contrast seasonal variation for all Northern and

Southern Hemispheres, especially for tropical and temperate areas (Figure 5a-c, Figure

7). There is a peak for warm weather (late spring to summer) and a crest for cold

weather (winter and early spring). Since nitrification and denitrification are highly

related to the activity of soil microbes, in the growing season, with higher temperature

and moisture, the microbes also become more active with a larger amount of bacteria

community. Meanwhile, the breaking down of deposited organic material provides

reaction with more reactants. In the end of the growing season, even with a similar

climate condition, the amount of reactants becomes smaller. Plants need fewer nutrients,

which makes the whole N cycle to be less dynamical. At dormancy stage, with less

nutrients needed for plants and less active of the reactors (related microbes), the

processes of nitrification and denitrification will slow down. Nitrous oxide, as the

product, becomes less in this period and is highly related to the temperature. But the

seasonal variation does not indicate that there must be less N2O production with low

temperature. The production of N2O is determined by the amount of reactants and

microbial activities. Based on stoichiometry theory, soils with ample organic carbon

should also be rich in organic nitrogen. In this view, soils of forests and grasslands will

not be limited by reactants amount. And the environmental conditions, such as

temperature, moisture, and pH, influence the activity of microbes that conduct the

reaction. It is traditionally believed that microbes in higher temperature should be more

active. However, each kind of microbes has its preference to environment. There may

be fewer kinds adapted for low temperature than those for warmer temperature, but they

do exist (Sprent). The distribution of emissions in July indicates that there is a

Page 41: Quantifying the global N2O emissions from natural

30

(a) rpctp

(b) bpctp

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2 d

ay)

0

0.5

1

1.5

2

2.5

Jan

-80

No

v-8

0

Sep

-81

Jul-

82

May

-83

Mar

-84

Jan

-85

No

v-8

5

Sep

-86

Jul-

87

May

-88

Mar

-89

Jan

-90

No

v-9

0

Sep

-91

Jul-

92

May

-93

Mar

-94

Jan

-95

No

v-9

5

Sep

-96

Jul-

97

May

-98

Mar

-99

Jan

-00

No

v-0

0

mg

N2O

/(m

2d

ay)

Page 42: Quantifying the global N2O emissions from natural

31

(c) Kntf

(d) k

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

Page 43: Quantifying the global N2O emissions from natural

32

(e) pn2

Figure 4. Sensitivity of key parameters: comparisons of N2O flux [mg N/(m2 day)]: (a) Comparisons

of N2O with original parameter set (blue line), and +5% rpctp (orange line); (b) Comparisons of N2O with

original parameter set (blue line), and +20% bpctp (orange line); (c) Comparison of N2O with original

parameter set (blue line), and +20% Kntf (orange line); (d) Comparison of N2O with original parameter

set (blue line), and +20% k (orange line); (e) Comparison of N2O with original parameter set (blue line),

and +20% pn2 (orange line).

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

Page 44: Quantifying the global N2O emissions from natural

33

significant amount of N2O emissions from high latitudes, especially from boreal forest

ecosystems.

Spatially, there are definite trends between area and N2O emissions. The emission

from tropical area is far more than that from high and middle latitudes, especially in

January, winter in North Hemisphere. Figure 5(a) shows in this month, most N2O is

produced in South Hemisphere, among which Amazon basin and central Africa are two

hotspots. High temperature and ample precipitation contribute to the high emission of

N. In July, emission form Northern Hemisphere becomes more (Figure 5 b, c). The

percentage of N2O from tropical area gets lower. For different latitudes, tropical and

subtropical zones have high emissions, especially from Amazon basin, Indonesia, and

central Africa, contributing more than 50% of the total emission. Compared to these

regions, East Asia, Europe and coastal area of North America have smaller area and

lower fluxes. N2O emissions from temperate zone of Northern Hemisphere give less

than 10%. Region in high latitude (>45 °N) even have more than temperate does, this

is because the former has a similar or even higher flux and larger area. Southern

Hemisphere (>23.5 °S) has much smaller land area. Without considering the emission

from ocean, this part, including Australia, South Africa and the southern part of South

America, only contributes a smaller fraction of the total emissions. Comparing

Figure.11 with Figure 5 (b), both displaying the data in July of 2000, there is a similar

trend between soil moisture and N2O emissions, especially in tropical and temperate

zone. Amazon basin, Indonesia, and central Africa are hotspots, which are also with

high soil moisture. For temperate area, the emission should be consistent with soil

moisture if we consider the farm land part, especially for East Asia. For North America,

there is more emission in east than in west due to the differences in precipitation and

Page 45: Quantifying the global N2O emissions from natural

34

soil moisture. In polar regions, although the soil moisture is lower compared to

temperate and tropical area, the high content of organic carbon and nitrogen in soil may

be an essential reason for the exceptionally high nitrous oxide emission.

With the consideration of vegetation types, forests contribute the most N2O. Boreal

forests, along with temperate forests, tropical forests and tundra, contribute more than

90% of N2O emissions in natural environment. It indicates that wet tundra, boreal forest,

and tropical forests have high fluxes compared to other types. Fluxes from vegetation

under dry conditions, such as xeric shrublands and grassland are much lower,

suggesting there are significant effects of precipitation on N2O production and

emissions. There are two factors influencing the total emission, including area and flux

density. Forests have proper humidity and temperature for microbes. Grasslands have

similar area with temperate forests, but the fluxes of N2O for forests are much higher.

The area for shrublands and woodlands are smaller and they have low productivity in

comparison with forests. Boreal forest and tundra provide more than 50% of the total

emission; this is because a large nitrogen pool in northern high latitudes is essential for

N2O emissions.

3.1.5 Seasonal and inter-annual variations of N2O emissions

Three grids are selected randomly to show the tendency in the past century, located

in the Arctic, temperate zone and tropical zone, respectively. All the three sites show

obvious seasonal variations as well. There is a decreasing trend of N2O emissions from

these grid cells. Figure 6(a) shows the 100-year simulation for the grid ( -101.5, 24.0),

which is located in tropical area of North America. The peak varies from 100 micro

grams per month to less than 20 per month, while the crest varies from 60 to 10. Grid

Page 46: Quantifying the global N2O emissions from natural

35

in Figure 6(b) is located in Nebraska, North America. This simulation shows a decrease

trend of peak value. The smallest value is near 0 in the

(a)

(b)

Page 47: Quantifying the global N2O emissions from natural

36

(c)

(d)

Figure 5. Global N2O emissions [mg N/(day m2)]: (a) January, 2000; (b) July, 2000; (c) July, 1925; (d)

July, 2000, estimated without carbon limitation

Page 48: Quantifying the global N2O emissions from natural

37

1900s, the change from 1900 to 2000 is not obvious. Figure 6(c) shows the result in

North Pole area. Different from the other two, there is decreasing no trend of N2O

emissions. A peak appears from the 1950s to 1970s and from the 1980s to 2000, there

is a decrease trend for the peak. As I mentioned, the Arctic has a large pool of organic

carbon and nitrogen. When air temperature approaches to a certain threshold, frozen

soils begin to thaw, causing the transformation of organic material. N2O, as a product

of a series of reactions, will be released.

At the global scale, there is a decreasing trend of N2O emissions, which is contrary

to my expectation (Figure 7). It decreases from year 1900, 1.92 Tg N2O month-1 in

summer and 1.46 Tg N2O month-1 in winter, to 1.3 Tg N2O month-1 in summer, 1.07 Tg

N2O month-1 in winter of year 2000. The global N2O emissions slightly decreased from

about 12.9 Tg N in 1900 to 9.1 Tg in year 2000. While the total global N2O emission

has increased since 1990 (Reay, 2012), my simulation shows that the total emission

from natural environment is decreased in the past century. There are two possible

reasons for it. First, all lands are considered to be natural, no fertilization or irrigation.

Previous studies indicate that the biogenic sources of N2O are largely from forest and

cultivated lands (Table 3). Another essential source is the biomass burning. Figure 9

indicates that in the past decades, the increase is principally due to fertilizers used in

agriculture. The second reason is due to global climate change. As Figure 5 shows, the

majority of emissions come from tropical forests and boreal forests. Global warming

leads to a decrease of precipitation in tropical and temperate area, and an increase of

that in cold areas (Figure 10). As a result, the total global precipitation keeps decreasing

since the 1970s (Figure l2).

Page 49: Quantifying the global N2O emissions from natural

38

(a)

(b)

-10

0

10

20

30

40

50

60

70

80

90

100

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

mg

N/(

day

m2 )

0

20

40

60

80

100

120

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

mg

N/(

day

m2 )

Page 50: Quantifying the global N2O emissions from natural

39

(c)

Figure 6. N2O emissions during 1900-2000 [mg N/(m2 day)]: (a) Grid in temperate area example (-97.5°,

41.0°); (b) Grid in tropical area example (-101.5°,24.0°);

-2

0

2

4

6

8

10

12

14

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

mg

N/(

day

m2 )

Page 51: Quantifying the global N2O emissions from natural

40

Figure 7. Estimated global N2O emissions from natural ecosystem soils for 1900-2000 (TgN/month).

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

19

00

19

03

19

07

19

10

19

14

19

17

19

21

19

24

19

28

19

31

19

35

19

38

19

42

19

45

19

49

19

52

19

56

19

59

19

63

19

66

19

70

19

73

19

77

19

80

19

84

19

87

19

91

19

94

19

98

TgN

/mo

nth

Page 52: Quantifying the global N2O emissions from natural

41

Consequently, emissions from tropical and temperate areas become less due to

droughts, but the nitrogen stored in frozen soil is released under warmer conditions. As

the total area in tropical and temperate zone is larger than that in the Arctic, the total

emission becomes less. Figure 12(a) shows a step in the 1940s. With the use of fertilizer

and biofuel, more ammonium releases into the atmosphere, leading to an increase of

NH4 deposition. However, the overall trend of N2O emissions is barely influenced by

ammonium deposition because the deposition amount is small compared to the soil N

pool (Figure 12(b)). It is possible that the sudden increase was led by the consideration

of other sources. But this also indicated that this part of NH4 is so essential fertilizers

used in agriculture. The second reason is due to global climate change. As Figure 5

shows, the majority of emissions come from tropical forests and boreal forests. Global

warming leads to a decrease of precipitation in tropical and temperate area, and an

increase of that in original cold areas (Figure 10). As a result, the total global

precipitation keeps decreasing since the 1970s (Figure l2). Consequently, emissions

from tropical and temperate areas become less due to droughts, but the nitrogen stored

in frozen soil is released under warmer conditions. As the total area in tropical and

temperate zone is larger than that in the Arctic, the total emission becomes less. Figure

12(a) shows a step in the 1940s. With the use of fertilizer and biofuel, more ammonium

releases into the atmosphere, leading to an increase of NH4 deposition. However, the

overall trend of N2O emissions is barely influenced by ammonium deposition because

the deposition amount is small compared to the soil N pool (Figure 12(b)). It is possible

that the sudden increase was led by the consideration of other sources. But this also

indicated that this part of NH4 is so essential that it cannot be neglected any longer.

In my analysis, the effect of organic carbon is only considered in denitrification

Page 53: Quantifying the global N2O emissions from natural

42

process (Figure 8). It is clear that when the limitation of soil carbon is not considered,

the emission of nitrogen increases significantly. Soil carbon is an energy source for N2O

emissions for microbes. The shortage of carbon will lead to uncompleted chemical

reactions related to soil nitrogen.

Figure 8. Comparison of monthly N2O emissions with (blue line) and without (orange line) considering

the effects of soil carbon limitation.

0

0.5

1

1.5

2

2.5

Jan

-80

De

c-8

0

No

v-8

1

Oct

-82

Sep

-83

Au

g-8

4

Jul-

85

Jun

-86

May

-87

Ap

r-8

8

Mar

-89

Feb

-90

Jan

-91

De

c-9

1

No

v-9

2

Oct

-93

Sep

-94

Au

g-9

5

Jul-

96

Jun

-97

May

-98

Ap

r-9

9

Mar

-00

mg

N2O

/(m

2d

ay)

Page 54: Quantifying the global N2O emissions from natural

43

Figure 9. Nitrous oxide emissions in IPCC report (Tg N2O, Source EDGAR 4.0)

Page 55: Quantifying the global N2O emissions from natural

44

Table 3. Present-day global budget of N2O emissions

Best estimate6

(Tg N /year)

Range7

(Tg N /year)

Model simulation8

(Tg N /year)

SOURCES, natural 9 6~12 9.05

Tropical soils 4 3~6 5.7

Temperate soils 2 0.6~4 1.7

SOURCES,anthropogenic 6 4~8

Cultivated soils 4 2~5

Biomass burning 0.5 0.2~0.7

Chemical industry 1.3 0.7~1.8

Livestock 0.4 0.2~0.5

SINK Stratosphere 12 9~16

6 2 Data source: http://acmg.seas.harvard.edu/people/faculty/djj/book/bookchap10.html

8 Results for 2000 global land

Page 56: Quantifying the global N2O emissions from natural

45

Figure 10. Average oil moisture data for July of 2000 (mm/month): (data source: NOAA NCEP CPC

Global Monthly Soil Moisture)

Page 57: Quantifying the global N2O emissions from natural

46

Figure. 11 Trends for global precipitation (land only) (mm/year) (Data source: climate data information,

CRU 5° gridded data set for observed precipitation over land area)

An

nu

al p

reci

pit

atio

n (

mm

)

Page 58: Quantifying the global N2O emissions from natural

47

(a)

(b)

Figure 12. (a) NH4 deposition from 1900 to 2000 (Zhuang et al., 2013); (b) N2O emissions due to NH4

deposition

0

5

10

15

20

25

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

NH

4T

g N

/ yea

r

0.005

0.0055

0.006

0.0065

0.007

0.0075

0.008

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

N2O

Tg N

/yea

r

Page 59: Quantifying the global N2O emissions from natural

48

Based on the report from United States Environment Protection Agency, there was

a clear increase of N2O emission in the United Stated in the past decade (Anderson, et

al, 2010). However, when we consider the major source of the emissions, the increase

was largely driven by agricultural activities. Fertilizers containing high nitrogen

provide sufficient reactants for nitrification and denitrification in soils. The trend for

total emissions from natural ecosystems is driven by different main factors. For the

global total emission, the increase is led by agriculture activities, transportation and

other industrial activities. Emissions from natural ecosystems are mainly driven by

temperature and moisture and microbial activities as well as N deposition (Butterbach-

Bahl et al., 2013).

3.2 Discussion

3.2.1 Environmental factors to N2O emission

Among all factors determining the amount of nitrous oxide emission, three of them

have an obvious trend with time. They are temperature, precipitation, and NH4

deposition. The anomaly displayed in Figure 13 shows an increasing temperature and a

relatively stable precipitation. From 1900 to 2000, temperature keeps going up, but

precipitation, especially from 1970 has a negative anomaly, which indicates a warmer

but drier environment. The covariance also indicates a positive relationship between

emissions and precipitation, and a negative relationship between temperature and

emissions. The small value of covariance coefficients is led by a large data set.

Nitrous oxide emissions are related to many factors, temperature and precipitation may

not be the only essential factors. Considering different seasons, the anomaly for

Page 60: Quantifying the global N2O emissions from natural

49

(a)

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

3

3.5

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

3

3.5

190

0

190

4

190

8

191

2

191

6

192

0

192

4

192

8

193

2

193

6

194

0

194

4

194

8

195

2

195

6

196

0

196

4

196

8

197

2

197

6

198

0

198

4

198

8

199

2

199

6

200

0 pre

cip

ita

tio

n (

mm

/mo

nth

)

tem

per

atu

re (

°C)

0

0.2

0.4

0.6

0.8

1

1.2

190

0

190

3

190

6

190

9

191

2

191

5

191

8

192

1

192

4

192

7

193

0

193

3

193

6

193

9

194

2

194

5

194

8

195

1

195

4

195

7

196

0

196

3

196

6

196

9

197

2

197

5

197

8

198

1

198

4

198

7

199

0

199

3

199

6

199

9

N2O

(T

gN

/yea

r)

Page 61: Quantifying the global N2O emissions from natural

50

(b)

-10

-5

0

5

10

15

-1

-0.5

0

0.5

1

1.5

190

0

190

4

190

8

191

2

191

6

192

0

192

4

192

8

193

2

193

6

194

0

194

4

194

8

195

2

195

6

196

0

196

4

196

8

197

2

197

6

198

0

198

4

198

8

199

2

199

6

200

0

pre

cip

ita

tio

n (

mm

/mo

nth

)

tem

per

atu

re(°

C)

0

0.2

0.4

0.6

0.8

1

1.2

190

01

90

31

90

61

90

91

91

21

91

51

91

81

92

11

92

41

92

71

93

01

93

31

93

61

93

91

94

21

94

51

94

81

95

11

95

41

95

71

96

01

96

31

96

61

96

91

97

21

97

51

97

81

98

11

98

41

98

71

99

01

99

31

99

61

99

9

N2O

(T

gN

/yea

r)

Page 62: Quantifying the global N2O emissions from natural

51

(c)

-80

-60

-40

-20

0

20

40

60

80

-3.5

-2.5

-1.5

-0.5

0.5

1.5

2.5

3.5

190

0

190

4

190

8

191

2

191

6

192

0

192

4

192

8

193

2

193

6

194

0

194

4

194

8

195

2

195

6

196

0

196

4

196

8

197

2

197

6

198

0

198

4

198

8

199

2

199

6

200

0

pre

cip

ita

tio

n(m

m/m

on

th)

tem

per

atu

re (

°C)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

190

0

190

3

190

6

190

9

191

2

191

5

191

8

192

1

192

4

192

7

193

0

193

3

193

6

193

9

194

2

194

5

194

8

195

1

195

4

195

7

196

0

196

3

196

6

196

9

197

2

197

5

197

8

198

1

198

4

198

7

199

0

199

3

199

6

199

9

N2O

(T

gN

/yea

r)

Page 63: Quantifying the global N2O emissions from natural

52

(d)

-50

-40

-30

-20

-10

0

10

20

30

40

50

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

19

00

19

04

19

08

19

12

19

16

19

20

19

24

19

28

19

32

19

36

19

40

19

44

19

48

19

52

19

56

19

60

19

64

19

68

19

72

19

76

19

80

19

84

19

88

19

92

19

96

20

00

pre

cip

ita

tio

n (

mm

/mo

nth

)

tem

per

atu

re (

°C)

0

0.2

0.4

0.6

0.8

1

1.2

19

00

19

04

19

08

19

12

19

16

19

20

19

24

19

28

19

32

19

36

19

40

19

44

19

48

19

52

19

56

19

60

19

64

19

68

19

72

19

76

19

80

19

84

19

88

19

92

19

96

20

00

N2O

(T

gN

/yea

r)

Page 64: Quantifying the global N2O emissions from natural

53

(e)

Figure 13 Upper panel: anomalies of precipitation and temperature; Lower panel: Trend for N2O emission:

(a) Winter: Covariance for T=-0.019, correlation for T=-0.37. Covariance for P=-0.00566, correlation for

P=-0.08847; (b) Spring: Covariance for T=- -0.026, correlation for T= -0.72, Covariance for P=- 0.082,

correlation for P=0.22; (c) Summer: Covariance for T= -0.068, correlation for T= -0.69, Covariance for

P= 1.17, correlation for P= 0.46; (d) Autumn: Covariance for T= -0.016, correlation for T= -0.56,

Covariance for P= 0.056, correlation for P= 0.045; (e) Whole year: Covariance for T= -0.016, correlation

for T= -0.56, Covariance for P= 0.056, correlation for P= 0.04

-5.00

-4.00

-3.00

-2.00

-1.00

0.00

1.00

2.00

3.00

4.00

5.00

-1.5

-1

-0.5

0

0.5

1

1.5

19

00

19

05

19

10

19

15

19

20

19

25

19

30

19

35

19

40

19

45

19

50

19

55

19

60

19

65

19

70

19

75

19

80

19

85

19

90

19

95

20

00

pre

cip

itat

ion

(cm

/ye

ar)

tem

pe

ratu

re (

°C)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

19

00

19

04

19

08

19

12

19

16

19

20

19

24

19

28

19

32

19

36

19

40

19

44

19

48

19

52

19

56

19

60

19

64

19

68

19

72

19

76

19

80

19

84

19

88

19

92

19

96

20

00

N2

O T

gN/m

on

th

Page 65: Quantifying the global N2O emissions from natural

54

precipitation in winter (December, January and February) (Figure 13(a)) for Northern

Hemisphere is much smaller than that in summer (June, July and August) (Figure 13(c)).

There are also more declines of emissions in summer months than in winter, partly as a

result of drought in temperate area.

When considering the emissions from different latitudes, tropical zone (23.5° S-

23.5° N) contributes to more than half of the emission (Figure 14), which is comparable

with previous study (Prinn et al. 1990, Bouwman et al. 1995, Hisch et al. 2006).

Although these studies considered all emission sources including arable land or even

industrial production, the percentage of their area is similar. Tropical zone has optimum

physical conditions for reaction with the most substrate, so it is natural to have the most

emission. Northern Hemisphere has broader land area compared to Southern

Hemisphere. Large amounts of fertilizers are applied in these areas each year. As a result

of that, the percentage of emissions led by human activities has increased year after

year, compared to the natural soil emissions (Saikawa et al., 2014). However, even

without these human activities, the storage of carbon and nitrogen in tundra and forests

are substantial, especially, when we consider their total area (Figure 17). From the first

half of the 20th century to the last decade, the emission percentage of tropical area get

larger but for temperate area, smaller. It can be led by the drought in temperate regions,

and the depletion of nitrogen by arable land. The percentage of high latitudes stays the

same, but contributing even more than temperate area. Given the large area of high

latitudes, any small change of emission flux density can be significant to the total

emissions.

Page 66: Quantifying the global N2O emissions from natural

55

In addition, the deposition of ammonia is also an important factor. Compared to the

nitrogen amount in soils, the deposition seems to be too trivial. Figure 13(a) shows an

obvious step in the 1940s. The step may be caused by different measure method, but it

shows an increase trend in general. Saikawa et al (2014) included the in-situ and aircraft

measurements, confirming an increasing fraction of N2O in the atmosphere, with the

rate of 0.1-0.7% per year. The use of nitrogen based fertilizer and tillage also led to a

higher content in the atmosphere. To some extent, N deposition reduces the depletion

of nitrogen in soils but contributes a little because the N pool in soil is quite large.

What’s more, the distribution of deposition is uneven. Heavily populated area gets more

deposition, because of industrial and agricultural activities.

3.2.2 Comparison with previous studies

My estimation of N2O emissions only contains the part form natural soils.

Traditionally, many studies focused on farm land as it is the major source for the

increasing emission, my study investigates the natural environment emission strengths

under various physical conditions. There are a few studies concerning the emission

from ocean, and fewer concerning natural soils. My research will provide another

estimate of this potent greenhouse gas.

Page 67: Quantifying the global N2O emissions from natural

56

(a)

(b)

18%

11%

63%

8%

1

2

3

4

18%

12%

61%

9%

1

2

3

4

Page 68: Quantifying the global N2O emissions from natural

57

(c)

Figure 14. N2O emissions for different latitudes (% of the total emissions) 1: 45°N-90°N, 2:23.5°N-

45°N, 3:23.5°S-23.5°N, 4:23.5°S-90°S. (a) 1900-2000; (b) 1900-1950; (c) 1980-2000

18%

8%

67%

7%

1

2

3

4

Page 69: Quantifying the global N2O emissions from natural

58

(a)

(b) (c)

(d) (e)

Figure 15. Comparison for of N2O emission in previous studies (%). 1: 30°S-90°S, 2:30°S-30°N,

3:30°N-90°N. (a) This study, for 1980-2000;(b)Prinn, R. et al (1990), for 1978-1988; (c)Bouwman et al

(1995), for 1990; (d)Hisch et al (2006), for 1998-2001; (e)Huang et al (2008), for 1997-2001

17%

267%

326%

112%

260%

328%

114%

260%

326%

12%

281%

317%

14%

277%

319%

Page 70: Quantifying the global N2O emissions from natural

59

Werner et al (2007) reckons that, although the percentage of N2O emissions driven

by human activities keeps increasing, especially in agriculture area, nitrous oxide from

natural soils and ocean still contribute more than 60% of the global emission. More

than one-third of the emissions are produced by fertilizer and manure management,

transportation, and biofuel combustion (Syakila et al, 2011), and less than 30% is from

oceans (Duce et al., 2008). Table 4 compared different studies for recent 20 years. The

total emission ranges from 14 to 21 TgN/year including all emissions for land and

oceans, and about 10 to 13.6 from land. Some of these results are from model

simulations, and some are based on observations. The percentage of natural land

emissions is ~40% or approximately 5.6 to 8.4 Tg N /year according to these studies.

Table 3 shows the percentage from different sources in 2000. Anthropogenic sources

(~4 TgN/year) are less than half of natural soil source (~9 TgN/year). My estimation

for 2000 is 9.05 TgN/year, which is comparable with these studies.

Table 3 shows the emission from tropical soils and temperate soils and they are

within a range of estimates from previous studies. My estimation for tropical and cold

regions is high and for temperate region is low. The key reason is that the model is

sensitive to precipitation. Ample rainfall in tropical forests and high moisture in boreal

forests led to high fluxes, while droughts in recent years in temperate region may

dampen the emissions. Figure 14 displays the results for different latitudes and Figure

15 shows that the results for 1980-2000. All these estimates fall the range of estimates

from previous studies for 1990-1990 and for 1997-2001. Tropical regions contribute

more while Polar regions emit less in comparison to the emissions from oceans in these

regions.

Page 71: Quantifying the global N2O emissions from natural

60

Table 4 Estimated N2O emissions from previous studies

Author Year

for data

Estimation9 Model or experiment

Huang, J.,

et

al.(2008)

1997-

2005

16.3 for 1997-2001, 15.4 for

2001-2005 (66% probability

error)

Model of Atmospheric Transport and

Chemistry (MATCH)

Prinn, R.,

et al (1990)

07/197

8-

06/198

8

20.5+/-2.4 total land and

ocean

Atmospheric Lifetime Experiment (ALE)

and Global Atmospheric Gases Experiment

(GAGE)

Wells,

K.C., et al

(2015)

04/201

0-

04/201

2

14.2 for total and and ocean,

3.2 for natural soil

GEOS-chem

Saikawa et

al (2014)

1995-

2008

18.1+/- 0.6 for 2002-2005,

16.8+/-0.64 for 1997-2001

AGAGE, NOAA CCGG, NOAA OTTO,

RITS, and CATS, CSIRO, NIES, and

Tohoku University

Hirch et al

(2006)

1998-

2001

9.7-13.6 for total land NOAA/ESRL GMD CCGG (the Carbon

Cycle Greenhouse Gases group in the

Global Monitoring Division at the NOAA

Earth System Research Laboratory in

Boulder, Colorado)

Thompson

et al (2014)

1999-

2009

17.5 to 20.1 Observation with time-dependent Bayesian

inversion technique

Thompson

et al (2014)

2006-

2009

16.1-19.7 Chemistry transport models (CTMs)

Zhuang et

al (2012)

2000 1.96-4.56 for natural soil Artificial neural network and field

observational data

3.2.3 Model limitation

First, the major uncertainty source is due to the structure of model. This model does

not include every section of nitrogen cycle. Processes in soil (mineralization,

nitrification, and denitrification) and the interaction between soil and plants were fully

considered. However, the N pool in the atmosphere is also essential. For the inflow

from atmosphere to the soil, I only considered the deposit of ammonia. Actually, the

deposit of nitrate is also an important source. This part will be complimented when I

get proper data source. N fixation is the only biological process making N2 in the

atmosphere accessible to plants. This process is involved in my model in the current

9 TgN per year

Page 72: Quantifying the global N2O emissions from natural

61

stage, but the simulation shown in this thesis does is without N fixation. This could be

a reason for the depletion of N in the soil.

Another source is from the quality and quantity of observational data, which not

only limits the quality of parameterization, but also makes it difficult to verify the model.

The first problem is the short period of observation. The observation at a site usually

lasts for several days to months only. However, in my study, I can only use one value

to represent a whole month’s fluxes in my parameterization. Continuous observations

for a long period (e.g., several years) at a site will be valuable for my future study. The

second problem is the contradiction of accuracy and observation period. Specifically,

instruments can measure fluxes at several seconds time interval, but this kind of

observation only lasts for days at most. For the long- time measurement, observational

data can have an uncertainty as large as several times of the observation. The third

problem is, that accurately measuring N2O fluxes is still difficult although many

methods have been applied. Chamber technique is a traditional method and many

observations were conducted in the 1980s with this method. However the measurement

could be significantly perturbed by chamber itself and the uncertainty can be 55%-145%

(Christensen et al, 1996). The uncertainty from eddy covariance flux measurements can

be more than 100%, which is greatly due to the small eddy flux of nitrous oxide (Kroon

et al., 2010). With the flux gradient method, the uncertainty is determined by estimating

gas diffusivity. The Bowen ratio method is good at getting air temperature and water

vapor pressure. The uncertainty mainly comes from the errors of sensors and low

available energy. In practice, the low energy exchange may lead to some missing data

of N2O fluxes. Aircraft-based measurement is better at providing spatial-averaged gas

flux measurement. The shortcoming is that the flight length is short compared to the

Page 73: Quantifying the global N2O emissions from natural

62

scale of convections. In addition, modeling should use the most recent data because,

with the development of more accurate instrument and methods, recent observations

tend to be more accurate. For some types of 11 vegetation types I modeled, the

observational data were collected in the 1980s. For others such as temperate forests or

tropical forests, the recent observational data were relatively more accurate and easy to

access. Different data sources give the inconstant uncertainty for various vegetation

types. NO3- deposition with acid rain also led to the change of substrate. There are some

difficulties in archiving the data in the past 100 years, so the increases of NO3-

deposition also need to be considered. I will consider this part in my future study. All

these will induce uncertainties in my current quantification.

Third, my simple hydrological model to link precipitation and soil moisture is not

adequate. A more sophisticated hydrological model that considers characteristics of

soils shall be used (e.g., VIC model, Lu and Zhuang, 2012, Zhu et al., 2014). In my

current simulation, for the analysis in the past 100 years, soil moisture data quality

before 1990 is much poorer than that for the recent 20 years (Figure 10). The coming

SMAP (Soil Moisture Active Passive) data products will be more accurate for the

analysis of topsoil layer moisture, which will be valuable to improve my study.

Fourth, the activity of microbes is also needed to be further considered. In current

model, soil organic carbon is simply treated as a factor (substrate), with a simple

algorithm. In fact, soil microbes carbon and nitrogen pools, should be explicitly

considered in my system model. Microbes need carbon to grow. Similarly, nitrogen is

also needed for them as an essential component of cells and enzyme. Therefore carbon

Page 74: Quantifying the global N2O emissions from natural

63

and nitrogen consumed by microbes are not negligible because the more active they are,

the more organic matter will need to be consumed, and subsequently, the more N2O

will be produced and released.

Finally, the uncertainty of the model due to uncertain parameters and input data

shall be conducted. Observational data with better quality and additional research on

the processes of nitrification and denitrification will help to reduce the prediction

uncertainty of the model. To improve the performance of model, first, better

observational data is needed to further constrain parameters. Second, using a better

statistical method for parameterization may also help. The data quality for

precipitation, temperature and soil moisture will be improved with better observation,

which will help constrain the simulation uncertainty.

Page 75: Quantifying the global N2O emissions from natural

64

CHAPTER 4 CONCLUSION AND FUTURE DIRECTIONS

A process-based model of N2O emissions was developed based on TEM. The

model was used to simulate the emissions of nitrous oxide at site-level and the global

scale. Eleven sites were selected to calibrate 5 key model parameters. Each vegetation

type has its own set of parameters. The calibrated model captured the trends of

observation at those observation sites. The results suggest that more observational data

with higher accuracy are needed for better simulations. For regional and global

simulations, the results show there are evident seasonal variations for N2O emissions.

There are regional differences, but no obvious relations between latitudes and emissions.

Tropical region have high emissions in the whole year. In the summer of the Northern

Hemisphere, emissions from polar region are quite significant. Considering the

limitation of soil organic carbon in the model, there is less N2O emissions compared to

those without carbon limitation (Figure 5b, d). From 1900 to 2000, there is a slightly

decreasing trend for the global emission, which is a result of many physical factors,

including temperature and precipitation.

My future research will focus on three aspects. First I will collect more

observational data to reduce the parameter uncertainty (Chatskikh et al.1996, Hirsh et

al 2006, Kroon et al. 2010, Yuesi et al. 2003). Meanwhile, a better set of parameters

will be obtained. Second, I will incorporate microbial traits and their effects on N2O

emissions, which needs developing a trait-based model (Cleveland et al. 2007, Li et al.

Page 76: Quantifying the global N2O emissions from natural

65

2005, Qin et al. 2013). Trait-based model links the microbe’s traits, such as metabolism,

propagation to the response of gas emissions (Buckeridge et al. 2009, Young et al. 2004).

Third, I will improve the coupling between C and N following the principles of their

stoichiometry in TEM modeling framework. Finally, I will analyze N2O simulations

from natural ecosystems during the 21st century under future climate conditions.

Page 77: Quantifying the global N2O emissions from natural

REFERENCES

Page 78: Quantifying the global N2O emissions from natural

66

REFERENCES

Anderson, B., et al. (2010) "Methane and nitrous oxide emissions from natural sources."

United States Environmental Protection Agency, Office of Atmospheric Programs,

Washington.

Allison, S., (2012). "A trait‐based approach for modelling microbial litter

decomposition." Ecology letters 15(9): 1058-1070.

Anderson, I. C. and M. A. Poth (1989). "Semiannual losses of nitrogen as NO and N2O

from unburned and burned chaparral." Global Biogeochemical Cycles 3(2): 121-135.

Ayres, R. U. (1996). Industrial metabolism and the grand nutrient cycles, INSEAD.

Bell, M. J., et al. (2011). "Simulation of soil nitrogen, nitrous oxide emissions and

mitigation scenarios at 3 European cropland sites using the ECOSSE model." Nutrient

Cycling in Agroecosystems 92(2): 161-181.

Bradbury, N., et al. (1993). "Modelling the fate of nitrogen in crop and soil in the years

following application of 15N-labelled fertilizer to winter wheat." The Journal of

Agricultural Science 121(03): 363-379.

Bremner, J. M., and K. Shaw (1958). "Denitrification in soil. II. Factors affecting

denitrification." The Journal of Agricultural Science 51.01: 40-52.

Page 79: Quantifying the global N2O emissions from natural

67

Brown, A. T., et al. (2013) "Stratospheric lifetimes of CFC-12, CCl4, CH4, CH3Cl and

N2O from measurements made by the Atmospheric Chemistry Experiment-Fourier

Transform Spectrometer (ACE-FTS)." Atmos. Chem. Phys 13.14: 6921-6950.

Buckeridge, K. M., et al. (2010). "Soil biogeochemistry during the early spring in low

arctic mesic tundra and the impacts of deepened snow and enhanced nitrogen

availability." Biogeochemistry 99(1-3): 127-141.

Butterbach-Bahl, K., et al. (2013) "Nitrous oxide emissions from soils: how well do we

understand the processes and their controls?." Philosophical Transactions of the Royal

Society of London B: Biological Sciences 368.1621: 20130122.

Chatskikh, D., et al. (2005). "Simulation of effects of soils, climate and management

on N2O emission from grasslands." Biogeochemistry 76(3): 395-419.

Christensen, S., et al (1996). "Nitrous oxide emission from an agricultural field:

comparison between measurements by flux chamber and micrometerological

techniques." Atmospheric Environment 30.24: 4183-4190.

Cleveland, C. C. and D. Liptzin (2007). "C: N: P stoichiometry in soil: is there a

“Redfield ratio” for the microbial biomass?" Biogeochemistry 85(3): 235-252.

Danevčič, T., et al. (2010). "Emissions of CO2, CH4 and N2O from Southern European

peatlands." Soil Biology and Biochemistry 42(9): 1437-1446.

Davidson, Eric A., et al. (2000) "Testing a Conceptual Model of Soil Emissions of

Nitrous and Nitric Oxides Using two functions based on soil nitrogen availability and

soil water content, the hole-in-the-pipe model characterizes a large fraction of the

observed variation of nitric oxide and nitrous oxide emissions from

soils." Bioscience 50.8: 667-680.

Page 80: Quantifying the global N2O emissions from natural

68

Delgrosso, S., et al. (2005). "DAYCENT model analysis of past and contemporary soil

NO and net greenhouse gas flux for major crops in the USA." Soil and Tillage Research

83(1): 9-24.

Dentener, F., et al. (2006). "Nitrogen and sulfur deposition on regional and global scales:

a multimodel evaluation." Global Biogeochemical Cycles 20(4).

Duce, R. A., et al. (2008) "Impacts of atmospheric anthropogenic nitrogen on the open

ocean." science 320.5878: 893-897.

Engel, T. and E. Priesack (1993). Expert-N - A Building Block System of Nitrogen

Models as Resource for Advice, Research, Water Management and Policy. Integrated

Soil and Sediment Research: A Basis for Proper Protection. H. P. Eijsackers and T.

Hamers, Springer Netherlands. 1: 503-507.

Eswaran, H., et al. (1993). "Organic carbon in soils of the world." Soil Science Society

of America Journal 57(1): 192-194.

Gorham, E. (1991). "Northern peatlands: role in the carbon cycle and probable

responses to climatic warming." Ecological applications 1(2): 182-195.

Griffiths, Bryan S., Annette Spilles, and Michael Bonkowski (2012). "C: N: P

stoichiometry and nutrient limitation of the soil microbial biomass in a grazed grassland

site under experimental P limitation or excess." Ecological Processes 1.1: 1-11.

He, Y., et al (2014). "Uncertainty in the fate of soil organic carbon: A comparison of

three conceptually different decomposition models at a larch plantation." Journal of

Geophysical Research: Biogeosciences 119.9: 1892-1905.

Hénault, C. and J. Germon (2000). "NEMIS, a predictive model of denitrification on

the field scale." European Journal of Soil Science 51(2): 257-270.

Page 81: Quantifying the global N2O emissions from natural

69

Hirsch, A. I., et al (2006). "Inverse modeling estimates of the global nitrous oxide

surface flux from 1998–2001." Global Biogeochemical Cycles 20.1.

Jones, D. L., and A. Hodge (1999). "Biodegradation kinetics and sorption reactions of

three differently charged amino acids in soil and their effects on plant organic nitrogen

availability." Soil Biology and Biochemistry 31.9: 1331-1342.

Kettunen, A. (2003). "Connecting methane fluxes to vegetation cover and water table

fluctuations at microsite level: a modeling study." Global Biogeochemical Cycles 17(2).

Kimble, J. M., et al. (1998). The potential of US cropland to sequester carbon and

mitigate the greenhouse effect, CRC Press.

Kirkby, C. A., et al. (2011) "Stable soil organic matter: a comparison of C: N: P: S ratios

in Australian and other world soils." Geoderma 163.3: 197-208.

Kroeze, C., et al. (1999). "Closing the global N2O budget: a retrospective analysis

1500–1994." Global Biogeochemical Cycles 13(1): 1-8.

Kroon, P. S., et al (2010). "Uncertainties in eddy covariance flux measurements

assessed from CH 4 and N 2 O observations." Agricultural and forest meteorology

150.6: 806-816.

Lal, R. (2004). "Soil carbon sequestration impacts on global climate change and food

security." Science 304(5677): 1623-1627.

Li, C. (1992). "A Model of Nitrous Oxide Evolution from Soil Driven by Rainfall

Events: 1. Model Structure and Sensitivity." Journal of Geophysical Research 97(D9):

9756-9776.

Li, C., et al. (2005). "Carbon sequestration in arable soils is likely to increase nitrous

oxide emissions, offsetting reductions in climate radiative forcing." Climatic Change

72(3): 321-338.

Page 82: Quantifying the global N2O emissions from natural

70

Linn, D. and J. Doran (1984). "Effect of water-filled pore space on carbon dioxide and

nitrous oxide production in tilled and nontilled soils." Soil Science Society of America

Journal 48(6): 1267-1272.

Liu, J., et al. (2011). "Responses of N2O Flux from Forest Soils to Land Use Change

in Subtropical China." The Botanical Review 77(3): 320-325.

Lu, X. and Q. Zhuang (2012). "Carbon and water dynamics in the Alaska Yukon River

Basin: Analysis using process-based biogeochemistry models and satellite data.".

Machida, T., et al (1995). "Increase in the atmospheric nitrous oxide concentration

during the last 250 years." Geophysical Research Letters 22.21: 2921-2924.

Maljanen, M., et al. (2006). "Methane and nitrous oxide fluxes, and carbon dioxide

production in boreal forest soil fertilized with wood ash and nitrogen." Soil Use and

Management 22(2): 151-157.

Mammarella, I., et al. (2010). "A case study of eddy covariance flux of N 2 O measured

within forest ecosystems: quality control and flux error analysis." Biogeosciences 7(2):

427-440.

Martikainen, P. J., et al. (1993). "Effect of a lowered water table on nitrous oxide fluxes

from northern peatlands." Nature 366(6450): 51-53.

Matson, P. A., et al. (1991). "Annual nitrous oxide flux and soil nitrogen characteristics

in sagebrush steppe ecosystems." Biogeochemistry 14(1): 1-12.

Mosier, A. R., et al(1993). "Methane and nitrous oxide flux in a Wyoming subalpine

meadow." Global Biogeochemical Cycles 7.4: 771-784.

Mosier, A. R., et al(1997). "Impact of agriculture on soil consumption of atmospheric

CH4 and a comparison of CH4 and N2O flux in subarctic, temperate and tropical

grasslands." Nutrient Cycling in Agroecosystems49.1-3: 71-83.

Page 83: Quantifying the global N2O emissions from natural

71

Nadelhoffer, K. J., et al. (1999). "Nitrogen deposition makes a minor contribution to

carbon sequestration in temperate forests." Nature 398(6723): 145-148.

Nevison, C. and E. Holland (1997). "A reexamination of the impact of

anthropogenically fixed nitrogen on atmospheric N2O and the stratospheric O3 layer."

Journal of Geophysical Research: Atmospheres (1984–2012) 102(D21): 25519-25536.

Oenema, Oene, et al. "Trends in global nitrous oxide emissions from animal production

systems." Nutrient Cycling in Agroecosystems 72.1 (2005): 51-65.

Parton, W., et al. (1996). "Generalized model for N2 and N2O production from

nitrification and denitrification." Global Biogeochemical Cycles 10(3): 401-412.

Portmann, R. W., J. S. Daniel, and A. R. Ravishankara. "Stratospheric ozone depletion

due to nitrous oxide: influences of other gases."Philosophical Transactions of the Royal

Society of London B: Biological Sciences 367.1593 (2012): 1256-1264.

Potter, C. S., et al. (1996). "Process modeling of controls on nitrogen trace gas

emissions from soils worldwide." Journal of Geophysical Research: Atmospheres

(1984–2012) 101(D1): 1361-1377.

Prinn, R., et al. "Atmospheric emissions and trends of nitrous oxide deduced from 10

years of ALE–GAGE data." Journal of Geophysical Research: Atmospheres (1984–

2012) 95.D11 (1990): 18369-18385.

Qin, Z., et al. (2013). "Carbon and nitrogen dynamics in bioenergy ecosystems: 1.

Model development, validation and sensitivity analysis." GCB Bioenergy.

Qin, Z., et al. (2013). "Carbon and nitrogen dynamics in bioenergy ecosystems: 2.

Potential greenhouse gas emissions and global warming intensity in the conterminous

United States." GCB Bioenergy.

Page 84: Quantifying the global N2O emissions from natural

72

Rösch, C., et al. (2002). "Biodiversity of denitrifying and dinitrogen-fixing bacteria in

an acid forest soil." Applied and Environmental Microbiology 68(8): 3818-3829.

Reay, D. S., et al. (2012). "Global agriculture and nitrous oxide emissions." Nature

Climate Change 2(6): 410-416.

Repo, M. E., et al. (2009). "Large N2O emissions from cryoturbated peat soil in tundra."

Nature Geoscience 2(3): 189-192.

Robertson, L. A. and J. G. Kuenen (1990). "Combined heterotrophic nitrification and

aerobic denitrification in Thiosphaera pantotropha and other bacteria." Antonie van

Leeuwenhoek 57(3): 139-152.

Rodhe, H. (1990). "A comparison of the contribution of various gases to the greenhouse

effect." Science 248(4960): 1217-1219.

Saikawa, E.,et al.(2014)"Global and regional emissions estimates for

N2O."Atmospheric Chemistry and Physics 14.9: 4617-4641.

Saikawa, E., et al. (2013). "Global modeling of soil nitrous oxide emissions from

natural processes." Global Biogeochemical Cycles 27(3): 972-989.

Sawyer, C. N., Wild, H. E., & McMahon, T. C. (1974). "Nitrification and denitrification

facilities wastewater treatment." EPA 625. Vol. 4. EPA.

Smith, A., et al. (2001). "Analyzing variety by environment data using multiplicative

mixed models and adjustments for spatial field trend." Biometrics 57(4): 1138-1147.

Smith, J., et al. (2010). "Estimating changes in Scottish soil carbon stocks using

ECOSSE. I. Model description and uncertainties." Climate Research 45: 179-192.

Smith, J., et al. (2010). "Estimating changes in Scottish soil carbon stocks using

ECOSSE. II. Application." Climate Research 45: 193-205.

Page 85: Quantifying the global N2O emissions from natural

73

Smith, K., et al. (1998). "Effects of temperature, water content and nitrogen fertilisation

on emissions of nitrous oxide by soils." Atmospheric Environment 32(19): 3301-3309.

Sprent, J.I.,. "Global Distribution of Legumes." Legume Nodulation: A Global

Perspective: 35-50.

Syakila, A., and C. Kroeze (2011). "The global nitrous oxide budget

revisited."Greenhouse Gas Measurement and Management 1.1: 17-26.

Thompson, R.L., et al. (1999) "Nitrous oxide emissions 1999 to 2009 from a global

atmospheric inversion." Atmospheric Chemistry and Physics 14.4 (2014): 1801-1817.

Thompson, R.L., et al. (2014) "Nitrous oxide emissions 1999 to 2009 from a global

atmospheric inversion." Atmospheric Chemistry and Physics 14.4: 1801-1817.

Thompson, R.L., et al. (2014) "TransCom N 2 O model inter-comparison–Part 2:

Atmospheric inversion estimates of N 2 O emissions." Atmospheric Chemistry and

Physics 14.12: 6177-6194.

Warneck, P.. Chemistry of the natural atmosphere (Vol. 71). Academic press

Wells, K. C., et al. (2015) "Simulation of atmospheric N2O with GEOS-Chem and its

adjoint: evaluation of observational constraints." Simulation 8: 5367-5418.

Werner, C., et al. (2007). "A global inventory of N2O emissions from tropical rainforest

soils using a detailed biogeochemical model." Global Biogeochemical Cycles 21(3).

William, J. Parton, M. H., Ojima, D., Schimel, D. (1998). "DAYCENT and its land

surface submodel: description and testing." Global and Planetary Change(19): 35-48.

Xing, G. (1998). "N2O emission from cropland in China." Nutrient Cycling in

Agroecosystems 52(2-3): 249-254.

Yan, X., et al. (2003). "Estimation of nitrous oxide, nitric oxide and ammonia emissions

from croplands in East, Southeast and South Asia." Glob Chang Biol 9(7): 1080-1096.

Page 86: Quantifying the global N2O emissions from natural

74

Young, I. M. and J. W. Crawford (2004). "Interactions and self-organization in the soil-

microbe complex." Science 304(5677): 1634-1637.

Yuesi, W. and W. Yinghong (2003). "Quick measurement of CH4, CO2 and N2O

emissions from a short-plant ecosystem." Advances in Atmospheric Sciences 20(5):

842-844.

Zhu, X., Q. Zhuang, X. Lu and L. Song (2014), Spatial scale-dependent land–

atmospheric methane exchanges in the northern high latitudes from 1993 to 2004,

Biogeosciences, 11, 1693-1704, doi:10.5194/bg-11-1693-2014.

Zhuang, Q., et al. (2002). "Modeling soil thermal and carbon dynamics of a fire

chronosequence in interior Alaska." Journal of Geophysical Research: Atmospheres

(1984–2012) 107(D1): FFR 3-1-FFR 3-26.

Zhuang, Q., et al. (2004). "Methane fluxes between terrestrial ecosystems and the

atmosphere at northern high latitudes during the past century: A retrospective analysis

with a process‐based biogeochemistry model." Global Biogeochemical Cycles 18(3).

Zhuang, Q., et al. (2006). "CO2 and CH4 exchanges between land ecosystems and the

atmosphere in northern high latitudes over the 21st century." Geophysical Research

Letters 33(17).

Zhuang, Q., et al. (2001). "Incorporation of a permafrost model into a large‐scale

ecosystem model: Evaluation of temporal and spatial scaling issues in simulating soil

thermal dynamics." Journal of Geophysical Research: Atmospheres (1984–2012)

106(D24): 33649-33670.

Zhuang, Q., et al. (2013) "Response of global soil consumption of atmospheric methane

to changes in atmospheric climate and nitrogen deposition." Global Biogeochemical

Cycles 27.3: 650-663.