downscaling the maximum carboxylation rate ( π‘Ž derived

84
Downscaling the Maximum Carboxylation Rate ( ) Derived from Satellite Sun-induced Chlorophyll Fluorescence Data Using High-resolution Remote Sensing Products by Jiye Leng A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Geography and Planning University of Toronto Β© Copyright by Jiye Leng, 2020

Upload: others

Post on 16-Nov-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

Downscaling the Maximum Carboxylation Rate (π‘‰π‘π‘šπ‘Žπ‘₯) Derived from Satellite Sun-induced Chlorophyll

Fluorescence Data Using High-resolution Remote Sensing Products

by

Jiye Leng

A thesis submitted in conformity with the requirements for the degree of Master of Science

Department of Geography and Planning University of Toronto

Β© Copyright by Jiye Leng, 2020

Page 2: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

ii

Downscaling the Maximum Carboxylation Rate (π‘‰π‘π‘šπ‘Žπ‘₯) Derived

from Satellite Sun-induced Chlorophyll Fluorescence Data Using

High-resolution Remote Sensing Products

Jiye Leng

Master of Science

Department of Geography & Planning

University of Toronto

2020

Abstract

The maximum carboxylation rate (π‘‰π‘π‘šπ‘Žπ‘₯) influences the magnitude of gross primary productivity

(GPP). Currently, reliable global π‘‰π‘π‘šπ‘Žπ‘₯ products derived from satellite sun-induced chlorophyll

fluorescence (SIF) data are at coarse resolutions, which cannot meet the demand of global

ecological research. In this thesis, the π‘‰π‘π‘šπ‘Žπ‘₯25 (π‘‰π‘π‘šπ‘Žπ‘₯ normalized to 25Β°C) dataset derived from

satellite SIF at a coarse resolution (0.1Β°, ~11 km) is downscaled to a higher resolution (1 km)

through a downscaling scheme using photochemical reflectance index (PRI) and spatial scaling

algorithms based on leaf chlorophyll content (LCC) and normalized difference vegetation index

(NDVI). The Boreal Ecosystem Productivity Simulator (BEPS) is used to evaluate the downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 using tower flux data. The results show that the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 data appreciatively

improve GPP simulations at the tower sites, indicating LCC as a feasible way for downscaling the

π‘‰π‘π‘šπ‘Žπ‘₯25 dataset. GPP estimations at the 0.1Β° resolution decrease by 2-7% after π‘‰π‘π‘šπ‘Žπ‘₯

25 downscaling.

Page 3: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

iii

Acknowledgments

I was fortunate to be admitted into University of Toronto. Life was unusual during this year with

the unprecedented pandemic, and I’m writing to express appreciation to the people around me

throughout my master's study.

I would like to first thank my supervisor, Prof. Jing Chen, for offering me the opportunity to pursue

an M.Sc. degree and a Ph.D. degree at this university. He has always been supportive, willing to

help, and enlighten me when I encountered obstacles in my research. He is a role model for me as

a scientist and a bright lighthouse who leads me to step into the research career and encourages

me to fulfill my dream.

I also would like to thank my course instructor and TA instructor, Prof. Jane Liu, for guiding me

into the new scientific field. The three individual classes built me with a foundation of the

knowledge for exploring this new field and fostered my ability of logical and critical thinking.

I want to thank all the group members for their advice, help, and care during this year. Special

thanks to Yihong Liu, Dr. Rong Wang, Xinyao Xie, and Cheryl Rogers for their valuable

contributions to this thesis. Yihong has given me enormous help as a friend and as a senior. He

shared his datasets and gave useful suggestions when I was struggling with the research. Rong is

so supportive that she always responded to my questions with a smile as well as shared her own

experience of studying in U of T, even when she returned to China. Xinyao taught me code writing,

illuminated me with her unique learning experience, and exchanged her perspectives with me.

Cheryl helped me improve my academic writing and discussed the research with me. Besides,

thanks also go to Dr. Weiliang Fan, Dr. Zhaoying Zhang, etc. for the happy time we had during

my life in Toronto.

Finally, I want to express my sincere gratitude to my parents for their support of my dream of

studying abroad all the time. They are always open-minded and giving me the freedom to take my

own road. A particular β€˜Thank you’ also goes to my dear girlfriend, Jing Zhang, for the company

and help during this difficult and emotionally challenging year.

Page 4: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

iv

Table of Contents

Acknowledgments.......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................ vii

List of Figures .............................................................................................................................. viii

Glossary of Acronyms and Abbreviations ..................................................................................... xi

Chapter 1 ..........................................................................................................................................1

Introduction .................................................................................................................................1

1.1 Introduction of π‘‰π‘π‘šπ‘Žπ‘₯ and methods of estimating π‘‰π‘π‘šπ‘Žπ‘₯ ................................................2

1.1.1 Definition of π‘‰π‘π‘šπ‘Žπ‘₯ ................................................................................................2

1.1.2 π‘‰π‘π‘šπ‘Žπ‘₯25 estimation from field measurements and flux measurements ................2

1.1.3 π‘‰π‘π‘šπ‘Žπ‘₯25 estimation from remote sensing .............................................................3

1.1.3.1 Direct correlations between π‘‰π‘π‘šπ‘Žπ‘₯25 and VIs ........................................3

1.1.3.2 Indirect estimation of π‘‰π‘π‘šπ‘Žπ‘₯25 from other parameters ...........................4

1.2 Introduction of spatial scaling, upscaling, and downscaling in remote sensing ..................5

1.2.1 Introduction of intra-pixel spatial heterogeneity......................................................5

1.2.2 Introduction to spatial scaling ..................................................................................5

1.2.3 Introduction to upscaling and downscaling .............................................................6

1.3 Significance of downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 ..............................................................................8

1.3.1 Introduction of the eddy covariance and flux tower measurements ........................8

1.3.2 Building a bridge linking π‘‰π‘π‘šπ‘Žπ‘₯25 from coarse to high resolutions .....................9

1.4 Objectives and main structure of this research ..................................................................10

1.4.1 Research objectives ................................................................................................10

1.4.2 Structure of this research .......................................................................................11

1.5 References ..........................................................................................................................13

Chapter 2 ........................................................................................................................................19

Page 5: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

v

Trial of photochemical reflectance index (PRI) on downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 ...........................19

2.1 Introduction ........................................................................................................................19

2.2 Data and methods ...............................................................................................................21

2.2.1 Data ........................................................................................................................21

2.2.2 Estimating GPP based on PRI................................................................................22

2.2.2.1 Trial of establishing generic correlations between PRI and LUE ...........22

2.2.2.2 Trial of establishing correlations between PRI and LUE at flux sites.....23

2.2.3 Retrieving π‘‰π‘π‘šπ‘Žπ‘₯25 based on GPP estimated from PRI .....................................24

2.2.3.1 Model description ....................................................................................24

2.2.3.2 Lookup-table establishment and π‘‰π‘π‘šπ‘Žπ‘₯25 searching ............................25

2.3 Discussion ..........................................................................................................................26

2.3.1 Results and problems found in the progress ..........................................................26

2.3.1.1 Trial of establishing generic PRI-LUE correlations for each PFT ..........26

2.3.1.2 Trial of establishing PRI-LUE correlations at each site ..........................30

2.3.2 Further work for solving the issues ........................................................................33

2.4 References ..........................................................................................................................34

Chapter 3 ........................................................................................................................................37

Leaf chlorophyll content (LCC) as a feasible way for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 .......................37

3.1 Introduction ........................................................................................................................37

3.2 Data and methods ...............................................................................................................39

3.2.1 Data ........................................................................................................................39

3.2.2 Algorithms for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 .................................................................41

3.2.3 Evaluation and Statistical Analysis ........................................................................42

3.3 Results and discussion .......................................................................................................43

3.3.1 The intra-pixel heterogeneity of the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 product .....................43

3.3.2 The π‘‰π‘π‘šπ‘Žπ‘₯25 seasonal variation ..........................................................................47

Page 6: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

vi

3.3.3 Evaluation of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 at sites.........................................................49

3.3.3.1 The comparison of GPP simulation results .............................................49

3.3.3.2 The seasonal variation of GPP simulation results ...................................51

3.3.3.3 Statistical analysis of GPP simulation results .........................................54

3.3.3.4 GPP responses to π‘‰π‘π‘šπ‘Žπ‘₯25 before and after downscaling ....................56

3.3.4 Applying the downscaling method to regional and global scales ..........................58

3.4 Conclusion .........................................................................................................................63

3.5 References ..........................................................................................................................65

Chapter 4 ........................................................................................................................................69

Summary ...................................................................................................................................69

4.1 Main conclusions ...............................................................................................................69

4.2 Limitation of current work and plan for further work .......................................................71

4.3 References ..........................................................................................................................73

Page 7: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

vii

List of Tables

Table 2-1 Basic information of datasets for establishing the lookup table. ................................. 22

Table 2-2 The abbreviations and full forms of the nine PFTs. ..................................................... 23

Table 2-3 Squared Pearson correlation coefficient between PRI and LUE using MODIS bands

10, 12, and 13 at 190 sites for nine plant functional types............................................................ 29

Table 2-4 Squared Pearson correlation coefficient between sPRI and LUE using MODIS band

10, 12, and 13 at five sites............................................................................................................. 30

Table 3-1 Site specifications for five flux sites ............................................................................ 39

Table 3-2 Statistical evaluation results of the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 based on GPP simulated

with π‘‰π‘π‘šπ‘Žπ‘₯25 before and after downscaling against GPP derived from tower flux measurements

....................................................................................................................................................... 55

Table 3-3 The summary of GPP responses to π‘‰π‘π‘šπ‘Žπ‘₯25 before and after downscaling .............. 56

Page 8: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

viii

List of Figures

Figure 1-1 Variations of LAI retrieval biases of coarse resolution pixels with different water area

fraction using an NDVI-based non-linear LAI retrieval algorithm, expressed as the relative

difference in LAI. Source: Chen (1999). ........................................................................................ 6

Figure 1-2 Basic operations involving upscaling and downscaling in remote sensing. Source:

Bierkens et al. (2000). ..................................................................................................................... 7

Figure 1-3 Variation of global annual GPP with π‘‰π‘π‘šπ‘Žπ‘₯25 in the sensitivity analysis of the High-

Dimensional Model Representation (HDMR). Source: Ziehn and Tomlin (2017). ....................... 9

Figure 1-4 The theoretical framework and flowchart of downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 and evaluation.11

Figure 2-1 The xanthophyll cycle involving the de-epoxidation and epoxidation of xanthophyll

pigments. Source: Demmig-Adams (1990). ................................................................................. 19

Figure 2-2 PRI-LUE correlations for each PFT, using band 10 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot................ 27

Figure 2-3 PRI-LUE correlations for each PFT, using band 12 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot................ 28

Figure 2-4 PRI-LUE correlations for each PFT, using band 13 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot................ 29

Figure 2-5 sPRI-LUE correlations at each site using bands 10, 12, and 13. Unit of LUE: gC/MJ.

The blue lines are regression lines of the scatter points in each subplot. ..................................... 32

Figure 3-1 Typical leaf reflectance spectra, from 400nm to 2500nm. Source: Croft and Chen

(2018). ........................................................................................................................................... 38

Figure 3-2 The correlation between MTCI and measured leaf chlorophyll content. Source: Croft

et al. (2014). .................................................................................................................................. 40

Figure 3-3 Intra-pixel heterogeneous π‘‰π‘π‘šπ‘Žπ‘₯25 distribution within the TROPOMI pixel of each

site on the closest cloud-free day to the day of year 190. The π‘‰π‘π‘šπ‘Žπ‘₯25 values at 1 kmΓ—1 km are

Page 9: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

ix

downscaled through LCC. The red point represents the relative location of the site (1 kmΓ—1 km)

in the TROPOMI pixel (0.1°×0.1Β°). ............................................................................................. 45

Figure 3-4 Intra-pixel heterogeneous π‘‰π‘π‘šπ‘Žπ‘₯25 distribution within the TROPOMI pixel of each

site on the closest cloud-free day to the day of year 190. The π‘‰π‘π‘šπ‘Žπ‘₯25 values at 1 kmΓ—1 km are

downscaled through NDVI. The red point represents the relative location of the site (1 kmΓ—1

km) in the TROPOMI pixel (0.1°×0.1Β°). ..................................................................................... 46

Figure 3-5 π‘‰π‘π‘šπ‘Žπ‘₯25 seasonal variation of each site from the day of year 100 to 300................ 48

Figure 3-6 GPP scatter plot of each site from the day of year 100 to 300. The Y-axis represents

GPP estimates and the X-axis represents flux measured GPP. ..................................................... 51

Figure 3-7 GPP scatter plot of all sites from the day of year 100 to 300...................................... 51

Figure 3-8 GPP seasonal variation of each site from the day of year 100 to 300. The daily EC

GPP data were merged from half-hourly EC measurements. The daily modeled GPP data were

merged from hourly BEPS GPP simulation results. ..................................................................... 52

Figure 3-9 Temporal patterns of the bias of GPP estimated with TROPOMI and downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 ...................................................................................................................................... 53

Figure 3-10 GPP responses to π‘‰π‘π‘šπ‘Žπ‘₯25 in the BEPS model. The Y-axis represents the GPP

estimates. Other inputs are kept consistent, using data on the day of year 124 at US-WCr site.

The X-axis represents π‘‰π‘π‘šπ‘Žπ‘₯25 values changing from 0 to 80 Β΅mol/m2/s. The blue curve shows

the GPP responses to change of π‘‰π‘π‘šπ‘Žπ‘₯25. The black line is the straightened version of the curve

between mean π‘‰π‘π‘šπ‘Žπ‘₯25 – SD and π‘‰π‘π‘šπ‘Žπ‘₯25 + SD. The blue point represents lumped GPP,

simulated using π‘‰π‘π‘šπ‘Žπ‘₯25 at the 0.1˚ resolution. The red point represents distributed, the mean

of GPP simulated based on the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 within the 0.1˚ pixel. .................... 57

Figure 3-11 Spatial distribution of standard deviation of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values in

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 pixels on the day of year 200. ................................................................... 58

Figure 3-12 Comparison of global π‘‰π‘π‘šπ‘Žπ‘₯25 maps on the day of year 200. a) The 0.1°×0.1Β°

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 map; b) The downscaled 1 kmΓ—1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map. ........................... 59

Page 10: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

x

Figure 3-13 Comparison of π‘‰π‘π‘šπ‘Žπ‘₯25 maps of North America on the day of year 150. a) The

0.1°×0.1Β° TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 map; b) The downscaled 1 kmΓ—1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map; c) and d)

Partially enlarged details of the red labeled region in a) and b); e) and f) Partially enlarged details

of the red labeled region in c) and d). ........................................................................................... 60

Figure 3-14 Comparison of π‘‰π‘π‘šπ‘Žπ‘₯25 maps of North America on the day of year 200. a) The

0.1°×0.1Β° TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 map; b) The downscaled 1 kmΓ—1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map; c) and d)

Partially enlarged details of the red labeled region in a) and b); e) and f) Partially enlarged details

of the red labeled region in c) and d). ........................................................................................... 61

Figure 3-15 Comparison of π‘‰π‘π‘šπ‘Žπ‘₯25 maps of North America on the day of year 250. a) The

0.1°×0.1Β° TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 map; b) The downscaled 1 kmΓ—1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map; c) and d)

Partially enlarged details of the red labeled region in a) and b); e) and f) Partially enlarged details

of the red labeled region in c) and d). ........................................................................................... 62

Page 11: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

xi

Glossary of Acronyms and Abbreviations

APAR Absorbed Photosynthetically Active Radiation

BEPS Boreal Ecosystem Productivity Simulator

CI Clumping Index

EC Eddy Covariance

FPAR Fraction of Photosynthetically Active Radiation Absorbed by Vegetation

GEE Google Earth Engine

GPP Gross Primary Productivity

IOA Index Of Agreement

LAI Leaf Area Index

LCC Leaf Chlorophyll Content

LUE Light Use Efficiency

MAE Mean Absolute Error

MERIS MEdium Resolution Imaging Spectrometer

MODIS MODerate-resolution Imaging Spectroradiometer

MTCI MERIS Terrestrial Chlorophyll Index

NDVI Normalized Difference Vegetation Index

PAR Photosynthetically Active Radiation

PFT Plant Functional Type

PRI Photochemical Reflectance Index

R2 Squared Pearson correlation coefficient

RMSE Root Mean Square Error

SIF Sun-Induced chlorophyll Fluorescence

TROPOMI TROPOspheric Monitoring Instrument

π‘‰π‘π‘šπ‘Žπ‘₯ The maximum carboxylation rate

π‘‰π‘π‘šπ‘Žπ‘₯25 π‘‰π‘π‘šπ‘Žπ‘₯ normalized to 25Β°C

VI Vegetation Index

Page 12: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

1

Chapter 1

Introduction

The carbon cycle is an essential part of the earth system dynamics. To understand the carbon cycle,

various models have been developed to simulate carbon cycle processes (IPCC, 2013). Terrestrial

ecosystems play an essential role in the climate system through carbon cycling among vegetation,

soil, and atmosphere (Cao and Woodward, 1998). Process-based models are important for

understanding the terrestrial carbon cycle. In the widely-adopted Farquhar’s scheme (Farquhar et

al., 1980), the maximum carboxylation velocity (π‘‰π‘π‘šπ‘Žπ‘₯) is a crucial parameter in modeling gross

primary productivity (GPP). The magnitude of π‘‰π‘π‘šπ‘Žπ‘₯ exerts an impact on the magnitude of GPP,

and the uncertainty in π‘‰π‘π‘šπ‘Žπ‘₯ will propagate through the model and be even magnified by the model

(Bonan et al., 2011, Chen et al., 2011), inducing errors in the final simulation results. Therefore,

reliable π‘‰π‘π‘šπ‘Žπ‘₯ datasets are prerequisite for accurate GPP modeling.

Remote sensing offers continuous observations of the globe, providing frequent and extensive

coverage of terrestrial ecosystems. Reflectance measurements taken by satellite sensors have been

successfully used for estimating π‘‰π‘π‘šπ‘Žπ‘₯, through vegetation indices (VIs), leaf chlorophyll content,

sun-induced chlorophyll fluorescence, etc. (Croft et al., 2017; He et al., 2019; Jin et al., 2012; Zhou

et al., 2014). Those measurements are at moderate spatial resolutions ranging from several hundred

meters (MERIS/ENVISAT, MODIS/TERRA) to several kilometers (TROPOMI/Sentinel-5P). At

such moderate resolutions, the field of view of the measurements would be heterogeneous due to

the nature of the land (Garrigues et al., 2006b). Radiometric sensors integrate the surface

reflectance over each pixel, so the intra-pixel heterogeneous information is lost during

measurement. The intra-pixel spatial heterogeneity causes biases in retrieved parameters if the

retrieval algorithm is nonlinear (Chen, 1999; Hu and Islam, 1997; Raffy, 1994; Tian et al., 2002).

Therefore, the intra-pixel spatial heterogeneity has been a subject of intensive studies for the

purpose to accurately retrieve land surface parameters (Chasmer et al., 2009; Duveiller and

Cescatti, 2016; Hong et al., 2011; Kim and Barros, 2002; Piles et al., 2011).

In this thesis, three factors were selected to investigate the intra-pixel spatial heterogeneity of

π‘‰π‘π‘šπ‘Žπ‘₯ and to examine if they can provide information for downscaling the π‘‰π‘π‘šπ‘Žπ‘₯ dataset derived

Page 13: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

2

from sun-induced chlorophyll fluorescence. The factors are photochemical reflectance index

(PRI), leaf chlorophyll content (LCC), and normalized difference vegetation index (NDVI).

This chapter introduces the background and main structure of this study. It serves three purposes:

1) to review various methods of estimating π‘‰π‘π‘šπ‘Žπ‘₯ ; 2) to review concepts of spatial scaling,

upscaling, and downscaling using remote sensing images; 3) to introduce the eddy covariance

techniques and state the significance of downscaling π‘‰π‘π‘šπ‘Žπ‘₯ ; and 4) to outline the research

objectives and main structure of this thesis.

1.1 Introduction of π‘‰π‘π‘šπ‘Žπ‘₯ and methods of estimating π‘‰π‘π‘šπ‘Žπ‘₯

1.1.1 Definition of π‘‰π‘π‘šπ‘Žπ‘₯

Terrestrial ecosystems β€œbreathe” in carbon dioxide through the photosynthetic process and

β€œrelease” carbon dioxide into the atmosphere by autotrophic respiration and heterotrophic

respiration (Schimel, 1995). The feedbacks from the terrestrial carbon cycle can dramatically exert

effects on the biosphere-atmosphere carbon fluxes and the future climate change (Schimel et al.,

2015). Photosynthesis is the key driver of the terrestrial carbon cycle (Cadule et al., 2010; Canadell

et al., 2007), and it is an essential part of carbon cycle models (Bonan et al., 2011; Sitch et al.,

2003). Within the photosynthetic process, carboxylation fixes carbon dioxide in the air into

carbohydrates by adding carbon dioxide to ribulose 1,5 bisphosphate. In the Farquhar–von

Caemmerer–Berry model (Farquhar et al., 1980), π‘‰π‘π‘šπ‘Žπ‘₯ (the maximum carboxylation rate) is a

fundamental parameter in simulating the photosynthetic activity of vegetation, which determines

the maximum photosynthetic capacity of leaves, and directly influences the amount of gross

primary productivity (GPP) in terrestrial ecosystems (Cramer and Field, 1999; Running et al.,

2004). In most process-based models, the π‘‰π‘π‘šπ‘Žπ‘₯ normalized to 25 Β°C (π‘‰π‘π‘šπ‘Žπ‘₯25 ) is used. There are

several methods to estimate π‘‰π‘π‘šπ‘Žπ‘₯25 , including the field measurements, the flux measurements, and

the remote sensing.

1.1.2 π‘‰π‘π‘šπ‘Žπ‘₯25 estimation from field measurements and flux measurements

π‘‰π‘π‘šπ‘Žπ‘₯25 can be estimated based on field measurements. As a key parameter of understanding the

capacity of a leaf for CO2 assimilation, π‘‰π‘π‘šπ‘Žπ‘₯25 can be retrieved through gas-exchange process

analysis (Harley and Baldocchi, 1995; Wullschleger, 1993). π‘‰π‘π‘šπ‘Žπ‘₯25 is obtained from the A/Ci

curves by measuring the photosynthesis rates under different carbon dioxide levels, where A

Page 14: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

3

represents the leaf photosynthesis rate and Ci stands for the intercellular carbon dioxide

concentration of the leaf. π‘‰π‘π‘šπ‘Žπ‘₯25 values of different species and different plant functional types

(PFTs) have been measured from field experiments in many studies (Kosugi and Matsuo, 2006;

Kosugi et al., 2003). Bahar et al. (2017) compared π‘‰π‘π‘šπ‘Žπ‘₯25 values of 210 species at 18 field sites in

tropical moist forests. These field measurements provide substantial π‘‰π‘π‘šπ‘Žπ‘₯25 data for global carbon

cycle simulation. However, the field gas exchange experiments cannot provide π‘‰π‘π‘šπ‘Žπ‘₯25 values for

large geographical areas, for it is time and labor-consuming for data collections.

To address the limitations of gas exchange experiments, many studies utilized flux measurements

using the eddy covariance technique to improve the efficiency and accuracy of π‘‰π‘π‘šπ‘Žπ‘₯25 estimation

(Wolf et al., 2006). Eddy covariance is a micro-meteorological method which can directly observe

the energy and gas exchange between ecosystems and the atmosphere (Liang et al., 2012a). The

eddy covariance technique can directly, accurately, and continuously measure the carbon exchange

and evapotranspiration of an ecosystem (Baldocchi, 2008). Many studies estimated π‘‰π‘π‘šπ‘Žπ‘₯25 based

on eddy covariance flux measurements. Wang et al. (2007) applied data assimilation techniques to

invert models to derive π‘‰π‘π‘šπ‘Žπ‘₯25 values from flux measurements. The π‘‰π‘π‘šπ‘Žπ‘₯

25 values estimated from

eddy covariance flux data agree well with the π‘‰π‘π‘šπ‘Žπ‘₯25 retrieved from field measurements, indicating

the usefulness of flux data for π‘‰π‘π‘šπ‘Žπ‘₯25 estimates (Zheng et al., 2017).

1.1.3 π‘‰π‘π‘šπ‘Žπ‘₯25 estimation from remote sensing

Remote sensing provides spatially continuous observation of the Earth's surface, covering

extensive spatial and temporal land surface processes at the global scale. Based on the reflectance

spectra of ground objects, data from satellite sensors offer information on biophysical and

physiological characteristics of terrestrial ecosystems. However, remotely measured reflectance

cannot estimate π‘‰π‘π‘šπ‘Žπ‘₯25 , because the variations of π‘‰π‘π‘šπ‘Žπ‘₯

25 cannot lead to directly detectable spectral

signals in the reflectance received by the sensors. To solve this issue, many efforts have been made

to retrieve π‘‰π‘π‘šπ‘Žπ‘₯25 indirectly from remotely sensed data.

1.1.3.1 Direct correlations between π‘‰π‘π‘šπ‘Žπ‘₯25 and VIs

Vegetation indices (VIs) are compositions of reflectances in several spectral bands of remotely

sensed data to assess a particular property of vegetation. VIs designed for assessing vegetation

physiological status are often composed of spectral bands that are sensitive to the physiological

Page 15: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

4

change of vegetation, often together with bands as references to exclude the influence of the plant

structure and the background. Various VIs have been designed to trace structural, biophysical and

physiological traits of vegetation from remote sensing data, and have been successfully used to

estimate structural parameters, including leaf area index (Chen and Cihlar, 1996; Zheng and

Moskal, 2009) and clumping index (Chen et al., 2005; He et al., 2012) and physiological

parameters, including leaf chlorophyll content (Croft et al., 2014). Wang et al. (2008) found the

correlation between π‘‰π‘π‘šπ‘Žπ‘₯25 and the broadband simple ratio for beech stands in the cold-temperate

zone of Japan and concluded different VI-π‘‰π‘π‘šπ‘Žπ‘₯25 correlations at different elevations. Zhou et al.

(2014) observed close VI-π‘‰π‘π‘šπ‘Žπ‘₯25 relationships in deciduous and mixed forests. Jin et al. (2012)

investigated the correlation between NDVI and π‘‰π‘π‘šπ‘Žπ‘₯25 and found tight exponential relationships,

showing the feasibility of NDVI to derive the interannual trajectory of photosynthetic capacity.

However, the interseasonal and interannual variations of VI- π‘‰π‘π‘šπ‘Žπ‘₯25 relationships restrict the

reliability of the retrieval of π‘‰π‘π‘šπ‘Žπ‘₯25 through VIs (Croft et al., 2020).

1.1.3.2 Indirect estimation of π‘‰π‘π‘šπ‘Žπ‘₯25 from other parameters

Sun-induced chlorophyll fluorescence (SIF) has been widely used to study and track vegetation

traits (Guan et al., 2016; Guanter et al., 2014; He et al., 2017; Joiner et al., 2013; KΓΆhler et al.,

2018; Sun et al., 2017). Zhang et al. (2014) generated unique relationships between SIF and π‘‰π‘π‘šπ‘Žπ‘₯25 ,

and obtained π‘‰π‘π‘šπ‘Žπ‘₯25 values using remotely sensed SIF through model inversion. He et al. (2019)

retrieved a π‘‰π‘π‘šπ‘Žπ‘₯25 dataset from a data assimilation system based on the significant correlation

between SIF and GPP, showing the feasibility of using satellite SIF data to retrieve the information

on leaf photosynthetic capacity. Liu (2019) adopted the main framework of He et al. (2019) and

used the SIF data from the newly launched sensor, TROPOMI onboard Sentinel-5P, to produce a

π‘‰π‘π‘šπ‘Žπ‘₯25 map. Besides, leaf chlorophyll content (LCC), which is an essential parameter responsible

for the light harvest as part of photosynthetic processes, can be retrieved from remote sensing data

(Croft et al., 2020; Croft et al., 2014; Croft et al., 2015). Croft et al. (2017) found a significant

relationship between LCC and π‘‰π‘π‘šπ‘Žπ‘₯25 , and assessed the feasibility of LCC retrieved from satellite

data to estimate π‘‰π‘π‘šπ‘Žπ‘₯25 . Other studies also show strong correlations between LCC and π‘‰π‘π‘šπ‘Žπ‘₯

25

(HomolovΓ‘ et al., 2013; Houborg et al., 2013), suggesting the potential for retrieving π‘‰π‘π‘šπ‘Žπ‘₯25 over

large areas based on LCC. π‘‰π‘π‘šπ‘Žπ‘₯25 can also be quantified considering leaf nitrogen content and

nitrogen use efficiency (Kattge et al., 2009).

Page 16: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

5

1.2 Introduction of spatial scaling, upscaling, and downscaling in remote sensing

1.2.1 Introduction of intra-pixel spatial heterogeneity

Remote sensing techniques provide observations covering large spatial extents at high temporal

frequencies. However, those observations are often at moderate or coarse spatial resolutions from

several hundred meters to several tens of kilometers, which may contain very heterogeneous land

surfaces in the footprint of the sensors (Garrigues et al., 2006b). The landscape features within a

moderate-resolution pixel, such as agricultural fields and vegetation patches, are relatively small

in comparison with the pixel dimension. Many such features are included within one pixel, and

this within-pixel heterogeneous information is lost in moderate resolution images. Therefore, it is

of great importance to explore how the intra-pixel spatial heterogeneity of remote sensing data

affects the retrieval of land surface parameters and the simulation results of satellite-data driven

models.

1.2.2 Introduction to spatial scaling

The spatial heterogeneity includes two components, the spatial variability and the spatial structures

(Garrigues et al., 2006b). To understand the subpixel information, intra-pixel heterogeneity and

their effects, spatial scaling has been widely adopted to link information across scales. Spatial

scaling refers to using the information at one scale to derive information at another scale (Jarvis,

1995). If the algorithms used for retrieving land surface parameters are non-linear, the intra-pixel

heterogeneity induces biases in the retrieved parameters over pixels at coarse resolutions because

the reflectance acquired at the coarse resolutions is an averaging process that masks subpixel

variations (Chen, 1999).

Several studies have been carried out to investigate the effect of spatial scaling and to correct the

biases of retrieved parameters at coarse resolutions. Garrigues et al. (2006a) founded that if the

pixel is heterogeneous and the transfer function from remote sensing data to LAI is not linear,

biases would exist in the computation of LAI, and then they proposed a model to estimate and

correct the errors in LAI estimates. Similar results in Chen (1999) showed that negative biases of

LAI estimates occurred in heterogeneous pixels. Chen et al. (2013) studied the effect of vegetation

heterogeneity and surface topography on net primary productivity (NPP) estimates at a coarse

resolution (1 km) using an eco-hydrological model applied to a fine resolution (30 m). Biases were

Page 17: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

6

introduced when averaging subpixel inputs for NPP estimation and a scaling algorithm was

developed to correct the biases in NPP retrieval at coarse resolutions. El Maayar and Chen (2006)

proposed a method to use subpixel information to correct the evapotranspiration at a coarse

resolution, considering the spatial heterogeneity of vegetation, topography, and soil texture.

Chasmer et al. (2009) assessed the influence of spatial heterogeneity on GPP estimation using

airborne light detection and ranging (Lidar), scaling from 1 m to 1000 m, and pointed out Lidar as

an appropriate method for scaling between tower flux GPP and satellite-based GPP products.

Therefore, it is essential to investigate the spatial scaling effect for accurate land surface

parameters derived from remote sensing data.

Figure 1-1 Variations of LAI retrieval biases of coarse resolution pixels with different water

area fraction using an NDVI-based non-linear LAI retrieval algorithm, expressed as the

relative difference in LAI. Source: Chen (1999).

1.2.3 Introduction to upscaling and downscaling

Differing from spatial scaling studies, which focus on the biases of retrieved surface parameters

from non-linear algorithms in heterogeneous pixels, upscaling and downscaling are aggregation

and disaggregation of the original dataset, combined with other data for better retrieval of land

surface parameters. Upscaling is defined as a decrease in spatial resolution or extrapolation from

point to grid (Bierkens et al., 2000). In situ measurements can be upscaled with remote sensing

data and models to regional scales. Ueyama et al. (2013) upscaled CO2 fluxes from 21 towers to

Page 18: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

7

estimate the Alaskan CO2 budget by combining remote sensing data based on a support vector

regression model and the predicted upscaled regional fluxes were found to be consistent with GPP

and respiration field observations. Kang et al. (2015) developed a regression Kriging model to

upscale soil moisture measurements with MODIS products and the upscaling model showed high

prediction accuracy. Fu et al. (2014) combined tower flux measurements with satellite data based

on an upscaling model framework to estimate net ecosystem exchange (NEE) at high spatial-

temporal resolutions. The modeled results showed consistency with observed data and they found

that higher spatial resolution remote sensing products with tower flux measurements resulted in

better upscaled results. Thus, the upscaling method can integrate in situ measurements with

satellite data for precise regional environmental monitoring.

Figure 1-2 Basic operations involving upscaling and downscaling in remote sensing. Source:

Bierkens et al. (2000).

On the contrary, downscaling is defined as an increase in spatial resolution and also referred to as

the disaggregation of the original dataset into finer spatial units (Bierkens et al., 2000). Data at

coarse resolutions provide useful information and the downscaled results at high resolutions are

obtained through various methods based on the coarse-resolution data. The downscaling process

requires useful and available information at that resolution and the downscaled results restore the

spatial variation at a finer scale (Price et al., 2000).

Downscaling methods have been extensively used in remote sensing to infer information at a fine

resolution from data at a coarse resolution. Hong et al. (2011) downscaled an evapotranspiration

map at 250 m resolution derived from MODIS data using Landsat imagery at 30 m resolution. The

Page 19: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

8

spatial distribution patterns of the disaggregated evapotranspiration maps were investigated from

the downscaled imagery. Duveiller and Cescatti (2016) performed a spatial downscaling of SIF,

which led to an improved temporal correlation between SIF and GPP. Their results supported that

the downscaled SIF could be used as new datasets for estimating GPP using satellite data. Kim

and Barros (2002) proposed a downscaling model to investigate the heterogeneous subpixel

information of remotely sensed soil moisture data. Their model adopted a modified fractal

interpolation method, which generated unique fractal surfaces to study the heterogeneity.

Therefore, downscaling can explore subpixel information and bring datasets from coarse to high

spatial resolutions.

1.3 Significance of downscaling π‘‰π‘π‘šπ‘Žπ‘₯25

1.3.1 Introduction of the eddy covariance and flux tower measurements

Eddy covariance, a micro-meteorological method, is prevailing in observing the exchanges of gas,

energy, and momentum between ecosystems and the atmosphere (Liang et al., 2012a, b). The eddy

covariance method can directly, precisely, and continuously measure the carbon, water, and heat

fluxes at various time scales ranging from hour, day, month to year. The spatial scales of

observations at each tower site extend through the flux footprint around the tower, ranging from

100 m to 1000 m (GΓΆckede et al., 2004). The eddy covariance technique has been proved to be the

most efficient way to measure the interactions between terrestrial ecosystems and the atmosphere

at the ecosystem scale (Baldocchi, 2008; Friend et al., 2007).

Photosynthesis happens during the day, absorbing solar energy and transforming it into vegetation

productivity, and stops at night. The observed flux at night represents the respiration activities of

the terrestrial ecosystem, including the plant autotrophic respiration and the heterotrophic soil

respiration. Without the influence of soil water, the respiration of terrestrial ecosystems (𝐸𝑅) can

be calculated as below (Liang et al., 2012a):

𝐸𝑅 = 𝐸𝑅0 Γ— 𝑄10

𝑇 βˆ’ 𝑇0𝑇0 (1-1)

where 𝐸𝑅0 is the respiration rate at the base temperature 𝑇0 . 𝑇 is the temperature at the

measurement time and 𝑄10 is the temperature-sensitivity factor of ecosystem respiration, which is

defined as the increase of the ecosystem respiration rate with the increase in temperature every 10

Page 20: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

9

Β°C. Therefore, when photosynthesis, plant autotrophic, and heterotrophic soil respiration co-occur

during the day, the observed flux represents the net ecosystem productivity (NEP) and GPP can

be obtained as:

𝐺𝑃𝑃 = 𝑁𝐸𝑃 + πΈπ‘…π‘‘π‘Žπ‘¦ (1-2)

where πΈπ‘…π‘‘π‘Žπ‘¦ represents the ecosystem respiration during the day, and it is positive for losing

carbon from the ecosystem. NEP is positive for carbon uptake by the ecosystem. In this way, GPP

at each tower site can be accurately and continuously measured.

1.3.2 Building a bridge linking π‘‰π‘π‘šπ‘Žπ‘₯25 from coarse to high resolutions

As one of the key parameters in terrestrial biosphere process-based models, π‘‰π‘π‘šπ‘Žπ‘₯25 accounts for the

photosynthetic capacity of plants. However, little attention has been drawn to the intra-pixel spatial

heterogeneity of π‘‰π‘π‘šπ‘Žπ‘₯25 . Since GPP is sensitive to the variation of π‘‰π‘π‘šπ‘Žπ‘₯

25 (Ziehn and Tomlin, 2017),

biases in π‘‰π‘π‘šπ‘Žπ‘₯25 would induce errors in GPP estimates, as shown in Figure 1-3. Currently, the

spatial resolution of π‘‰π‘π‘šπ‘Žπ‘₯25 datasets derived from remote sensing data ranges from several to

several tens of kilometers while tower flux footprints are only about one kilometer. Therefore,

there is an apparent mismatch between the spatial resolutions of tower flux footprints and available

π‘‰π‘π‘šπ‘Žπ‘₯25 datasets derived from satellite data. This mismatch will induce biases in GPP estimates using

flux tower measurements as model inputs if the intra-pixel π‘‰π‘π‘šπ‘Žπ‘₯25 values are not homogeneous.

Thus, the intra-pixel spatial distribution of π‘‰π‘π‘šπ‘Žπ‘₯25 datasets needs to be explored.

Figure 1-3 Variation of global annual GPP with π‘½π’„π’Žπ’‚π’™πŸπŸ“ in the sensitivity analysis of the High-

Dimensional Model Representation (HDMR). Source: Ziehn and Tomlin (2017).

Page 21: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

10

In order to better understand the intra-pixel heterogeneous information of π‘‰π‘π‘šπ‘Žπ‘₯25 datasets, it is

necessary to downscale π‘‰π‘π‘šπ‘Žπ‘₯25 with the useful and available information at a finer resolution. This

will also enhance the accuracy of GPP estimates at a finer resolution within the pixel such as tower

flux sites and support investigation into the intra-pixel spatial distribution of π‘‰π‘π‘šπ‘Žπ‘₯25 . Hence, the

downscaling process builds a bridge linking the π‘‰π‘π‘šπ‘Žπ‘₯25 datasets from coarse to high spatial

resolutions and resolves the mismatch between the π‘‰π‘π‘šπ‘Žπ‘₯25 datasets and tower flux footprints.

1.4 Objectives and main structure of this research

1.4.1 Research objectives

This research is designed to quantitively analyze the heterogeneity and the spatial distribution of

π‘‰π‘π‘šπ‘Žπ‘₯25 within the pixels of the π‘‰π‘π‘šπ‘Žπ‘₯

25 dataset produced by Liu (2019) and downscale the π‘‰π‘π‘šπ‘Žπ‘₯25

dataset to 1 km, same as the footprint of tower sites. This dataset was produced by firstly,

estimating GPP from angularly normalized TROPOMI SIF data, and then retrieval π‘‰π‘π‘šπ‘Žπ‘₯25 at the

0.1°×0.1Β° resolution from an Ensemble Kalman filter (EnKF) data assimilation system (He et al.,

2019). Photochemical reflectance index (PRI), normalized difference vegetation index (NDVI),

and leaf chlorophyll content (LCC) are selected to investigate whether they can provide useful

information at a high spatial resolution for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 . A spatial scaling algorithm is also

developed for the downscaling purpose.

The theoretical framework of this study is presented in the flowchart below.

This thesis research encompasses the following four objectives:

1. To quantify the heterogeneity within the pixels of the π‘‰π‘π‘šπ‘Žπ‘₯25 product derived from satellite sun-

induced chlorophyll fluorescence;

2. To explore a feasible and effective way to downscale π‘‰π‘π‘šπ‘Žπ‘₯25 using high-resolution remote

sensing images;

3. To demonstrate the improvement of GPP simulation at tower flux sites using the downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 .

Page 22: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

11

4. To produce the first global 1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map with the downscaling method satisfactorily

evaluated with the flux-derived GPP.

Figure 1-4 The theoretical framework and flowchart of downscaling π‘½π’„π’Žπ’‚π’™πŸπŸ“ and evaluation.

1.4.2 Structure of this research

The thesis is organized following the research objectives mentioned in Chapter 1.4.1.

In Chapter 2, PRI is selected for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 derived from satellite sun-induced chlorophyll

fluorescence from Liu (2019) at 0.1-degree resolution (approximately 11 km) to 1 km resolution.

Generic correlations of PRI-LUE for each plant functional type and unique correlations of PRI-

LUE of five sites are established based on historical data from FLUXNET and AMERIFLUX,

respectively. GPP estimates at the downscaled resolution are then derived following the

established PRI-LUE correlations. A lookup table approach for searching π‘‰π‘π‘šπ‘Žπ‘₯25 values is

established based on BEPS simulations that determine the relationship between π‘‰π‘π‘šπ‘Žπ‘₯25 and GPP

Page 23: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

12

under given meteorological conditions. The downscaling results are found to be unsatisfactory and

are analyzed, leading to the development of alternatives shown in Chapter 3.

In Chapter 3, LCC and NDVI are selected for downscaling the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset mentioned above. A

scaling algorithm for obtaining downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 at 1 km resolution is developed and a scaling

ratio is designed. To evaluate the results, the original π‘‰π‘π‘šπ‘Žπ‘₯25 and the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 are used as

inputs to the BEPS model to simulate GPP. The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 is evaluated over five different

sites from AMERIFLUX with available GPP measurements in 2018, by comparing GPP from EC

towers with GPP values simulated from the original π‘‰π‘π‘šπ‘Žπ‘₯25 and the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 . After a

downscaling method is satisfactorily evaluated with the flux-derived GPP at the tower sites, a

π‘‰π‘π‘šπ‘Žπ‘₯25 map at 1 km resolution for North America is produced as an example.

In Chapter 4, the main conclusions of this thesis are summarized. The limitations are discussed

and a plan for future work is proposed.

Page 24: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

13

1.5 References

Bahar, N.H., Ishida, F.Y., Weerasinghe, L.K., Guerrieri, R., O'Sullivan, O.S., Bloomfield, K.J.,

Asner, G.P., Martin, R.E., Lloyd, J., Malhi, Y., Phillips, O.L., Meir, P., Salinas, N., Cosio,

E.G., Domingues, T.F., Quesada, C.A., Sinca, F., Escudero Vega, A., Zuloaga Ccorimanya,

P.P., Del Aguila-Pasquel, J., Quispe Huaypar, K., Cuba Torres, I., Butron Loayza, R., Pelaez

Tapia, Y., Huaman Ovalle, J., Long, B.M., Evans, J.R., & Atkin, O.K. (2017). Leaf-level

photosynthetic capacity in lowland Amazonian and high-elevation Andean tropical moist

forests of Peru. New Phytologist, 214, 1002-1018

Baldocchi, D. (2008). Breathing of the terrestrial biosphere: lessons learned from a global network

of carbon dioxide flux measurement systems. Australian Journal of Botany, 56, 1-26

Bierkens, M.F.P., Finke, P.A., & Willigen, P.d. (2000). Upscaling and downscaling methods for

environmental research. Dordrecht: Kluwer

Bonan, G.B., Lawrence, P.J., Oleson, K.W., Levis, S., Jung, M., Reichstein, M., Lawrence, D.M.,

& Swenson, S.C. (2011). Improving canopy processes in the Community Land Model

version 4 (CLM4) using global flux fields empirically inferred from FLUXNET data.

Journal of Geophysical Research, 116

Cadule, P., Friedlingstein, P., Bopp, L., Sitch, S., Jones, C.D., Ciais, P., Piao, S.L., & Peylin, P.

(2010). Benchmarking coupled climate-carbon models against long-term atmospheric

CO2measurements. Global Biogeochemical Cycles, 24, n/a-n/a

Canadell, J.G., Le Quere, C., Raupach, M.R., Field, C.B., Buitenhuis, E.T., Ciais, P., Conway, T.J.,

Gillett, N.P., Houghton, R.A., & Marland, G. (2007). Contributions to accelerating

atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of

natural sinks. Proc Natl Acad Sci U S A, 104, 18866-18870

Cao, M., & Woodward, F.I. (1998). Dynamic responses of terrestrial ecosystem carbon cycling to

global climate change. Nature, 393, 249-252

Chasmer, L., Barr, A., Hopkinson, C., McCaughey, H., Treitz, P., Black, A., & Shashkov, A. (2009).

Scaling and assessment of GPP from MODIS using a combination of airborne lidar and

eddy covariance measurements over jack pine forests. Remote Sensing of Environment, 113,

82-93

Chen, H., Dickinson, R.E., Dai, Y., & Zhou, L. (2011). Sensitivity of simulated terrestrial carbon

assimilation and canopy transpiration to different stomatal conductance and carbon

assimilation schemes. Climate Dynamics, 36, 1037-1054

Chen, J.M. (1999). Spatial Scaling of a Remotely Sensed Surface Parameter by Contexture.

Remote Sensing of Environment, 69, 30-42

Chen, J.M., Chen, X., & Ju, W. (2013). Effects of vegetation heterogeneity and surface topography

on spatial scaling of net primary productivity. Biogeosciences, 10, 4879-4896

Chen, J.M., & Cihlar, J. (1996). Retrieving leaf area index of boreal conifer forests using Landsat

Page 25: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

14

TM images. Remote Sensing of Environment, 55, 153-162

Chen, J.M., Menges, C.H., & Leblanc, S.G. (2005). Global mapping of foliage clumping index

using multi-angular satellite data. Remote Sensing of Environment, 97, 447-457

Cramer, W., & Field, C.B. (1999). Comparing global models of terrestrial net primary productivity

(NPP): introduction. Global Change Biology, 5, iii-iv

Croft, H., Chen, J.M., Luo, X., Bartlett, P., Chen, B., & Staebler, R.M. (2017). Leaf chlorophyll

content as a proxy for leaf photosynthetic capacity. Global Change Biology, 23, 3513-3524

Croft, H., Chen, J.M., Wang, R., Mo, G., Luo, S., Luo, X., He, L., Gonsamo, A., Arabian, J., Zhang,

Y., Simic-Milas, A., Noland, T.L., He, Y., HomolovΓ‘, L., MalenovskΓ½, Z., Yi, Q., Beringer,

J., Amiri, R., Hutley, L., Arellano, P., Stahl, C., & Bonal, D. (2020). The global distribution

of leaf chlorophyll content. Remote Sensing of Environment, 236, 111479

Croft, H., Chen, J.M., & Zhang, Y. (2014). The applicability of empirical vegetation indices for

determining leaf chlorophyll content over different leaf and canopy structures. Ecological

Complexity, 17, 119-130

Croft, H., Chen, J.M., Zhang, Y., Simic, A., Noland, T.L., Nesbitt, N., & Arabian, J. (2015).

Evaluating leaf chlorophyll content prediction from multispectral remote sensing data

within a physically-based modelling framework. ISPRS Journal of Photogrammetry and

Remote Sensing, 102, 85-95

Duveiller, G., & Cescatti, A. (2016). Spatially downscaling sun-induced chlorophyll fluorescence

leads to an improved temporal correlation with gross primary productivity. Remote Sensing

of Environment, 182, 72-89

El Maayar, M., & Chen, J.M. (2006). Spatial scaling of evapotranspiration as affected by

heterogeneities in vegetation, topography, and soil texture. Remote Sensing of Environment,

102, 33-51

Farquhar, G.D., von Caemmerer, S., & Berry, J.A. (1980). A biochemical model of photosynthetic

CO2 assimilation in leaves of C3 species. Planta, 149, 78-90

Friend, A.D., Arneth, A., Kiang, N.Y., Lomas, M., OgÉE, J., RÖDenbeck, C., Running, S.W.,

Santaren, J.-D., Sitch, S., Viovy, N., Ian Woodward, F., & Zaehle, S. (2007). FLUXNET

and modelling the global carbon cycle. Global Change Biology, 13, 610-633

Fu, D., Chen, B., Zhang, H., Wang, J., Black, T.A., Amiro, B.D., Bohrer, G., Bolstad, P., Coulter,

R., Rahman, A.F., Dunn, A., McCaughey, J.H., Meyers, T., & Verma, S. (2014). Estimating

landscape net ecosystem exchange at high spatial–temporal resolution based on Landsat

data, an improved upscaling model framework, and eddy covariance flux measurements.

Remote Sensing of Environment, 141, 90-104

Garrigues, S., Allard, D., Baret, F., & Weiss, M. (2006a). Influence of landscape spatial

heterogeneity on the non-linear estimation of leaf area index from moderate spatial

resolution remote sensing data. Remote Sensing of Environment, 105, 286-298

Page 26: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

15

Garrigues, S., Allard, D., Baret, F., & Weiss, M. (2006b). Quantifying spatial heterogeneity at the

landscape scale using variogram models. Remote Sensing of Environment, 103, 81-96

GΓΆckede, M., Rebmann, C., & Foken, T. (2004). A combination of quality assessment tools for

eddy covariance measurements with footprint modelling for the characterisation of

complex sites. Agricultural and Forest Meteorology, 127, 175-188

Guan, K., Berry, J.A., Zhang, Y., Joiner, J., Guanter, L., Badgley, G., & Lobell, D.B. (2016).

Improving the monitoring of crop productivity using spaceborne solar-induced

fluorescence. Global Change Biology, 22, 716-726

Guanter, L., Zhang, Y., Jung, M., Joiner, J., Voigt, M., Berry, J.A., Frankenberg, C., Huete, A.R.,

Zarco-Tejada, P., Lee, J.E., Moran, M.S., Ponce-Campos, G., Beer, C., Camps-Valls, G.,

Buchmann, N., Gianelle, D., Klumpp, K., Cescatti, A., Baker, J.M., & Griffis, T.J. (2014).

Global and time-resolved monitoring of crop photosynthesis with chlorophyll fluorescence.

Proc Natl Acad Sci U S A, 111, E1327-1333

Harley, P.C., & Baldocchi, D.D. (1995). Scaling carbon dioxide and water vapour exchange from

leaf to canopy in a deciduous forest. I. Leaf model parametrization. Plant, Cell &

Environment, 18, 1146-1156

He, L., Chen, J.M., Liu, J., Mo, G., & Joiner, J. (2017). Angular normalization of GOME-2 Sun-

induced chlorophyll fluorescence observation as a better proxy of vegetation productivity.

Geophysical Research Letters, 44, 5691-5699

He, L., Chen, J.M., Liu, J., Zheng, T., Wang, R., Joiner, J., Chou, S., Chen, B., Liu, Y., Liu, R., &

Rogers, C. (2019). Diverse photosynthetic capacity of global ecosystems mapped by

satellite chlorophyll fluorescence measurements. Remote Sensing of Environment, 232,

111344

He, L., Chen, J.M., Pisek, J., Schaaf, C.B., & Strahler, A.H. (2012). Global clumping index map

derived from the MODIS BRDF product. Remote Sensing of Environment, 119, 118-130

HomolovΓ‘, L., MalenovskΓ½, Z., Clevers, J.G.P.W., GarcΓ­a-Santos, G., & Schaepman, M.E. (2013).

Review of optical-based remote sensing for plant trait mapping. Ecological Complexity, 15,

1-16

Hong, S.-H., Hendrickx, J.M.H., & Borchers, B. (2011). Down-scaling of SEBAL derived

evapotranspiration maps from MODIS (250 m) to Landsat (30 m) scales. International

Journal of Remote Sensing, 32, 6457-6477

Houborg, R., Cescatti, A., Migliavacca, M., & Kustas, W.P. (2013). Satellite retrievals of leaf

chlorophyll and photosynthetic capacity for improved modeling of GPP. Agricultural and

Forest Meteorology, 177, 10-23

Hu, Z., & Islam, S. (1997). A framework for analyzing and designing scale invariant remote

sensing algorithms. IEEE Transactions on Geoscience and Remote Sensing, 35, 747-755

IPCC (2013). Climate Change 2013: The Physical Science Basis. Contribution of Working Group

Page 27: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

16

I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.

Cambridge, United Kingdom and New York, NY, USA: Cambridge University Press

Jarvis, P.G. (1995). Scaling processes and problems. Plant, Cell & Environment, 18, 1079-1089

Jin, P., Wang, Q., Iio, A., & Tenhunen, J. (2012). Retrieval of seasonal variation in photosynthetic

capacity from multi-source vegetation indices. Ecological Informatics, 7, 7-18

Joiner, J., Guanter, L., Lindstrot, R., Voigt, M., Vasilkov, A.P., Middleton, E.M., Huemmrich, K.F.,

Yoshida, Y., & Frankenberg, C. (2013). Global monitoring of terrestrial chlorophyll

fluorescence from moderate-spectral-resolution near-infrared satellite measurements:

methodology, simulations, and application to GOME-2. Atmospheric Measurement

Techniques, 6, 2803-2823

Kang, J., Jin, R., & Li, X. (2015). Regression Kriging-Based Upscaling of Soil Moisture

Measurements From a Wireless Sensor Network and Multiresource Remote Sensing

Information Over Heterogeneous Cropland. IEEE Geoscience and Remote Sensing Letters,

12, 92-96

Kattge, J., Knorr, W., Raddatz, T., & Wirth, C. (2009). Quantifying photosynthetic capacity and its

relationship to leaf nitrogen content for global-scale terrestrial biosphere models. Global

Change Biology, 15, 976-991

Kim, G., & Barros, A.P. (2002). Downscaling of remotely sensed soil moisture with a modified

fractal interpolation method using contraction mapping and ancillary data. Remote Sensing

of Environment, 83, 400-413

KΓΆhler, P., Guanter, L., Kobayashi, H., Walther, S., & Yang, W. (2018). Assessing the potential of

sun-induced fluorescence and the canopy scattering coefficient to track large-scale

vegetation dynamics in Amazon forests. Remote Sensing of Environment, 204, 769-785

Kosugi, Y., & Matsuo, N. (2006). Seasonal fluctuations and temperature dependence of leaf gas

exchange parameters of co-occurring evergreen and deciduous trees in a temperate broad-

leaved forest. Tree Physiology, 26, 1173-1184

Kosugi, Y., Shibata, S., & Kobashi, S. (2003). Parameterization of the CO2 and H2O gas exchange

of several temperate deciduous broad-leaved trees at the leaf scale considering seasonal

changes. Plant, Cell & Environment, 26, 285-301

Liang, S., Li, X., & Wang, J. (2012a). Chapter 16 - Vegetation Production in Terrestrial Ecosystems.

In S. Liang, X. Li, & J. Wang (Eds.), Advanced Remote Sensing (pp. 501-531). Boston:

Academic Press

Liang, S., Li, X., & Wang, J. (2012b). Chapter 18 - Terrestrial Evapotranspiration. Advanced

Remote Sensing (pp. 557-588). Boston: Academic Press

Liu, Y. (2019). Master Thesis: Global Retrieval of the Maximum Carboxylation Rate of Plant

Leaves by Solar-induced Chlorophyll Fluorescence from TROPOMI. In, Department of

Geography & Planning: University of Toronto

Page 28: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

17

Piles, M., Camps, A., Vall-llossera, M., Corbella, I., Panciera, R., Rudiger, C., Kerr, Y.H., & Walker,

J. (2011). Downscaling SMOS-Derived Soil Moisture Using MODIS Visible/Infrared Data.

IEEE Transactions on Geoscience and Remote Sensing, 49, 3156-3166

Price, D.T., McKenney, D.W., Nalder, I.A., Hutchinson, M.F., & Kesteven, J.L. (2000). A

comparison of two statistical methods for spatial interpolation of Canadian monthly mean

climate data. Agricultural and Forest Meteorology, 101, 81-94

Raffy, M. (1994). Change of scale theory: a capital challenge for space observation of earth.

International Journal of Remote Sensing, 15, 2353-2357

Running, S.W., Nemani, R.R., Heinsch, F.A., Zhao, M., Reeves, M., & Hashimoto, H. (2004). A

Continuous Satellite-Derived Measure of Global Terrestrial Primary Production.

BioScience, 54, 547-560

Schimel, D., Stephens, B.B., & Fisher, J.B. (2015). Effect of increasing CO2 on the terrestrial

carbon cycle. Proc Natl Acad Sci U S A, 112, 436-441

Schimel, D.S. (1995). Terrestrial ecosystems and the carbon cycle. Global Change Biology, 1, 77-

91

Sitch, S., Smith, B., Prentice, I.C., Arneth, A., Bondeau, A., Cramer, W., Kaplan, J.O., Levis, S.,

Lucht, W., Sykes, M.T., Thonicke, K., & Venevsky, S. (2003). Evaluation of ecosystem

dynamics, plant geography and terrestrial carbon cycling in the LPJ dynamic global

vegetation model. Global Change Biology, 9, 161-185

Sun, Y., Frankenberg, C., Wood, J.D., Schimel, D.S., Jung, M., Guanter, L., Drewry, D.T., Verma,

M., Porcar-Castell, A., Griffis, T.J., Gu, L., Magney, T.S., Kohler, P., Evans, B., & Yuen, K.

(2017). OCO-2 advances photosynthesis observation from space via solar-induced

chlorophyll fluorescence. Science, 358

Tian, Y., Woodcock, C.E., Wang, Y., Privette, J.L., Shabanov, N.V., Zhou, L., Zhang, Y., Buermann,

W., Dong, J., Veikkanen, B., HΓ€me, T., Andersson, K., Ozdogan, M., Knyazikhin, Y., &

Myneni, R.B. (2002). Multiscale analysis and validation of the MODIS LAI product: I.

Uncertainty assessment. Remote Sensing of Environment, 83, 414-430

Ueyama, M., Ichii, K., Iwata, H., Euskirchen, E.S., Zona, D., Rocha, A.V., Harazono, Y., Iwama,

C., Nakai, T., & Oechel, W.C. (2013). Upscaling terrestrial carbon dioxide fluxes in Alaska

with satellite remote sensing and support vector regression. Journal of Geophysical

Research: Biogeosciences, 118, 1266-1281

Wang, Q., Iio, A., & Kakubari, Y. (2008). Broadband simple ratio closely traced seasonal trajectory

of canopy photosynthetic capacity. Geophysical Research Letters, 35, n/a-n/a

Wang, Y.P., Baldocchi, D., Leuning, R.A.Y., Falge, E.V.A., & Vesala, T. (2007). Estimating

parameters in a land-surface model by applying nonlinear inversion to eddy covariance

flux measurements from eight FLUXNET sites. Global Change Biology, 13, 652-670

Wolf, A., Akshalov, K., Saliendra, N., Johnson, D.A., & Laca, E.A. (2006). Inverse estimation of

Page 29: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

18

Vcmax, leaf area index, and the Ball-Berry parameter from carbon and energy fluxes.

Journal of Geophysical Research, 111

Wullschleger, S.D. (1993). Biochemical Limitations to Carbon Assimilation in C3 Plantsβ€”A

Retrospective Analysis of the A/Ci Curves from 109 Species. Journal of Experimental

Botany, 44, 907-920

Zhang, Y., Guanter, L., Berry, J.A., Joiner, J., van der Tol, C., Huete, A., Gitelson, A., Voigt, M.,

& Kohler, P. (2014). Estimation of vegetation photosynthetic capacity from space-based

measurements of chlorophyll fluorescence for terrestrial biosphere models. Global Change

Biology, 20, 3727-3742

Zheng, G., & Moskal, L.M. (2009). Retrieving Leaf Area Index (LAI) Using Remote Sensing:

Theories, Methods and Sensors. Sensors, 9, 2719-2745

Zheng, T., Chen, J., He, L., Arain, M.A., Thomas, S.C., Murphy, J.G., Geddes, J.A., & Black, T.A.

(2017). Inverting the maximum carboxylation rate ( V cmax ) from the sunlit leaf

photosynthesis rate derived from measured light response curves at tower flux sites.

Agricultural and Forest Meteorology, 236, 48-66

Zhou, Y., Ju, W., Sun, X., Hu, Z., Han, S., Black, T.A., Jassal, R.S., & Wu, X. (2014). Close

relationship between spectral vegetation indices and Vcmax in deciduous and mixed forests.

Tellus B: Chemical and Physical Meteorology, 66, 23279

Ziehn, T., & Tomlin, A.S. (2017). Chapter 15 - Efficient Tools for Global Sensitivity Analysis

Based on High-Dimensional Model Representation. In G.P. Petropoulos, & P.K. Srivastava

(Eds.), Sensitivity Analysis in Earth Observation Modelling (pp. 297-318): Elsevier

Page 30: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

19

Chapter 2

Trial of photochemical reflectance index (PRI)

on downscaling π‘‰π‘π‘šπ‘Žπ‘₯25

2.1 Introduction

When plants receive excessive photosynthetic active radiation, the xanthophyll pigments in leaves

dissipate the excessive energy and prevent plants from photodamage. This energy dissipation is

achieved through the interconversion of three xanthophyll pigments in the xanthophyll cycling.

When plants receive excess energy, the light-harvesting xanthophyll, violaxanthin is de-

epoxidized to the energy quenching xanthophyll, zeaxanthin via antheraxanthin, and this process

is reversed under light-limiting conditions (Demmig-Adams, 1990; Demmig-Adams and Adams,

1996). Therefore, the xanthophyll cycle is associated with the efficiency of light harvesting, thus

the photosynthetic efficiency.

Under excessive light conditions, the content of zeaxanthin increases in the xanthophyll cycle and

a decrement in the leaf reflectance at around 531nm (𝑅531) has been observed (Gamon et al., 1990).

Then, Gamon et al. (1992) proposed a narrow band index, the physiological reflectance index

(PRI), using the signal at 𝑅531 and a reference band (π‘…π‘Ÿπ‘’π‘“) to minimize the effects of other factors

on the xanthophyll signal. The original formula was presented by:

𝑃𝑅𝐼 = π‘…π‘Ÿπ‘’π‘“ βˆ’ 𝑅531

π‘…π‘Ÿπ‘’π‘“ + 𝑅531 (2-1)

Figure 2-1 The xanthophyll cycle involving the de-epoxidation and epoxidation of

xanthophyll pigments. Source: Demmig-Adams (1990).

Page 31: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

20

Gamon et al. (1992) used 𝑅550 as the reference band and observed its capacity to track diurnal

photosynthetic efficiency. The relationship between PRI and LUE was also tested under nitrogen

stressed and water stress situations in sunflower canopies. PeΓ‘Uelas et al. (1995) used 𝑅570 as the

reference band and renamed the index as photochemical reflectance index (PRI), which was

calculated as:

𝑃𝑅𝐼 = 𝑅531 βˆ’ 𝑅570

𝑅531 + 𝑅570

(2-2)

The rearranged PRI was found to yield a positive correlation with LUE and this formula became

the PRI definition.

Many studies have been done to use remotely sensed PRI to capture LUE changes across different

species, at sites, and over regional areas. Rahman et al. (2004) first used PRI derived from band

11 (bandwidth 526-536nm) and band 12 (546-556nm) of the Moderate Resolution Imaging

Spectroradiometer (MODIS) onboard the Terra and Aqua platforms to track the seasonal variation

of LUE. The correlation between PRI and LUE was used to improve the LUE model (Monteith,

1972) for better net primary productivity simulation. Drolet et al. (2005) tested MODIS bands 1

(620-670nm), 4 (545-565nm), 12, and 13 (662-672nm) as potential reference bands for PRI and

selected band 13 to correlate the scaled PRI with LUE over a boreal trembling aspen site. Drolet

et al. (2008) later tested the relationship between MODIS-derived PRI and LUE over eight boreal

eddy covariance (EC) towers and found a strong PRI-LUE correlation when all sites points were

combined. Goerner et al. (2009) used MODIS-derived PRI to track LUE under seasonal drought

conditions in a Mediterranean forest. The relationship between MODIS PRI and LUE has been

studied and tested in many studies in various ecosystems (Garbulsky et al., 2008; Guarini et al.,

2014; Middleton et al., 2016; Moreno et al., 2012), showing the capacity of MODIS PRI to track

LUE changes.

Therefore, PRI can provide useful physiological information of plants at a higher spatial resolution.

In this chapter, the PRI data obtained from MODIS are used to downscale the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset. In

Section 2.3, the data preprocessing process and the designed methodology are introduced. The

BEPS model used for establishing the lookup table is also introduced. In Section 2.4, the results

are presented. The problems found in the process are discussed, and further works for solving those

issues are proposed.

Page 32: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

21

2.2 Data and methods

2.2.1 Data

A total number of 190 sites from FLUXNET (https://fluxnet.fluxdata.org) were selected to

establish generic correlations between PRI and LUE for nine plant functional types (The basic

information of the sites can be found at https://fluxnet.fluxdata.org/sites/site-list-and-

pages/?view=table). Five sites from AMERIFLUX (https://ameriflux.lbl.gov/) were selected to

establish unique correlations between PRI and LUE at each site, including US-Bi2 (Sanchez et al.,

2017-), US-Los (Sulman et al., 2009), US-Rpf (Ueyama et al., 2019), US-Tw4 (Sanchez et al.,

2013-), and US-WCr (Cook et al., 2004). For PFT-level correlations between PRI and LUE, all

historical data from FLUXNET (2001-2015) were used. For site-level correlations between PRI

and LUE, data of five sites in 2017 were used. The established correlations were used for the GPP

estimation of 2018. The meteorological data as inputs for GPP simulation were obtained from the

flux measurements, including shortwave radiation, air temperature, vapor pressure deficit,

precipitation, and wind speed.

Google earth engine (GEE) as a cloud-based platform for planetary-scale geospatial analysis

provides an efficient cloud-computing tool for remote sensing studies (Gorelick et al., 2017).

Surface reflectance data were collected from the MODIS Terra surface spectral reflectance data

product (MODOCGA) at a resolution of 1 km from GEE. LAI and FPAR data were collected from

MODIS LAI/FPAR 4-Day 500 m product (MCD15A3H) from GEE. And the LAI/FPAR data

were smoothed and resampled to a daily 1 km product. Clumping index (CI) data used for

calculating the gap fraction of a canopy were derived from the MODIS BRDF product (He et al.,

2016). Global π‘‰π‘π‘šπ‘Žπ‘₯25 data were derived from TROPOMI SIF datasets at a resolution of 0.1 degrees

(Liu, 2019).

For establishing the lookup table to search the π‘‰π‘π‘šπ‘Žπ‘₯25 values with GPP derived from PRI, the input

data of the BEPS model corresponds to the data used by Liu (2019), as shown in Table 2-1. The

data were all spatially interpolated to a 0.1Β° Γ— 0.1Β° grid.

Page 33: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

22

Table 2-1 Basic information of datasets for establishing the lookup table.

Input data Data source Original Spatial

Resolution References

Surface wind speed

Air temperature

Shortwave radiation

Surface precipitation

Atmosphere pressure

MERRA-2 0.625Β° Γ— 0.5Β°

LAI MODIS & AVHRR 8 km Γ— 8 km Liu et al. (2012)

CI MODIS 36 km Γ— 36 km He et al. (2016)

PFT MODIS 1 km Γ— 1 km Friedl et al. (2002)

2.2.2 Estimating GPP based on PRI

The LUE model was adopted to link PRI and GPP. The LUE model was initially established from

the linear relationship between the amount of PAR absorbed by plants, and their GPP (Monteith,

1972; Monteith et al., 1977), where the APAR can be further estimated by introducing a factor

referring to as the fraction of incident PAR absorbed by plants (FPAR).

𝐺𝑃𝑃 = πΏπ‘ˆπΈ βˆ— 𝐴𝑃𝐴𝑅 (2-3)

The LUE values of each day with available data can be obtained by:

πΏπ‘ˆπΈ = 𝐺𝑃𝑃

𝑃𝐴𝑅 βˆ— 𝐹𝑃𝐴𝑅(2-4)

2.2.2.1 Trial of establishing generic correlations between PRI and LUE

To estimate the global GPP distributions of 2018 at 1 km resolution, reliable PRI-LUE correlations

should be initially established. In this part, the generic correlations were firstly established based

on all historical flux measurements from FLUXNET. Band 11 (centered at 531nm) was selected

as the signal band of PRI. Band 10 (centered at 488nm), band 12 (centered at 551nm), and band

13 (centered at 667nm) were selected as the reference band of PRI, respectively. PRI was

calculated as:

Page 34: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

23

𝑃𝑅𝐼 = 𝐡11 βˆ’ π΅π‘Ÿπ‘’π‘“

𝐡11 + π΅π‘Ÿπ‘’π‘“

(2-5)

From the first day with available data to the ending day of each site, PRIs using those three bands

were derived on cloud-free days on the pixels of each site. Then noise filters were applied to

exclude noises and errors.

Considering different physiological characteristics among different plant species, nine PFTs were

selected. The abbreviations and full forms of the nine PFTs are listed in Table 2-2 below.

Table 2-2 The abbreviations and full forms of the nine PFTs.

Abbreviation Full form

DBF Deciduous Broadleaf Forest

EBF Evergreen Broadleaf Forest

DNF Deciduous Needleleaf Forest

ENF Evergreen Needleleaf Forest

MF Mixed Forest

GRA Grassland

CRO Cropland

SH Shrubs

WET Wetlands

The GPP and PAR data derived from half-hourly flux measurements during all daytime were

averaged to daily GPP and PAR, to obtain daily LUE data. All available PRIs of each PFT were

aggregated to perform regression analysis against the corresponding LUE. The PRI reference band

with the best correlations was finally selected to derive the generic correlations for each PFT,

which would be further used to estimate the GPP of 2018 at a 1 km resolution.

2.2.2.2 Trial of establishing correlations between PRI and LUE at flux sites

In this part, unique PRI-LUE correlations at each site were established based on flux

measurements, the MODIS reflectance product, and the MODIS LAI/FPAR product (mentioned

in Chapter 2.3.1) of 2017. Band 10 (centered at 488nm), band 12 (centered at 551nm), and band

13 (centered at 667nm) were selected as the reference band of PRI, respectively. The LUE data

were calculated from the GPP and PAR data derived from half-hourly flux measurements from 10

Page 35: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

24

am to 12 pm at local time, which approximately corresponds to the crossing time of MODIS Terra

(Drolet et al., 2008). In order to obtain only positive values, PRIs using those four bands were

scaled using the mathematical transformation (Rahman et al., 2004):

𝑠𝑃𝑅𝐼 = 𝑃𝑅𝐼 + 1

2(2-6)

The scaled PRI (sPRI) derived on cloud-free days on the pixels of each site and noise filters were

applied to exclude noises and errors of sPRI. The sPRI reference band with the best correlations

was finally selected to derive the unique correlation between sPRI and LUE for that site. Then

following the established correlations of each site, the GPP estimates were retrieved.

2.2.3 Retrieving π‘‰π‘π‘šπ‘Žπ‘₯25 based on GPP estimated from PRI

2.2.3.1 Model description

In this study, the Boreal Ecosystem Productivity Simulator (BEPS) was adopted to link π‘‰π‘π‘šπ‘Žπ‘₯25 with

the GPP estimated from PRI. The BEPS model is a diagnostic enzyme kinetic model, which has

been frequently used for regional and global carbon cycle simulation (Chen et al., 2019; Gonsamo

et al., 2013; Wang et al., 2004). BEPS was first developed for the Boreal Ecosystem-Atmosphere

Study, coupling water and carbon cycles at regional scales (Liu et al., 1999, 2002; Liu et al., 1997).

BEPS adopts a two-leaf theory, separating sunlit and shaded leaves during the modeling process

(Chen et al., 1999). The total photosynthesis (Ac) at the canopy level is modeled by combining the

photosynthetic rates of sunlit and shaded leaves (𝐴𝑠𝑒𝑛 and π΄π‘ β„Žπ‘Žπ‘‘π‘’). GPP at the canopy level is

modeled by multiplying the corresponding LAI to the photosynthetic rates of sunlit and shaded

leaf groups:

𝐺𝑃𝑃 = 𝐴𝑠𝑒𝑛𝐿𝐴𝐼𝑠𝑒𝑛 + π΄π‘ β„Žπ‘Žπ‘‘π‘’πΏπ΄πΌπ‘ β„Žπ‘Žπ‘‘π‘’ (2-7)

In this study, the MODIS LAI represents the total LAI of the canopy, where 𝐿𝐴𝐼𝑠𝑒𝑛 and πΏπ΄πΌπ‘ β„Žπ‘Žπ‘‘π‘’

stand for LAI of sunlit and shaded leaves. When separating 𝐿𝐴𝐼𝑠𝑒𝑛 and πΏπ΄πΌπ‘ β„Žπ‘Žπ‘‘π‘’, the clumping

index (Ξ©) is adopted to consider the non-random distribution patterns of the leaves inside the

canopy (Chen, 1996). The clumping index characterizes the extent of clumping of the foliage,

where the Ξ© equals to 1 in canopies with randomly distributed foliage. The more clumping of the

leaves, the smaller the Ξ© is. Then 𝐿𝐴𝐼𝑠𝑒𝑛 and πΏπ΄πΌπ‘ β„Žπ‘Žπ‘‘π‘’ are calculated as:

Page 36: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

25

𝐿𝐴𝐼𝑠𝑒𝑛 = 2π‘π‘œπ‘ πœƒ (1 βˆ’ π‘’βˆ’0.5Ω𝐿𝐴𝐼

π‘π‘œπ‘ πœƒ ) (2-8)

πΏπ΄πΌπ‘ β„Žπ‘Žπ‘‘π‘’ = 𝐿𝐴𝐼 βˆ’ 𝐿𝐴𝐼𝑠𝑒𝑛 (2-9)

BEPS follows Farquhar’s principle (Farquhar et al., 1980) to calculate the photosynthetic rate of a

leaf as the minimum value between π‘Šπ‘— (radiation-limited gross photosynthesis rate) and π‘Šπ‘

(Rubisco-limited gross photosynthetic rate), where:

π‘Šπ‘— = 𝐽𝐢𝑖 βˆ’ Ξ“

4.5𝐢𝑖 + 10.5Ξ“(2-10)

π‘Šπ‘ = π‘‰π‘š

𝐢𝑖 βˆ’ Ξ“

𝐢𝑖 + 𝐾(2-11)

𝐴𝑐 = min{π‘Šπ‘— , π‘Šπ‘} βˆ’ 𝑅𝑑 (2-12)

𝐽, π‘‰π‘š, 𝐢𝑖 , Ξ“, 𝐾, and 𝑅𝑑 represent the radiation-dependent electron transport rate, the maximum

carboxylation rate, intercellular carbon dioxide concentration, temperature-dependent carbon

dioxide compensation point without dark respiration, the temperature-dependent function of

enzyme kinetics, and daytime leaf dark respiration, respectively. Following this method, with

given π‘‰π‘π‘šπ‘Žπ‘₯25 values, LAI and clumping data, meteorological data, and other inputs, BEPS can

simulate hourly GPP of that day. Likewise, when given a GPP value, together with other

environmental conditions, the π‘‰π‘π‘šπ‘Žπ‘₯25 value can be estimated.

2.2.3.2 Lookup-table establishment and π‘‰π‘π‘šπ‘Žπ‘₯25 searching

In order to invert an ecological model, for example, to assess the physiological conditions of plants

based on remote sensing or flux measurements, many methods have been employed, including

iterative optimization methods, neural networks, lookup-tables, etc. (Croft and Chen, 2018).

Among various methods, the lookup-table method provides a computationally efficient way to

model the physiological characteristics. In this study, the BEPS model was iteratively run using

the input data described in Chapter 2.3.1. The simulated GPP and meteorological data were

averaged to each day, and the daily GPP, daily meteorological conditions, PFT, LAI, and π‘‰π‘π‘šπ‘Žπ‘₯25

formed the lookup-table. Then, the GPP estimated from PRI at 1 km resolution was used to search

Page 37: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

26

the nearest π‘‰π‘π‘šπ‘Žπ‘₯25 value from the lookup-table as the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 . The criteria here were: 1)

whether the scatters of PRI and LUE were clustered; 2) whether the Pearson coefficients of PRI

and LUE were good enough to achieve reliable GPP estimations. However, due to the

unsatisfactory accuracy of GPP estimates from PRI, this part of work was not completed.

2.3 Discussion

2.3.1 Results and problems found in the progress

2.3.1.1 Trial of establishing generic PRI-LUE correlations for each PFT

The generic PRI-LUE correlations were established using MODIS bands 10, 12, and 13 as the

reference band. The PRI data were derived using MODIS bands 10, 12, and 13 as the reference

band (𝑃𝑅𝐼𝐡10, 𝑃𝑅𝐼𝐡12, and 𝑃𝑅𝐼𝐡13), respectively.

As shown in Figure 2-2, the scatter points are dispersed, indicating weak correlations between

either 𝑃𝑅𝐼𝐡10, 𝑃𝑅𝐼𝐡12 or 𝑃𝑅𝐼𝐡13 and LUE. Similar results can also be observed in Figure 2-3, and

Figure 2-4. The R2 values are all low though most of the correlations are significant (p<0.001), as

shown in Table 2-3.

Page 38: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

27

Figure 2-2 PRI-LUE correlations for each PFT, using band 10 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot.

Page 39: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

28

Figure 2-3 PRI-LUE correlations for each PFT, using band 12 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot.

Page 40: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

29

Figure 2-4 PRI-LUE correlations for each PFT, using band 13 as the reference band. Unit of

LUE: gC/MJ. The blue lines are regression lines of the scatter points in each subplot.

Table 2-3 Squared Pearson correlation coefficient between PRI and LUE using MODIS

bands 10, 12, and 13 at 190 sites for nine plant functional types.

Plant Functional Type 𝑃𝑅𝐼𝐡10 𝑃𝑅𝐼𝐡12 𝑃𝑅𝐼𝐡13

Page 41: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

30

ENF 0.02* 0.02* 0.01*

EBF 0.15* 0.04* 0.00*

DNF 0.06* 0.02 0.01

DBF 0.00 0.00 0.11*

MF 0.00 0.01* 0.01*

SH 0.10* 0.06* 0.07*

GRA 0.08* 0.02* 0.05*

WET 0.02* 0.02* 0.01*

CRO 0.00* 0.00 0.14*

*p<0.001

2.3.1.2 Trial of establishing PRI-LUE correlations at each site

The PRI-LUE correlations at each site also used MODIS bands 10, 12, and 13 as the reference

band. As shown in Figure 2-5, the PRI data of five sites were derived using MODIS band 10, 12,

and 13 as the reference band (𝑃𝑅𝐼𝐡10, 𝑃𝑅𝐼𝐡12, and 𝑃𝑅𝐼𝐡13), respectively.

Table 2-4 Squared Pearson correlation coefficient between sPRI and LUE using MODIS

band 10, 12, and 13 at five sites.

Site Code 𝑠𝑃𝑅𝐼𝐡10 𝑠𝑃𝑅𝐼𝐡12 𝑠𝑃𝑅𝐼𝐡13

US-Bi2 0.29* 0.26* 0.49*

US-Los 0.00 0.02 0.25*

US-Rpf 0.10* 0.13* 0.04

US-Tw4 0.12* 0.01 0.37*

US-WCr 0.11* 0.09* 0.28*

*p<0.001

Page 42: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

31

Page 43: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

32

Figure 2-5 sPRI-LUE correlations at each site using bands 10, 12, and 13. Unit of LUE:

gC/MJ. The blue lines are regression lines of the scatter points in each subplot.

Most of the correlations between sPRI and LUE are weak, as shown in Figure 2-5 and Table 2-4.

Though 𝑠𝑃𝑅𝐼𝐡13 correlates well with LUE at some sites and the 𝑠𝑃𝑅𝐼𝐡13-LUE correlations are

significant, the scatter points are dispersed and the correlations can induce bias in creating the LUE

map and GPP estimation.

In establishing PRI-LUE correlations, several problems were found:

1) Though the correlations of MODIS PRI and LUE have been tested at several sites (Drolet et al.,

2008; Goerner et al., 2009), a generic relationship between the two parameters cannot be

established. In this chapter, generic relationships of PRI-LUE for each plant functional type were

proposed to be established. However, the regressions are weak and cannot support further GPP

estimation based on the LUE derived from MODIS PRI. The locations and climate conditions of

sites were not considered. In establishing unique PRI-LUE correlations at five sites, GPP

measurements from 10 am to 12 pm at the local time were merged to derive LUE at Terra passing

time but the results were unsatisfying.

2) There are many noisy data points in the MODIS reflectance product but the filtering conditions

failed to exclude the outliers. In establishing PRI-LUE correlations, many PRI outliers exist and

exert negative impacts on the correlations. The data points in the regression are scattered widely

and the regressions are weak. However, the p values indicate that the correlations between PRI

and LUE are statistically significant or informative, indicating a certain value of PRI.

3) LUE, defined as the ratio of gross primary productivity and absorbed photosynthetically active

ration, can be influenced by temperature, moisture, and phenology (Liang et al., 2012). In

establishing PRI-LUE correlations at five sites, the air temperature was considered as an additional

factor but its consideration does not lead to improved results. The weak correlations after

considering environmental factors originate from noisy data in the MODIS reflectance product.

Page 44: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

33

2.3.2 Further work for solving the issues

1) In addition to plant functional types, the geographical locations and climate conditions will also

be considered in establishing generic PRI-LUE correlations. GPP at corresponding satellite

passing time will be merged for LUE calculation following the method by Drolet et al. (2008).

2) Some improvements will be made in preprocessing MODIS reflectance data. Instead of using

the MODIS reflectance product, MODIS at-sensor radiance data (MOD/MYD021) will be used.

Besides, the corresponding geolocation data (MOD/MYD03) and cloud mask data

(MOD/MYD35) will be used to generate geolocations and remove the contaminated data,

following the method by Zheng (2017) and Drolet et al. (2008).

3) Environmental factors including air temperature and soil moisture will be considered in

establishing PRI-LUE correlations, by adopting the factors as coefficients in the correlations.

Besides, NDVI will also be adopted to generate PRI*NDVI-LUE correlations following the

method by (Zheng and Chen, 2017).

Overall, some future works are proposed to enhance the correlations between PRI and LUE.

Reliable GPP data can only be estimated if the PRI-LUE correlations are significant and thus the

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 dataset can be derived.

Page 45: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

34

2.4 References

Camilo Rey Sanchez, Cove Sturtevant, Daphne Szutu, Dennis Baldocchi, Elke Eichelmann, Sara

Knox (2013-) AmeriFlux US-Tw4 Twitchell East End Wetland, Dataset.

https://doi.org/10.17190/AMF/1246151

Camilo Rey-Sanchez, Daphne Szutu, Dennis Baldocchi, Kyle Hemes (2017-) AmeriFlux US-Bi2

Bouldin Island corn, Dataset. https://doi.org/10.17190/AMF/1419513

Chen, J.M. (1996). Optically-based methods for measuring seasonal variation of leaf area index in

boreal conifer stands. Agricultural and Forest Meteorology, 80, 135-163

Chen, J.M., Ju, W., Ciais, P., Viovy, N., Liu, R., Liu, Y., & Lu, X. (2019). Vegetation structural

change since 1981 significantly enhanced the terrestrial carbon sink. Nature

Communication, 10, 4259

Chen, J.M., Liu, J., Cihlar, J., & Goulden, M.L. (1999). Daily canopy photosynthesis model

through temporal and spatial scaling for remote sensing applications. Ecological Modelling,

124, 99-119

Cook, B.D., Davis, K.J., Wang, W., Desai, A., Berger, B.W., Teclaw, R.M., Martin, J.G., Bolstad,

P.V., Bakwin, P.S., Yi, C., & Heilman, W. (2004). Carbon exchange and venting anomalies

in an upland deciduous forest in northern Wisconsin, USA. Agricultural and Forest

Meteorology, 126, 271-295

Croft, H., & Chen, J.M. (2018). Leaf Pigment Content. Comprehensive Remote Sensing (pp. 117-

142)

Demmig-Adams, B. (1990). Carotenoids and photoprotection in plants: A role for the xanthophyll

zeaxanthin. Biochimica et Biophysica Acta (BBA) - Bioenergetics, 1020, 1-24

Demmig-Adams, B., & Adams, W.W. (1996). Xanthophyll cycle and light stress in nature: uniform

response to excess direct sunlight among higher plant species. Planta, 198, 460-470

Drolet, G.G., Huemmrich, K.F., Hall, F.G., Middleton, E.M., Black, T.A., Barr, A.G., & Margolis,

H.A. (2005). A MODIS-derived photochemical reflectance index to detect inter-annual

variations in the photosynthetic light-use efficiency of a boreal deciduous forest. Remote

Sensing of Environment, 98, 212-224

Drolet, G.G., Middleton, E.M., Huemmrich, K.F., Hall, F.G., Amiro, B.D., Barr, A.G., Black, T.A.,

McCaughey, J.H., & Margolis, H.A. (2008). Regional mapping of gross light-use

efficiency using MODIS spectral indices. Remote Sensing of Environment, 112, 3064-3078

Farquhar, G.D., von Caemmerer, S., & Berry, J.A. (1980). A biochemical model of photosynthetic

CO2 assimilation in leaves of C3 species. Planta, 149, 78-90

Friedl, M.A., McIver, D.K., Hodges, J.C.F., Zhang, X.Y., Muchoney, D., Strahler, A.H., Woodcock,

C.E., Gopal, S., Schneider, A., Cooper, A., Baccini, A., Gao, F., & Schaaf, C. (2002). Global

Page 46: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

35

land cover mapping from MODIS: algorithms and early results. Remote Sensing of

Environment, 83, 287-302

Gamon, J.A., Field, C.B., Bilger, W., BjΓΆrkman, O., Fredeen, A.L., & PeΓ±uelas, J. (1990). Remote

sensing of the xanthophyll cycle and chlorophyll fluorescence in sunflower leaves and

canopies. Oecologia, 85, 1-7

Gamon, J.A., PeΓ±uelas, J., & Field, C.B. (1992). A narrow-waveband spectral index that tracks

diurnal changes in photosynthetic efficiency. Remote Sensing of Environment, 41, 35-44

Garbulsky, M.F., PeΓ‘Uelas, J., Papale, D., & Filella, I. (2008). Remote estimation of carbon

dioxide uptake by a Mediterranean forest. Global Change Biology, 14, 2860-2867

Goerner, A., Reichstein, M., & Rambal, S. (2009). Tracking seasonal drought effects on ecosystem

light use efficiency with satellite-based PRI in a Mediterranean forest. Remote Sensing of

Environment, 113, 1101-1111

Gonsamo, A., Chen, J.M., Price, D.T., Kurz, W.A., Liu, J., Boisvenue, C., Hember, R.A., Wu, C.,

& Chang, K.-H. (2013). Improved assessment of gross and net primary productivity of

Canada's landmass. Journal of Geophysical Research: Biogeosciences, 118, 1546-1560

Gorelick, N., Hancher, M., Dixon, M., Ilyushchenko, S., Thau, D., & Moore, R. (2017). Google

Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sensing of

Environment, 202, 18-27

Guarini, R., Nichol, C., Clement, R., Loizzo, R., Grace, J., & Borghetti, M. (2014). The utility of

MODIS-sPRI for investigating the photosynthetic light-use efficiency in a Mediterranean

deciduous forest. International Journal of Remote Sensing, 35, 6157-6172

He, L., Liu, J., Chen, J.M., Croft, H., Wang, R., Sprintsin, M., Zheng, T., Ryu, Y., Pisek, J.,

Gonsamo, A., Deng, F., & Zhang, Y. (2016). Inter- and intra-annual variations of clumping

index derived from the MODIS BRDF product. International Journal of Applied Earth

Observation and Geoinformation, 44, 53-60

Liang, S., Li, X., & Wang, J. (2012). Chapter 16 - Vegetation Production in Terrestrial Ecosystems.

In S. Liang, X. Li, & J. Wang (Eds.), Advanced Remote Sensing (pp. 501-531). Boston:

Academic Press

Liu, J., Chen, J.M., Cihlar, J., & Chen, W. (1999). Net primary productivity distribution in the

BOREAS region from a process model using satellite and surface data. Journal of

Geophysical Research: Atmospheres, 104, 27735-27754

Liu, J., Chen, J.M., Cihlar, J., & Chen, W. (2002). Net primary productivity mapped for Canada at

1-km resolution. Global Ecology and Biogeography, 11, 115-129

Liu, J., Chen, J.M., Cihlar, J., & Park, W.M. (1997). A process-based boreal ecosystem productivity

simulator using remote sensing inputs. Remote Sensing of Environment, 62, 158-175

Liu, Y. (2019). Master Thesis: Global Retrieval of the Maximum Carboxylation Rate of Plant

Page 47: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

36

Leaves by Solar-induced Chlorophyll Fluorescence from TROPOMI. In, Department of

Geography & Planning: University of Toronto

Liu, Y., Liu, R., & Chen, J.M. (2012). Retrospective retrieval of long-term consistent global leaf

area index (1981–2011) from combined AVHRR and MODIS data. Journal of Geophysical

Research: Biogeosciences, 117

Middleton, E.M., Huemmrich, K.F., Landis, D.R., Black, T.A., Barr, A.G., & McCaughey, J.H.

(2016). Photosynthetic efficiency of northern forest ecosystems using a MODIS-derived

Photochemical Reflectance Index (PRI). Remote Sensing of Environment, 187, 345-366

Monteith, J.L. (1972). Solar Radiation and Productivity in Tropical Ecosystems. Journal of

Applied Ecology, 9, 747-766

Monteith, J.L., Moss, C.J., Cooke, G.W., Pirie, N.W., & Bell, G.D.H. (1977). Climate and the

efficiency of crop production in Britain. Philosophical Transactions of the Royal Society

of London. B, Biological Sciences, 281, 277-294

Moreno, A., Maselli, F., Gilabert, M.A., Chiesi, M., MartΓ­nez, B., & Seufert, G. (2012).

Assessment of MODIS imagery to track light-use efficiency in a water-limited

Mediterranean pine forest. Remote Sensing of Environment, 123, 359-367

PeΓ‘Uelas, J., Filella, I., & Gamon, J.A. (1995). Assessment of photosynthetic radiation-use

efficiency with spectral reflectance. New Phytologist, 131, 291-296

Rahman, A.F., Cordova, V.D., Gamon, J.A., Schmid, H.P., & Sims, D.A. (2004). Potential of

MODIS ocean bands for estimating CO2 flux from terrestrial vegetation: A novel approach.

Geophysical Research Letters, 31

Sulman, B.N., Desai, A.R., Cook, B.D., Saliendra, N., & Mackay, D.S. (2009). Contrasting carbon

dioxide fluxes between a drying shrub wetland in Northern Wisconsin, USA, and nearby

forests. Biogeosciences, 6, 1115-1126

Ueyama, M., Iwata, H., Nagano, H., Tahara, N., Iwama, C., & Harazono, Y. (2019). Carbon dioxide

balance in early-successional forests after forest fires in interior Alaska. Agricultural and

Forest Meteorology, 275, 196-207

Wang, Q., Tenhunen, J., Falge, E., Bernhofer, C., Granier, A., & Vesala, T. (2004). Simulation and

scaling of temporal variation in gross primary production for coniferous and deciduous

temperate forests. Global Change Biology, 10, 37-51

Zheng, T. (2017). Retrieving Canopy-Level Light Use Efficiency (LUE) and Maximum

Carboxylation Velocity (Vcmax) Using the Photochemical Reflectance Ratio. In,

Department of Geography and Planning: University of Toronto

Zheng, T., & Chen, J.M. (2017). Photochemical reflectance ratio for tracking light use efficiency

for sunlit leaves in two forest types. ISPRS Journal of Photogrammetry and Remote

Sensing, 123, 47-61

Page 48: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

37

Chapter 3

Leaf chlorophyll content (LCC) as a feasible way for

downscaling π‘‰π‘π‘šπ‘Žπ‘₯25

3.1 Introduction

Leaf chlorophyll is essential to the cycles of carbon, water, and energy between the biosphere and

atmosphere, and to the functioning of terrestrial ecosystems (Croft et al., 2017; Croft et al., 2020).

As a crucial constituent involved in plant photosynthetic processes, chlorophyll is responsible for

harvesting photons and transporting electrons to facilitate the production of biochemical energy,

which supports the Calvin-Benson cycle (Alton, 2017; Porcar-Castell et al., 2014). Many studies

have been carried out to estimate LCC using remotely sensed data, through vegetation indices

(VIs) and biophysical modeling approaches. Leaf reflectance is controlled by the existence of

foliar constituents, including chlorophyll, carotenoids, nitrogen, and water (Ustin et al., 2004). The

spectral signals in the red-edge region (690-750 nm) are particularly sensitive to the changes in

chlorophyll content (Curran et al., 1990). Several narrowband VIs using reflectance bands in the

red-edge region have been tested to estimate chlorophyll content (Dash and Curran, 2004;

Haboudane et al., 2004; Wu et al., 2008). Besides, several studies coupled canopy models and leaf

models to retrieve chlorophyll content, and a two-step model inversion method has been

demonstrated to perform well. Zhang et al. (2008) combined the 4-Scale model (Chen and Leblanc,

1997) and the PROSPECT model (Jacquemoud and Baret, 1990) and retrieved chlorophyll content

at 20 m resolution using airborne hyperspectral data over a boreal landscape. Similar results were

also achieved by Croft et al. (2013) and Croft et al. (2015). As one of the key constituents

associated with plant photosynthetic activities, chlorophyll content can reflect vegetation growth

status and has been found to correlate well with plant photosynthetic capacity. Croft et al. (2017)

found that chlorophyll content strongly correlates with π‘‰π‘π‘šπ‘Žπ‘₯25 for four tree species in a temperate

forest and a global chlorophyll distribution map was later derived by Croft et al. (2020). Luo et al.

(2019) showed that GPP estimation at 124 tower sites (555 site-years) distributed around the globe

is significantly improved with R2 increasing from 0.74 to 0.84 after considering the seasonal

variation of π‘‰π‘π‘šπ‘Žπ‘₯25 at these sites using a global LCC map. Therefore, LCC can capture informative

vegetation traits, and the LCC at 1 km resolution can be a proxy for downscaling the π‘‰π‘π‘šπ‘Žπ‘₯25 product

at about 11 km resolution.

Page 49: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

38

Figure 3-1 Typical leaf reflectance spectra, from 400nm to 2500nm. Source: Croft and Chen

(2018).

Besides, VIs have been widely adopted to monitor vegetation greenness (Boegh et al., 2002;

Haboudane et al., 2004; Huete et al., 2002), to track light use efficiency (PeΓ‘Uelas et al., 1995),

to estimate canopy nitrogen (Serrano et al., 2002), and to retrieve canopy water content (Ceccato

et al., 2001; Jackson et al., 2004). As one of the widely used VIs, normalized difference vegetation

index (NDVI) was first used to monitor vegetation growth conditions in rangeland (Rouse et al.,

1973), calculated as:

𝑁𝐷𝑉𝐼 = π‘…π‘›π‘–π‘Ÿ βˆ’ π‘…π‘Ÿπ‘’π‘‘

π‘…π‘›π‘–π‘Ÿ + π‘…π‘Ÿπ‘’π‘‘

(3-1)

where π‘…π‘Ÿπ‘’π‘‘ and π‘…π‘›π‘–π‘Ÿ represent the spectral reflectance measurements obtained in the visible red

and near-infrared regions, respectively. NDVI has been tested to strongly correlate with green leaf

biomass, foliar nitrogen content, and maximum canopy CO2 uptake (Gamon et al., 1995). Jin et al.

(2012) correlated NDVI with π‘‰π‘π‘šπ‘Žπ‘₯ and found an exponential relationship after combining all sites

and similar conclusions were reached by Zhou et al. (2014), showing the capacity of NDVI for

capturing plant growth and photosynthetic status.

Page 50: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

39

Therefore, LCC and NDVI can provide useful information for vegetation traits at higher spatial

resolutions than the existing π‘‰π‘π‘šπ‘Žπ‘₯25 product. In this chapter, LCC and NDVI data are used to

downscale the π‘‰π‘π‘šπ‘Žπ‘₯25 product. In Section 3.3, the data preprocessing process and the spatial scaling

algorithm are introduced, using the information of LCC and NDVI, respectively. The validation

and statistical analysis methods are also introduced. In Section 3.4, the downscaled results are

presented and discussed. The conclusions are reached in Section 3.5.

3.2 Data and methods

3.2.1 Data

Five flux sites with available GPP and meteorological measurements of 2018 were selected in this

research, including US-Bi2 (Sanchez et al., 2017-), US-Los (Sulman et al., 2009), US-Rpf

(Ueyama et al., 2019), US-Tw4 (Sanchez et al., 2013-), and US-WCr (Cook et al., 2004). Data of

these five sites were collected from AMERIFLUX ((https://ameriflux.lbl.gov/). The biological and

geographical specifications of these sites are summarized in Table 3-1. The clumping indices were

derived from the product by He et al. (2016), and the LAI data of each site were derived from

GLOBMAP LAI product by Liu et al. (2012). The LAI data were resampled to a daily LAI dataset

with a 1 kmΓ—1 km resolution.

Table 3-1 Site specifications for five flux sites

Site Code Longitude Latitude Plant Functional Type Year Clumping Index Peak LAI

US-Bi2 -121.535Β° 38.109Β° CRO 2018 0.70 7.25

US-Los -89.979Β° 46.083Β° WET 2018 0.64 8.18

US-Rpf -147.429Β° 65.120Β° DBF 2018 0.66 3.92

US-Tw4 -121.641Β° 38.103Β° WET 2018 0.69 2.18

US-WCr -90.080Β° 45.806Β° DBF 2018 0.64 6.14

*CRO: croplands, WET: wetlands, DBF: deciduous broadleaf forest

Page 51: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

40

Leaf chlorophyll content data were derived from the Medium Resolution Imaging Spectrometer

(MERIS) Terrestrial Chlorophyll Index (MTCI), according to a linear equation outlined in Croft

et al. (2014):

𝐿𝐢𝐢 =𝑀𝑇𝐢𝐼 + 0.16

0.04(3-2)

where the MTCI was originally calculated as (Dash and Curran, 2004):

𝑀𝑇𝐢𝐼 = 𝑅754 βˆ’ 𝑅709

𝑅709 βˆ’ 𝑅681

(3-3)

The MTCI has demonstrated that it could use red edge positions bands to estimate chlorophyll

content over a large spatial area (Dash and Curran, 2006). However, due to lack of data from

MERIS since 2012, in this study, the Sentinel-2 data were used as a substitution. The red edge

position bands of Sentinel-2 have been evaluated to provide estimates of biophysical variables in

Figure 3-2 The correlation between MTCI and measured leaf chlorophyll content. Source:

Croft et al. (2014).

vegetation (Clevers and Gitelson, 2013; Frampton et al., 2013). Hence, the modified MTCI

(π‘€π‘‡πΆπΌπ‘š) using reference bands of Sentinel-2 was calculated as (Frampton et al., 2013):

π‘€π‘‡πΆπΌπ‘š = 𝐡6 βˆ’ 𝐡5

𝐡5 βˆ’ 𝐡4(3-4)

where B4, B5, and B6 center at 665 nm, 705 nm, and 740 nm, with a bandwidth of 30 nm, 15 nm,

and 15nm and a spatial resolution of 10 m, 20 m, and 20 m, respectively. All the images were

Page 52: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

41

collected from the Sentinel-2 Multi-Spectral Instrument Level-1C dataset (COPERNICUS/S2) and

resampled to a 1 kmΓ—1 km resolution on Google Earth Engine. Only cloud-free images were

selected for further analysis.

NDVI data were collected from the Terra Vegetation Indices 16-Day Global 500 m dataset

(MOD13A1) and resampled to a 1 kmΓ—1 km resolution on Google Earth Engine. The SummaryQA

band was used to exclude outliers.

3.2.2 Algorithms for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25

The TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 dataset was firstly gap-filled by assigning the nearest available values to the

days with no data available. The TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 dataset was assumed to be numerically reliable

and the downscaling factors only provided spatial variation information for distributing the

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 data within each 0.1°×0.1Β° pixel. The downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 values were then

obtained by applying a spatial scaling ratio 𝑆𝑆𝑅 to TROPOMI Vcmax, as described in the

following equation:

π‘‰π‘π‘šπ‘Žπ‘₯βˆ’π‘˜π‘š = 𝑆𝑆𝑅 Γ— π‘‰π‘π‘šπ‘Žπ‘₯βˆ’π‘‘π‘” (3-5)

where π‘‰π‘π‘šπ‘Žπ‘₯βˆ’π‘˜π‘š, π‘‰π‘π‘šπ‘Žπ‘₯βˆ’π‘‘π‘” represent the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 with a 1 kmΓ—1 km resolution, and

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 with a 0.1°×0.1Β° resolution, respectively.

The spatial scaling ratios of each pixel were derived as described in the following equations:

𝑆𝑆𝑅𝐿𝐢𝐢𝑖,𝑗=

𝐿𝐢𝐢𝑖,𝑗

𝐿𝐢𝐢̅̅ Μ…Μ… Μ…(3-6)

𝑆𝑆𝑅𝑁𝐷𝑉𝐼𝑖,𝑗=

𝑁𝐷𝑉𝐼𝑖,𝑗

𝑁𝐷𝑉𝐼̅̅ Μ…Μ… Μ…Μ… Μ…Μ…(3-7)

where 𝐿𝐢𝐢𝑖,𝑗 and 𝑁𝐷𝑉𝐼𝑖,𝑗 represent LCC and NDVI of that 1 kmΓ—1k m pixel; 𝐿𝐢𝐢̅̅ Μ…Μ… Μ… and 𝑁𝐷𝑉𝐼̅̅ Μ…Μ… Μ…Μ… Μ…Μ…

represent the mean LCC and NDVI of that 0.1°×0.1Β° pixel, respectively. When calculating the

mean LCC and the mean NDVI, outliers were replaced with the nearest available data in the

0.1°×0.1Β° pixel. The mean values of all downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values in 0.1°×0.1Β° pixels were kept

the same as the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 data. The downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 data for each site in this study using

Page 53: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

42

two methods mentioned above were obtained, after gap-filling the days with missing data using

available π‘‰π‘π‘šπ‘Žπ‘₯25 values of the nearest day.

3.2.3 Evaluation and Statistical Analysis

The Boreal Ecosystem Productivity Simulator (BEPS) was adopted in this study to evaluate the

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 , and it is described in Chapter 2.3.3.1. The original π‘‰π‘π‘šπ‘Žπ‘₯

25 data and downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 data from LCC and NDVI were input to BEPS for GPP simulation. To evaluate the

performance of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 , simulated GPP based on original TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 data,

NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data, and LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 data were compared with eddy

covariance (EC) flux GPP measurements, where squared Pearson correlation coefficient (R2),

mean absolute error (MAE) and root mean square error (RMSE) were used. The performance of

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 was also evaluated using the agreement index ( 𝐼𝑂𝐴 ), which was firstly

proposed by Willmott (1981, 1982) and has been widely adopted in model assessments (Gu et al.,

2002; Wang et al., 2001; Zhou et al., 2016). This index of agreement represents the ratio of the

mean square error and the potential error, taking both R2 and RMSE into consideration. The index

𝐼𝑂𝐴 is calculated as:

𝐼𝑂𝐴 = 1 βˆ’βˆ‘ (π‘ƒπ‘˜ βˆ’ π‘‚π‘˜)2𝑛

π‘˜=1

βˆ‘ (|π‘ƒβ€²π‘˜| βˆ’ |𝑂′

π‘˜|)2π‘›π‘˜=1

(3-8)

where π‘ƒβ€²π‘˜ = π‘ƒπ‘˜ βˆ’ οΏ½Μ…οΏ½ and π‘‚β€²π‘˜ = π‘‚π‘˜ βˆ’ οΏ½Μ…οΏ½ ; 𝑛 is the total number of measurements; π‘ƒπ‘˜ and π‘‚π‘˜

represent GPP simulated from BEPS and EC GPP measurements, respectively; οΏ½Μ…οΏ½ is the mean

value of EC GPP. The value of 𝐼𝑂𝐴 ranges from 𝐼𝑂𝐴 = 0, for no agreement between π‘ƒπ‘˜ and π‘‚π‘˜,

to 𝐼𝑂𝐴 = 1, for perfect agreement between simulations and observations.

Based on the downscaling factor with higher GPP estimation accuracy, a global 1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map

was produced. To evaluate the spatial distribution and variation within the TROPOMI pixels, the

standard deviation (𝑆𝐷0.1𝐷) of each TROPOMI pixel was calculated and the standard deviations

of all global pixels (𝑆𝐷) were averaged, as described in the following equations:

𝑆𝐷0.1𝐷 = βˆšβˆ‘(π‘‰π‘π‘šπ‘Žπ‘₯𝑖

25 βˆ’ π‘‰π‘π‘šπ‘Žπ‘₯25Μ…Μ… Μ…Μ… Μ…Μ… Μ…)

𝑛1π‘˜π‘š

(3-9)

Page 54: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

43

𝑆𝐷 = βˆ‘π‘†π·0.1𝐷𝑖

𝑛0.1𝐷

(3-10)

where π‘‰π‘π‘šπ‘Žπ‘₯𝑖25 , π‘‰π‘π‘šπ‘Žπ‘₯

25Μ…Μ… Μ…Μ… Μ…Μ… Μ…, 𝑛1π‘˜π‘š , and 𝑛0.1𝐷 stand for the 1 km π‘‰π‘π‘šπ‘Žπ‘₯25 values, the average of 1 km

π‘‰π‘π‘šπ‘Žπ‘₯25 values within the TROPOMI pixel, the total number of 1 km π‘‰π‘π‘šπ‘Žπ‘₯

25 pixels within the

TROPOMI pixel, and the total number of global TROPOMI pixels, respectively.

3.3 Results and discussion

Due to the availability of the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset, the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 starts from March 7th, 2018

(the day of year 66) to October 31st, 2018 (the day of year 304). The following analysis and results

are all in the range of the day of year 100 to 300, from the beginning to the end of the growing

season.

3.3.1 The intra-pixel heterogeneity of the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 product

As shown in Figure 3-3, significant intra-pixel heterogeneities exist in the 0.1°×0.1Β° π‘‰π‘π‘šπ‘Žπ‘₯25 pixels

on the day of year 190, in the middle of the growing season. The red points in the figures are the

relative location of the sites in the TROPOMI pixel. The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values were retrieved

by applying the scaling ratio derived from LCC, 𝑆𝑆𝑅𝐿𝐢𝐢, of that day. The blue, green and yellow

pixels represent low, moderate and high π‘‰π‘π‘šπ‘Žπ‘₯25 values of that downscaled pixel. The vertical and

horizontal axes indicate the relative location of each downscaled pixel in the TROPOMI pixel.

US-Bi2 is a crop site, with growing corn on an island in the Sacramento San Joaquin Delta. The

site has deep peat, thus providing ideal environments for farming and the π‘‰π‘π‘šπ‘Žπ‘₯25 values are high in

that region. The site is located at the bottom of that TROPOMI pixel and there are some high

estimates of π‘‰π‘π‘šπ‘Žπ‘₯25 in the middle right of that TROPOMI pixel. The main landcover in that

TROPOMI pixel is crop, but the growing status of various subpixels differs, leading to

heterogeneous downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values. US-Los is a shrub wetland site with some coniferous and

grassy stands. It is located in the top left of that TROPOMI pixel and the landcover within that

TROPOMI pixel is intensively heterogeneous. Different types of plants have individual growing

patterns and physiological characteristics, which indicates that the variation of π‘‰π‘π‘šπ‘Žπ‘₯25 differs

among plant functional types. The middle right downscaled pixels have lower π‘‰π‘π‘šπ‘Žπ‘₯25 while the top

right downscaled pixels have higher values. US-Rpf is a deciduous broadleaf forest in Alaska. It

Page 55: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

44

is located in the middle right of that TROPOMI pixel and there are lower downscaled π‘‰π‘π‘šπ‘Žπ‘₯25

estimates in the left part of that TROPOMI pixel. US-Tw4 is a permanent wetland site with woody

vegetation. It is located at the bottom of that TROPOMI pixel and intensive variation of π‘‰π‘π‘šπ‘Žπ‘₯25 can

be seen from the figure of US-Tw4. Lower values are situated in the left part and higher values

situate in the bottom right corner. US-WCr is a deciduous broadleaf forest site in Wisconsin,

mainly with sugar maple and basswood. It is located in the bottom-left part of that TROPOMI

pixel and lower downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values cluster around the bottom right corner.

Page 56: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

45

Figure 3-3 Intra-pixel heterogeneous π‘½π’„π’Žπ’‚π’™πŸπŸ“ distribution within the TROPOMI pixel of each

site on the closest cloud-free day to the day of year 190. The π‘½π’„π’Žπ’‚π’™πŸπŸ“ values at 1 kmΓ—1 km are

downscaled through LCC. The red point represents the relative location of the site (1 kmΓ—1

km) in the TROPOMI pixel (0.1°×0.1Β°).

Page 57: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

46

Figure 3-4 Intra-pixel heterogeneous π‘½π’„π’Žπ’‚π’™πŸπŸ“ distribution within the TROPOMI pixel of

each site on the closest cloud-free day to the day of year 190. The π‘½π’„π’Žπ’‚π’™πŸπŸ“ values at 1 kmΓ—1

km are downscaled through NDVI. The red point represents the relative location of the site

(1 kmΓ—1 km) in the TROPOMI pixel (0.1°×0.1Β°).

Page 58: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

47

3.3.2 The π‘‰π‘π‘šπ‘Žπ‘₯25 seasonal variation

With the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 using 𝑆𝑆𝑅𝐿𝐢𝐢 and 𝑆𝑆𝑅𝑁𝐷𝑉𝐼 , the temporal patterns of π‘‰π‘π‘šπ‘Žπ‘₯

25 at each site

can be generalized. At some sites, the seasonal variation of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 shows similar

patterns to the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 product, while at other sites, the patterns differ a lot.

As shown in Figure 3-5, in US-Bi2 and US-WCr, the LCC-downscaled and NDVI-downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 curves intersect at two points with the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 curve. Interestingly, the

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values at sites are higher than the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 values in the interval

between the two intercepts, which is in the middle of the growing season. The LCC-downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 values at sites are higher than the NDVI-downscaled values at sites between the intercepts.

Beyond the two intercepts, i.e., before and after the growing season, the opposite situation occurs.

The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values are lower than the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 values. In US-Bi2, the peak

π‘‰π‘π‘šπ‘Žπ‘₯25 values are very high mainly because the soil fertility and the values are low before sowing

in spring and harvesting in autumn. In US-Los, the LCC downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values are all lower

than the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 values and the seasonal variation of the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 is

similar to that of the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 , indicating the 𝑆𝑆𝑅𝐿𝐢𝐢 of this site undulates slightly near

1. The downscaled NDVI π‘‰π‘π‘šπ‘Žπ‘₯25 curve is found to fluctuate moderately near the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25

curve. In US-Rpf, over the period of the day of year 180 to 220, a slight decrease occurs to the

LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values while the trends of the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 and the TROPOMI

π‘‰π‘π‘šπ‘Žπ‘₯25 become consistent afterwards. In US-Tw4, a steady upward trend is observed in all the

π‘‰π‘π‘šπ‘Žπ‘₯25 curves, in which the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 curve stays near the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 curve

but the NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 curve diverges drastically.

Page 59: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

48

Figure 3-5 π‘½π’„π’Žπ’‚π’™πŸπŸ“ seasonal variation of each site from the day of year 100 to 300.

Page 60: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

49

3.3.3 Evaluation of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 at sites

Due to the lack of available π‘‰π‘π‘šπ‘Žπ‘₯25 measurements for validation, the downscaled results were

evaluated indirectly through terrestrial biosphere models. The original and downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data

were adopted as the input parameter into the BEPS model for GPP simulation.

3.3.3.1 The comparison of GPP simulation results

In Figure 3-6, the vertical and horizontal axes represent the GPP simulation and EC GPP

measurements. The blue, orange, and green points indicate the GPP simulation results with

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 , LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 , and NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 , and the blue, orange,

and green lines indicate the linear regression of the scatter points, respectively. It can be observed

that at all sites, the regression lines of LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 scatter points are most aligned with

the 1:1 line and similar results are also exhibited after combining points from all 5 sites, as shown

in Figure 3-7. However, some of the regression lines of NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 scatter points

are more aligned with the 1:1 line compared to the original TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 ones while some are

less aligned. This is because NDVI catches up to the greenness of plants but the physiological

connection between NDVI and photosynthetic capacity is weak.

Page 61: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

50

Page 62: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

51

Figure 3-6 GPP scatter plot of each site from the day of year 100 to 300. The Y-axis represents

GPP estimates and the X-axis represents flux measured GPP.

Figure 3-7 GPP scatter plot of all sites from the day of year 100 to 300.

3.3.3.2 The seasonal variation of GPP simulation results

In Figure 3-8, the vertical and horizontal axes represent the daily GPP simulation or GPP

measurements and day of year. The red, orange, blue, and green curves indicate the GPP seasonal

variation of EC measurements, simulation with TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 , simulation with LCC-

Page 63: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

52

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 , and simulation with NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 . Overall, the GPP simulation

results with LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 show most aligned patterns to the EC GPP measurements. In

Figure 3-8 GPP seasonal variation of each site from the day of year 100 to 300. The daily EC

GPP data were merged from half-hourly EC measurements. The daily modeled GPP data

were merged from hourly BEPS GPP simulation results.

Page 64: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

53

Figure 3-9 Temporal patterns of the bias of GPP estimated with TROPOMI and downscaled

π‘½π’„π’Žπ’‚π’™πŸπŸ“

US-Bi2, LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data considerably improve the GPP simulation between the day

of year 150 to the day of year 250 and NDVI-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data also improve the result to

Page 65: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

54

some extent. The underestimates of GPP simulation during the growing season originate from a

lack of consideration of fertilization and irrigation. In US-Los and US-Rpf, significant

overestimates of simulated GPP are observed before the day of year 150, though GPP simulation

results with LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 are closer to the EC measurements. There are overestimates

of GPP simulation in US-Tw4 and the estimates with LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 show a gradual

decline from the estimates with TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 . In US-WCr, GPP estimates with LCC-

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 capture the fluctuation of daily variation best. As shown in Figure 3-9, the

temporal patterns of the bias of GPP estimation with TROPOMI and downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 are

presented. Due to the consistency of other parameters except π‘‰π‘π‘šπ‘Žπ‘₯25 in BEPS modeling, the

temporal patterns of the bias are similar. However, the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 reduces the bias of

GPP estimates at all sites.

3.3.3.3 Statistical analysis of GPP simulation results

In Table 3-2, the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 data and the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 data are used as input of BEPS

to simulate GPP. The simulated GPP data are compared with EC GPP measurements. Squared

Pearson correlation coefficient (R2), mean absolute error (MAE), root mean square error (RMSE),

and the agreement index (𝐼𝑂𝐴) are calculated for statistical analysis. The results of each site and

combined-all-sites are presented at all sites the GPP simulation results with LCC-downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 are observed to achieve the least RMSE and MAE, and highest IOA. After combining points

from all sites, LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data remarkably improve the simulation results from R2 =

0.80 to R2 = 0.85 and IOA = 0.93 to IOA = 0.95. The RMSE and MAE decrease from RMSE =

2.49 to RMSE = 2.15 gC/m2/d and MAE = 1.74 to MAE = 1.55 gC/m2/d. Although NDVI-

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data improve the IOA and RMSE slightly, the same R2 and worse MAE indicate

the limits of NDVI for downscaling. However, the overall improvement of GPP estimates with

LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 is not large because the improvements are small at four sites except US-

Bi2. Only US-Bi2 has obvious differences between the downscaled and TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 values

while the other four sites have the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values close to the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 values,

as shown in Figure 3-5. These small improvements could be mostly because the landscapes over

the available tower flux sites are relatively homogeneous so that the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values do

not differ greatly from original values before downscaling. Besides, LAI and CI can affect the

magnitude of simulated GPP (Chen et al., 2012; Liu et al., 2018). Errors in the input data such as

Page 66: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

55

LAI and CI to BEPS used to evaluate π‘‰π‘π‘šπ‘Žπ‘₯25 downscaling through simulated GPP would also have

considerable influence on the accuracy of simulated GPP and hence on the π‘‰π‘π‘šπ‘Žπ‘₯25 evaluation.

Table 3-2 Statistical evaluation results of the downscaled π‘½π’„π’Žπ’‚π’™πŸπŸ“ based on GPP simulated

with π‘½π’„π’Žπ’‚π’™πŸπŸ“ before and after downscaling against GPP derived from tower flux

measurements

Site Code π‘‰π‘π‘šπ‘Žπ‘₯25 Source R2 RMSE MAE IOA

US-Bi2

TROPOMI 0.91 3.69 2.37 0.92

LCC-Downscaled 0.92 2.95 1.94 0.95

NDVI-Downscaled 0.91 3.25 2.09 0.94

US-Los

TROPOMI 0.84 1.75 1.32 0.94

LCC-Downscaled 0.87 1.47 1.10 0.95

NDVI-Downscaled 0.85 1.81 1.37 0.93

US-Rpf

TROPOMI 0.84 1.86 1.43 0.94

LCC-Downscaled 0.82 1.80 1.38 0.94

NDVI-Downscaled 0.81 1.92 1.47 0.94

US-Tw4

TROPOMI 0.81 1.80 1.43 0.89

LCC-Downscaled 0.82 1.59 1.37 0.93

NDVI-Downscaled 0.80 2.47 2.18 0.84

US-WCr

TROPOMI 0.79 2.75 2.15 0.92

LCC-Downscaled 0.81 2.53 1.96 0.94

NDVI-Downscaled 0.79 2.64 2.04 0.93

All Sites

TROPOMI 0.80 2.49 1.74 0.93

LCC-Downscaled 0.85 2.15 1.55 0.95

NDVI-Downscaled 0.80 2.47 1.83 0.94

*Unit of RMSE and MAE: gC/m2/d

Page 67: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

56

3.3.3.4 GPP responses to π‘‰π‘π‘šπ‘Žπ‘₯25 before and after downscaling

The percentage change of GPP estimates before and after downscaling is related to the standard

deviation of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 in 0.1˚ pixels, LAI, and the landcover. GPP was estimated at each

1 km pixel using LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 and then averaged to 0.1˚ pixels, named as the distributed

GPP. GPP estimated using TROPOMI-π‘‰π‘π‘šπ‘Žπ‘₯25 data, i.e., one value for each 0.1˚ pixel, named as the

lumped GPP, was compared with the distributed GPP and thus the percentage change was

obtained. The standard deviation of LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 was calculated at each 0.1˚ pixel. At

all 0.1˚ pixels of sites, GPP estimations at 0.1˚ resolution decrease by 2-7% after π‘‰π‘π‘šπ‘Žπ‘₯25

downscaling using the LCC method, as shown in Table 3-3.

Table 3-3 The summary of GPP responses to π‘½π’„π’Žπ’‚π’™πŸπŸ“ before and after downscaling

Site Code π‘‰π‘π‘šπ‘Žπ‘₯25 𝑆𝐷

π‘‰π‘π‘šπ‘Žπ‘₯25 𝑆𝐷

π‘‰π‘π‘šπ‘Žπ‘₯25Μ…Μ… Μ…Μ… Μ…Μ… Μ…Μ…

Percentage

Change Mean LAI

US-Bi2 37.39 0.32 -2.14 2.29

US-Los 6.54 0.21 -2.03 4.09

US-Rpf 10.76 0.30 -2.16 1.76

US-Tw4 76.29 0.65 -5.33 1.47

US-WCr 8.12 0.23 -7.21 3.00

*SD: Standard Deviation, Unit of π‘‰π‘π‘šπ‘Žπ‘₯25 : Β΅mol/m2/s

The distributed GPP estimates were lower than the lumped GPP at all sites because of the nonlinear

relationship between GPP and π‘‰π‘π‘šπ‘Žπ‘₯25 . As shown in Figure 3-10, the blue curve represents GPP

responses to changes of π‘‰π‘π‘šπ‘Žπ‘₯25 in the BEPS model. The black line is the straightened version of

the blue curve between π‘‰π‘π‘šπ‘Žπ‘₯25 – SD and π‘‰π‘π‘šπ‘Žπ‘₯

25 + SD. The black straight line provides quick

estimations of distributed GPP because the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 is most located in that range. The

GPP estimates were retrieved using environmental data on the day of year 124 at US-WCr site

with π‘‰π‘π‘šπ‘Žπ‘₯25 ranging from 0 to 80 Β΅mol/m2/s. As π‘‰π‘π‘šπ‘Žπ‘₯

25 gets higher, GPP estimates grow slower.

The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 within the 0.1˚ pixel fluctuates in a certain range with a mean value equal

to the π‘‰π‘π‘šπ‘Žπ‘₯25 value of the 0.1˚ pixel. However, the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 lower than the mean value

Page 68: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

57

exerts a greater influence on the change of GPP than the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 higher than the mean

value. The red point is the middle point of the black straight line, indicating the mean value of

distributed GPP. Therefore, a decrement of GPP estimates can be observed after downscaling and

thus the overestimation of lumped GPP can be corrected.

Figure 3-10 GPP responses to π‘½π’„π’Žπ’‚π’™πŸπŸ“ in the BEPS model. The Y-axis represents the GPP

estimates. Other inputs are kept consistent, using data on the day of year 124 at US-WCr

site. The X-axis represents π‘½π’„π’Žπ’‚π’™πŸπŸ“ values changing from 0 to 80 Β΅mol/m2/s. The blue curve

shows the GPP responses to change of π‘½π’„π’Žπ’‚π’™πŸπŸ“ . The black line is the straightened version of

the curve between mean π‘½π’„π’Žπ’‚π’™πŸπŸ“ – SD and π‘½π’„π’Žπ’‚π’™

πŸπŸ“ + SD. The blue point represents lumped

GPP, simulated using π‘½π’„π’Žπ’‚π’™πŸπŸ“ at the 0.1˚ resolution. The red point represents distributed, the

mean of GPP simulated based on the LCC-downscaled π‘½π’„π’Žπ’‚π’™πŸπŸ“ within the 0.1˚ pixel.

Page 69: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

58

3.3.4 Applying the downscaling method to regional and global scales

As discussed in Chapter 3.4.3, the LCC downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 achieves better performance in GPP

simulation, indicating LCC as a feasible way to downscale the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset. Therefore, the LCC-

downscaling method was further applied from site to regional and global scales. For the global

map, as shown in Figure 3-12, the distributions of π‘‰π‘π‘šπ‘Žπ‘₯25 values in the original and the downscaled

π‘‰π‘π‘šπ‘Žπ‘₯25 map show similar patterns because the downscaling considers the spatial distribution of

LCC within each TROPOMI pixel while the average π‘‰π‘π‘šπ‘Žπ‘₯25 for the pixel is unchanged. . However,

the global averaged intra-pixel standard deviation is 6.72 Β΅mol/m2/s, showing the heterogeneity in

TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯25 pixels and the spatial variation of the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 map, as shown in

Figure 3-11. For the regional maps of North America, as shown in Figure 3-13, Figure 3-14, and

Figure 3-15, details can be observed in the enlarged portions of regional maps. For example, the

spatial distribution and variation of π‘‰π‘π‘šπ‘Žπ‘₯25 can be clearly captured in the subfigure f) in each figure.

The seasonal variation patterns of π‘‰π‘π‘šπ‘Žπ‘₯25 are also shown in the comparison of the three figures. The

downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 maps preserve the values of the coarse-resolution π‘‰π‘π‘šπ‘Žπ‘₯

25 dataset and

simultaneously contain the within-pixel spatial information of LCC.

Figure 3-11 Spatial distribution of standard deviation of downscaled π‘½π’„π’Žπ’‚π’™πŸπŸ“ values in

TROPOMI π‘½π’„π’Žπ’‚π’™πŸπŸ“ pixels on the day of year 200.

Page 70: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

59

Figure 3-12 Comparison of global π‘½π’„π’Žπ’‚π’™πŸπŸ“ maps on the day of year 200. a) The 0.1°×0.1Β°

TROPOMI π‘½π’„π’Žπ’‚π’™πŸπŸ“ map; b) The downscaled 1 kmΓ—1 km π‘½π’„π’Žπ’‚π’™

πŸπŸ“ map.

Page 71: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

60

Figure 3-13 Comparison of π‘½π’„π’Žπ’‚π’™πŸπŸ“ maps of North America on the day of year 150. a) The 0.1°×0.1Β° TROPOMI π‘½π’„π’Žπ’‚π’™

πŸπŸ“ map; b)

The downscaled 1 kmΓ—1 km π‘½π’„π’Žπ’‚π’™πŸπŸ“ map; c) and d) Partially enlarged details of the red labeled region in a) and b); e) and f)

Partially enlarged details of the red labeled region in c) and d).

Page 72: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

61

Figure 3-14 Comparison of π‘½π’„π’Žπ’‚π’™πŸπŸ“ maps of North America on the day of year 200. a) The 0.1°×0.1Β° TROPOMI π‘½π’„π’Žπ’‚π’™

πŸπŸ“ map; b)

The downscaled 1 kmΓ—1 km π‘½π’„π’Žπ’‚π’™πŸπŸ“ map; c) and d) Partially enlarged details of the red labeled region in a) and b); e) and f)

Partially enlarged details of the red labeled region in c) and d).

Page 73: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

62

Figure 3-15 Comparison of π‘½π’„π’Žπ’‚π’™πŸπŸ“ maps of North America on the day of year 250. a) The 0.1°×0.1Β° TROPOMI π‘½π’„π’Žπ’‚π’™

πŸπŸ“ map; b)

The downscaled 1 kmΓ—1 km π‘½π’„π’Žπ’‚π’™πŸπŸ“ map; c) and d) Partially enlarged details of the red labeled region in a) and b); e) and f)

Partially enlarged details of the red labeled region in c) and d).

Page 74: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

63

3.4 Conclusion

Due to the nature of the heterogeneous land surface, the intra-pixel heterogeneity and the scaling

effect have been widely discussed in previous studies (Chen, 1999; Chen et al., 2013; Garrigues et

al., 2006). However, the heterogeneity in π‘‰π‘π‘šπ‘Žπ‘₯25 derived from remotely sensed data has not yet

been studied. In this chapter, the intra-pixel heterogeneities of a π‘‰π‘π‘šπ‘Žπ‘₯25 dataset derived from sun-

induced chlorophyll fluorescence by (Liu, 2019) are quantitatively investigated. A spatial scaling

method to downscale the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset to a resolution appropriate for comparison with flux

measurements at eddy covariance towers was developed. The spatial scaling ratio (SSR) is

calculated as the proportion of the informative factor in the 1 km pixel and the averaged

informative factor in the TROPOMI pixel. Two separate downscaling factors, leaf chlorophyll

content (LCC) and normalized difference vegetation index (NDVI) were used to provide useful

information at a higher spatial resolution for downscaling. The seasonal variation patterns and

intra-pixel heterogeneities of π‘‰π‘π‘šπ‘Žπ‘₯25 were presented and analyzed. The downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 with the

two factors mentioned above were evaluated by adopting the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 into BEPS for

GPP simulation and by validating the GPP simulation results with GPP derived from EC

measurements. The other parameters of BEPS except π‘‰π‘π‘šπ‘Žπ‘₯25 were kept consistent for the evaluation

of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 . The bias of GPP estimates, temporal patterns of GPP, and statistical analysis

of simulated GPP were presented. Based on the SSR with better GPP estimates, the first global 1

km π‘‰π‘π‘šπ‘Žπ‘₯25 map was produced.

The main conclusions of this chapter are: 1) Intra-pixel heterogeneities of the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset

retrieved from the TROPOMI SIF data at 0.1-degree resolution were significant at selected 5

locations. The seasonal variation pattern of downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 differed considerably from those at

the original coarse resolution at some sites. When adopting the coarse π‘‰π‘π‘šπ‘Žπ‘₯25 data into terrestrial

ecosystem models, the intra-pixel heterogeneity will induce bias in simulation results over tower

flux sites with footprints much smaller than the pixel size. 2) A spatial scaling algorithm was

developed and LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 achieve the best results in terms of matching simulated

GPP with tower-measured GPP. The downscaling processes were implemented at five sites with

available GPP measurements in 2018. The spatial scaling ratios derived from LCC and NDVI were

experimented and the LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 improved the GPP simulation most. LCC has been

tested as the proxy of the photosynthetic capacity (Croft et al., 2017). Thus LCC is more

Page 75: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

64

physiologically related to plant photosynthetic process than NDVI, indicating that a factor with

tighter connections with plant photosynthetic capacity can better support π‘‰π‘π‘šπ‘Žπ‘₯25 downscaling. On

the contrary, if the informative factor for downscaling is not physiologically connected with plant

photosynthesis, the downscaling will cause error propagation, leading to unreliable results. 3) The

first 1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map was produced following the LCC downscaling method. Within-pixel details

of π‘‰π‘π‘šπ‘Žπ‘₯25 can be observed after the downscaling.

Overall, intra-pixel heterogeneities indeed exist within the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset derived from TROPOMI

SIF measurements. The spatial scaling algorithm is reliable and LCC can provide a feasible method

for downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 datasets and studying the intra-pixel spatial distribution pattern of the

π‘‰π‘π‘šπ‘Žπ‘₯25 dataset. Since LCC datasets are not widely available, NDVI can also be used for

downscaling π‘‰π‘π‘šπ‘Žπ‘₯25 as the first-order approximation, but because NDVI is a surrogate of canopy

structure (e.g., LAI) and leaf greenness, it is not as ideal as LCC, which represents a leaf trait that

is closely linked to leaf photosynthetic capacity.

Page 76: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

65

3.5 References

Alton, P.B. (2017). Retrieval of seasonal Rubisco-limited photosynthetic capacity at global

FLUXNET sites from hyperspectral satellite remote sensing: Impact on carbon modeling.

Agricultural and Forest Meteorology, 232, 74-88

Boegh, E., Soegaard, H., Broge, N., Hasager, C.B., Jensen, N.O., Schelde, K., & Thomsen, A.

(2002). Airborne multispectral data for quantifying leaf area index, nitrogen concentration,

and photosynthetic efficiency in agriculture. Remote Sensing of Environment, 81, 179-193

Camilo Rey Sanchez, Cove Sturtevant, Daphne Szutu, Dennis Baldocchi, Elke Eichelmann, Sara

Knox (2013-) AmeriFlux US-Tw4 Twitchell East End Wetland, Dataset.

https://doi.org/10.17190/AMF/1246151

Camilo Rey-Sanchez, Daphne Szutu, Dennis Baldocchi, Kyle Hemes (2017-) AmeriFlux US-Bi2

Bouldin Island corn, Dataset. https://doi.org/10.17190/AMF/1419513

Ceccato, P., Flasse, S., Tarantola, S., Jacquemoud, S., & GrΓ©goire, J.-M. (2001). Detecting

vegetation leaf water content using reflectance in the optical domain. Remote Sensing of

Environment, 77, 22-33

Chen, J.M. (1999). Spatial Scaling of a Remotely Sensed Surface Parameter by Contexture.

Remote Sensing of Environment, 69, 30-42

Chen, J.M., Chen, X., & Ju, W. (2013). Effects of vegetation heterogeneity and surface topography

on spatial scaling of net primary productivity. Biogeosciences, 10, 4879-4896

Chen, J.M., & Leblanc, S.G. (1997). A four-scale bidirectional reflectance model based on canopy

architecture. IEEE Transactions on Geoscience and Remote Sensing, 35, 1316-1337

Chen, J.M., Mo, G., Pisek, J., Liu, J., Deng, F., Ishizawa, M., & Chan, D. (2012). Effects of foliage

clumping on the estimation of global terrestrial gross primary productivity. Global

Biogeochemical Cycles, 26

Clevers, J.G.P.W., & Gitelson, A.A. (2013). Remote estimation of crop and grass chlorophyll and

nitrogen content using red-edge bands on Sentinel-2 and -3. International Journal of

Applied Earth Observation and Geoinformation, 23, 344-351

Cook, B.D., Davis, K.J., Wang, W., Desai, A., Berger, B.W., Teclaw, R.M., Martin, J.G., Bolstad,

P.V., Bakwin, P.S., Yi, C., & Heilman, W. (2004). Carbon exchange and venting anomalies

in an upland deciduous forest in northern Wisconsin, USA. Agricultural and Forest

Meteorology, 126, 271-295

Croft, H., & Chen, J.M. (2018). Leaf Pigment Content. Comprehensive Remote Sensing (pp. 117-

142)

Croft, H., Chen, J.M., Luo, X., Bartlett, P., Chen, B., & Staebler, R.M. (2017). Leaf chlorophyll

content as a proxy for leaf photosynthetic capacity. Global Change Biology, 23, 3513-3524

Page 77: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

66

Croft, H., Chen, J.M., Wang, R., Mo, G., Luo, S., Luo, X., He, L., Gonsamo, A., Arabian, J., Zhang,

Y., Simic-Milas, A., Noland, T.L., He, Y., HomolovΓ‘, L., MalenovskΓ½, Z., Yi, Q., Beringer,

J., Amiri, R., Hutley, L., Arellano, P., Stahl, C., & Bonal, D. (2020). The global distribution

of leaf chlorophyll content. Remote Sensing of Environment, 236, 111479

Croft, H., Chen, J.M., & Zhang, Y. (2014). The applicability of empirical vegetation indices for

determining leaf chlorophyll content over different leaf and canopy structures. Ecological

Complexity, 17, 119-130

Croft, H., Chen, J.M., Zhang, Y., & Simic, A. (2013). Modelling leaf chlorophyll content in

broadleaf and needle leaf canopies from ground, CASI, Landsat TM 5 and MERIS

reflectance data. Remote Sensing of Environment, 133, 128-140

Croft, H., Chen, J.M., Zhang, Y., Simic, A., Noland, T.L., Nesbitt, N., & Arabian, J. (2015).

Evaluating leaf chlorophyll content prediction from multispectral remote sensing data

within a physically-based modelling framework. ISPRS Journal of Photogrammetry and

Remote Sensing, 102, 85-95

Curran, P.J., Dungan, J.L., & Gholz, H.L. (1990). Exploring the relationship between reflectance

red edge and chlorophyll content in slash pine. Tree Physiology, 7, 33-48

Dash, J., & Curran, P.J. (2004). The MERIS terrestrial chlorophyll index. International Journal of

Remote Sensing, 25, 5403-5413

Dash, J., & Curran, P.J. (2006). Relationship between herbicide concentration during the 1960s

and 1970s and the contemporary MERIS Terrestrial Chlorophyll Index (MTCI) for

southern Vietnam. International Journal of Geographical Information Science, 20, 929-

939

Frampton, W.J., Dash, J., Watmough, G., & Milton, E.J. (2013). Evaluating the capabilities of

Sentinel-2 for quantitative estimation of biophysical variables in vegetation. ISPRS Journal

of Photogrammetry and Remote Sensing, 82, 83-92

Gamon, J.A., Field, C.B., Goulden, M.L., Griffin, K.L., Hartley, A.E., Joel, G., Penuelas, J., &

Valentini, R. (1995). Relationships Between NDVI, Canopy Structure, and Photosynthesis

in Three Californian Vegetation Types. Ecological Applications, 5, 28-41

Garrigues, S., Allard, D., Baret, F., & Weiss, M. (2006). Quantifying spatial heterogeneity at the

landscape scale using variogram models. Remote Sensing of Environment, 103, 81-96

Gu, L., Baldocchi, D., Verma, S.B., Black, T.A., Vesala, T., Falge, E.M., & Dowty, P.R. (2002).

Advantages of diffuse radiation for terrestrial ecosystem productivity. Journal of

Geophysical Research: Atmospheres, 107, ACL 2-1-ACL 2-23

Haboudane, D., Miller, J.R., Pattey, E., Zarco-Tejada, P.J., & Strachan, I.B. (2004). Hyperspectral

vegetation indices and novel algorithms for predicting green LAI of crop canopies:

Modeling and validation in the context of precision agriculture. Remote Sensing of

Environment, 90, 337-352

Page 78: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

67

He, L., Liu, J., Chen, J.M., Croft, H., Wang, R., Sprintsin, M., Zheng, T., Ryu, Y., Pisek, J.,

Gonsamo, A., Deng, F., & Zhang, Y. (2016). Inter- and intra-annual variations of clumping

index derived from the MODIS BRDF product. International Journal of Applied Earth

Observation and Geoinformation, 44, 53-60

Huete, A., Didan, K., Miura, T., Rodriguez, E.P., Gao, X., & Ferreira, L.G. (2002). Overview of

the radiometric and biophysical performance of the MODIS vegetation indices. Remote

Sensing of Environment, 83, 195-213

Jackson, T.J., Chen, D., Cosh, M., Li, F., Anderson, M., Walthall, C., Doriaswamy, P., & Hunt, E.R.

(2004). Vegetation water content mapping using Landsat data derived normalized

difference water index for corn and soybeans. Remote Sensing of Environment, 92, 475-

482

Jacquemoud, S., & Baret, F. (1990). PROSPECT: A model of leaf optical properties spectra.

Remote Sensing of Environment, 34, 75-91

Jin, P., Wang, Q., Iio, A., & Tenhunen, J. (2012). Retrieval of seasonal variation in photosynthetic

capacity from multi-source vegetation indices. Ecological Informatics, 7, 7-18

Liu, Y. (2019). Master Thesis: Global Retrieval of the Maximum Carboxylation Rate of Plant

Leaves by Solar-induced Chlorophyll Fluorescence from TROPOMI. In, Department of

Geography & Planning: University of Toronto

Liu, Y., Liu, R., & Chen, J.M. (2012). Retrospective retrieval of long-term consistent global leaf

area index (1981–2011) from combined AVHRR and MODIS data. Journal of Geophysical

Research: Biogeosciences, 117

Liu, Y., Xiao, J., Ju, W., Zhu, G., Wu, X., Fan, W., Li, D., & Zhou, Y. (2018). Satellite-derived LAI

products exhibit large discrepancies and can lead to substantial uncertainty in simulated

carbon and water fluxes. Remote Sensing of Environment, 206, 174-188

Luo, X., Croft, H., Chen, J.M., He, L., & Keenan, T.F. (2019). Improved estimates of global

terrestrial photosynthesis using information on leaf chlorophyll content. Global Change

Biology, 25, 2499-2514

PeΓ‘Uelas, J., Filella, I., & Gamon, J.A. (1995). Assessment of photosynthetic radiation-use

efficiency with spectral reflectance. New Phytologist, 131, 291-296

Porcar-Castell, A., TyystjΓ€rvi, E., Atherton, J., van der Tol, C., Flexas, J., PfΓΌndel, E.E., Moreno,

J., Frankenberg, C., & Berry, J.A. (2014). Linking chlorophyll a fluorescence to

photosynthesis for remote sensing applications: mechanisms and challenges. Journal of

Experimental Botany, 65, 4065-4095

Rouse, J.W., Haas, R.H., Schell, J.A., & Deering, D.W. (1973). Monitoring vegetation systems in

the Great Plains with ERTS. NASA

Serrano, L., PeΓ±uelas, J., & Ustin, S.L. (2002). Remote sensing of nitrogen and lignin in

Mediterranean vegetation from AVIRIS data: Decomposing biochemical from structural

Page 79: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

68

signals. Remote Sensing of Environment, 81, 355-364

Sulman, B.N., Desai, A.R., Cook, B.D., Saliendra, N., & Mackay, D.S. (2009). Contrasting carbon

dioxide fluxes between a drying shrub wetland in Northern Wisconsin, USA, and nearby

forests. Biogeosciences, 6, 1115-1126

Ueyama, M., Iwata, H., Nagano, H., Tahara, N., Iwama, C., & Harazono, Y. (2019). Carbon dioxide

balance in early-successional forests after forest fires in interior Alaska. Agricultural and

Forest Meteorology, 275, 196-207

Ustin, S.L., Roberts, D.A., Gamon, J.A., Asner, G.P., & Green, R.O. (2004). Using Imaging

Spectroscopy to Study Ecosystem Processes and Properties. BioScience, 54, 523-534

Wang, Y.-P., Leuning, R., Cleugh, H.A., & Coppin, P.A. (2001). Parameter estimation in surface

exchange models using nonlinear inversion: how many parameters can we estimate and

which measurements are most useful? Global Change Biology, 7, 495-510

Willmott, C.J. (1981). ON THE VALIDATION OF MODELS. Physical Geography, 2, 184-194

Willmott, C.J. (1982). Some Comments on the Evaluation of Model Performance. Bulletin of the

American Meteorological Society, 63, 1309-1313

Wu, C., Niu, Z., Tang, Q., & Huang, W. (2008). Estimating chlorophyll content from hyperspectral

vegetation indices: Modeling and validation. Agricultural and Forest Meteorology, 148,

1230-1241

Zhang, Y., Chen, J.M., Miller, J.R., & Noland, T.L. (2008). Leaf chlorophyll content retrieval from

airborne hyperspectral remote sensing imagery. Remote Sensing of Environment, 112,

3234-3247

Zhou, Y., Ju, W., Sun, X., Hu, Z., Han, S., Black, T.A., Jassal, R.S., & Wu, X. (2014). Close

relationship between spectral vegetation indices and Vcmax in deciduous and mixed forests.

Tellus B: Chemical and Physical Meteorology, 66, 23279

Zhou, Y., Wu, X., Ju, W., Chen, J.M., Wang, S., Wang, H., Yuan, W., Andrew Black, T., Jassal, R.,

Ibrom, A., Han, S., Yan, J., Margolis, H., Roupsard, O., Li, Y., Zhao, F., Kiely, G., Starr,

G., Pavelka, M., Montagnani, L., Wohlfahrt, G., D'Odorico, P., Cook, D., Arain, M.A.,

Bonal, D., Beringer, J., Blanken, P.D., Loubet, B., Leclerc, M.Y., Matteucci, G., Nagy, Z.,

Olejnik, J., Paw U, K.T., & Varlagin, A. (2016). Global parameterization and validation of

a two-leaf light use efficiency model for predicting gross primary production across

FLUXNET sites. Journal of Geophysical Research: Biogeosciences, 121, 1045-1072

Page 80: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

69

Chapter 4

Summary

4.1 Main conclusions

1) Heterogeneities exist within the pixels of the π‘½π’„π’Žπ’‚π’™πŸπŸ“ product derived from TROPOMI sun-

induced chlorophyll fluorescence measurements. A downscaling algorithm was designed for

downscaling the π‘½π’„π’Žπ’‚π’™πŸπŸ“ dataset.

The intra-pixel heterogeneity of the π‘‰π‘π‘šπ‘Žπ‘₯25 dataset was explored at five tower flux sites in the USA.

Obvious heterogeneity was observed within the TROPOMI pixels over these sites. A downscaling

algorithm was designed to downscale π‘‰π‘π‘šπ‘Žπ‘₯25 by multiplying the TROPOMI π‘‰π‘π‘šπ‘Žπ‘₯

25 with a spatial

scaling ratio (SSR) at a higher resolution. The SSR considered informative downscaling factors,

such as leaf chlorophyll content (LCC) and normalized difference vegetation index (NDVI) in this

study. The SSR of each pixel was calculated as the proportion of the downscaling factor at that 1

km pixel and the averaged downscaling factor at that TROPOMI pixel at 0.1Β° resolution.

2) Leaf chlorophyll content (LCC) can be a feasible method for downscaling the π‘½π’„π’Žπ’‚π’™πŸπŸ“ data

from 0.1Β° Γ— 0.1Β° to 1 km Γ— 1 km. The LCC-downscaled π‘½π’„π’Žπ’‚π’™πŸπŸ“ data achieve better GPP

simulation results.

The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data were evaluated by using the π‘‰π‘π‘šπ‘Žπ‘₯

25 as inputs to the Boreal Ecosystem

Productivity Simulator (BEPS). The GPP simulation with LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data was

improved with better statistical metrics. As a proxy of plant photosynthetic capacity (Croft et al.,

2017), remotely sensed LCC can be informative at a higher spatial resolution for downscaling the

π‘‰π‘π‘šπ‘Žπ‘₯25 dataset. The improvement in GPP simulation after downscaling with LCC is not very large

because some of the tower flux sites selected in this study have downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 values close to

the values of the TROPOMI pixels.

3) Downscaling with a factor having tight physiological connections with π‘½π’„π’Žπ’‚π’™πŸπŸ“ can achieve

satisfying results, such as chlorophyll content in this study. Otherwise, it will introduce other

factors, such as NDVI, which may cause error propagation.

Page 81: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

70

Although some studies have indicated the correlations between NDVI and π‘‰π‘π‘šπ‘Žπ‘₯25 (Jin et al., 2012;

Zhou et al., 2014), weak physiological connections exist between those two parameters. NDVI

captures the greenness of the vegetation canopy, which is not only influenced by leaf optical

properties by also canopy structure, while π‘‰π‘π‘šπ‘Žπ‘₯25 indicates the maximum photosynthetic capacity

of leaves. The NDVI-based SSR improved the GPP estimates at some sites but worsened the

estimates at other sites, indicating interference of the variation in the canopy structure in the

downscaling process.

4) The first global 1 km π‘½π’„π’Žπ’‚π’™πŸπŸ“ map was produced based on the LCC-downscaling method

satisfactorily tested at sites.

The global downscaled 1 km π‘‰π‘π‘šπ‘Žπ‘₯25 map can capture details and intra-pixel spatial distribution of

π‘‰π‘π‘šπ‘Žπ‘₯25 . The downscaled map comprehensively considers the accurate numerical values of the

coarse-resolution π‘‰π‘π‘šπ‘Žπ‘₯25 dataset derived from TROPOMI solar-induced chlorophyll fluorescence

data and the within-pixel spatial information of LCC derived from Sentinel-2 data. The

downscaling method eliminates the mismatch between satellite-based π‘‰π‘π‘šπ‘Žπ‘₯25 datasets and the

footprint of tower fluxes and this new map will further enhance accurate global GPP estimates.

5) The correlations between PRI and LUE are not as pronounced as expected, though

statistically significant. The PRI-LUE correlations need to be further explored for the

purpose of downscaling π‘½π’„π’Žπ’‚π’™πŸπŸ“ .

The PRI-LUE correlations were experimented for each plant functional types using historical eddy

covariance measurements from 190 sites and for each of the five AmeriFlux sites available for

2018. The MODIS ocean band 11 (centered at 531nm) covers the signal band of PRI (Drolet et al.,

2005; Drolet et al., 2008; Middleton et al., 2016) and band 10 (centered at 488nm), band 12

(centered at 551nm), and band 13 (centered at 667nm) were tested as the reference band of PRI.

However, weak correlations were observed and the PRI-LUE points were widely dispersed from

their regression line. However, the PRI-LUE relationships are still statistically significant (p<0.05)

at most individual sites and for most plant functional types, suggesting that PRI contains

information for π‘‰π‘π‘šπ‘Žπ‘₯25 that may be useful for its downscaling in some ways which need to be

further explored.

Page 82: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

71

4.2 Limitation of current work and plan for further work

1) The design of the spatial scaling algorithm

In this study, a spatial scaling factor was adopted in the downscaling process, which was derived

as the proportion of the informative downscaling factor of the 1 km pixel and the averaged factor

of the TROPOMI pixel. However, the surface heterogeneity originates from two sources of

ecosystem heterogeneity, endogenous (biotic) and exogenous (abiotic) one (Chen et al., 2013).

Other factors, such as land cover, will be considered in further optimization of the spatial scaling

algorithm. Besides, temporal filtering will also be introduced to exclude outliers and generate the

seasonal variation of the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 .

2) The evaluation of the downscaled π‘½π’„π’Žπ’‚π’™πŸπŸ“

The downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 data were evaluated indirectly by using the downscaled π‘‰π‘π‘šπ‘Žπ‘₯

25 as the input

to the BEPS model for GPP simulation. The simulated GPP data were then compared with EC

GPP measurements for evaluation. Although in the GPP simulation all other parameters except

π‘‰π‘π‘šπ‘Žπ‘₯25 were consistent, the GPP simulation results were not only affected solely by π‘‰π‘π‘šπ‘Žπ‘₯

25 , but also

soil moisture, leaf area index, etc. (Liu et al., 1999; Liu et al., 1997). The LCC-downscaled π‘‰π‘π‘šπ‘Žπ‘₯25

improved the GPP simulation while some overestimations and underestimations were observed.

In further research, ground-based π‘‰π‘π‘šπ‘Žπ‘₯25 data or π‘‰π‘π‘šπ‘Žπ‘₯

25 products at higher resolutions when

available can be used to directly evaluate the accuracy of the downscaled π‘‰π‘π‘šπ‘Žπ‘₯25 .

3) The establishment of PRI-LUE correlations

The correlations of PRI-LUE for nine plant function types and for five sites were weak and could

not generate reliable GPP estimates through the LUE model (Monteith, 1972). Due to the lack of

an ideal reference band of PRI among MODIS reflectance bands, some substituting bands were

used. Many other complicating factors affect the signal of PRI (Gitelson et al., 2017a) and were

not considered in this study. In order to use remotely sensed PRI, careful consideration of the

optional definitions such as LUE formulation is also needed (Gitelson et al., 2017b). As discussed

in Section 2.4.1 and Section 2.4.2, the PRI retrieval and PRI-LUE correlations need to be improved

in further work. Besides, the Fluorescence Explorer (FLEX) mission of the European Space

Agency will be launched in the near future and the fluorescence imaging spectrometer (FLORIS)

Page 83: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

72

including a PRI band will be on board. The PRI band has a bandwidth of 500-600 nm and a spectral

resolution of 3 nm (Coppo et al., 2017), which will significantly enhance the quality of remotely

sensed PRI. With the FLEX PRI, better PRI-LUE correlations may be derived for various purposes

including π‘‰π‘π‘šπ‘Žπ‘₯25 downscaling.

Page 84: Downscaling the Maximum Carboxylation Rate ( π‘Ž Derived

73

4.3 References

Chen, J.M., Chen, X., & Ju, W. (2013). Effects of vegetation heterogeneity and surface topography

on spatial scaling of net primary productivity. Biogeosciences, 10, 4879-4896

Coppo, P., Taiti, A., Pettinato, L., Francois, M., Taccola, M., & Drusch, M. (2017). Fluorescence

imaging spectrometer (FLORIS) for ESA FLEX mission. Remote Sensing, 9, 649

Croft, H., Chen, J.M., Luo, X., Bartlett, P., Chen, B., & Staebler, R.M. (2017). Leaf chlorophyll

content as a proxy for leaf photosynthetic capacity. Global Change Biology, 23, 3513-3524

Drolet, G.G., Huemmrich, K.F., Hall, F.G., Middleton, E.M., Black, T.A., Barr, A.G., & Margolis,

H.A. (2005). A MODIS-derived photochemical reflectance index to detect inter-annual

variations in the photosynthetic light-use efficiency of a boreal deciduous forest. Remote

Sensing of Environment, 98, 212-224

Drolet, G.G., Middleton, E.M., Huemmrich, K.F., Hall, F.G., Amiro, B.D., Barr, A.G., Black, T.A.,

McCaughey, J.H., & Margolis, H.A. (2008). Regional mapping of gross light-use

efficiency using MODIS spectral indices. Remote Sensing of Environment, 112, 3064-3078

Gitelson, A.A., Gamon, J.A., & Solovchenko, A. (2017a). Multiple drivers of seasonal change in

PRI: Implications for photosynthesis 1. Leaf level. Remote Sensing of Environment, 191,

110-116

Gitelson, A.A., Gamon, J.A., & Solovchenko, A. (2017b). Multiple drivers of seasonal change in

PRI: Implications for photosynthesis 2. Stand level. Remote Sensing of Environment, 190,

198-206

Jin, P., Wang, Q., Iio, A., & Tenhunen, J. (2012). Retrieval of seasonal variation in photosynthetic

capacity from multi-source vegetation indices. Ecological Informatics, 7, 7-18

Liu, J., Chen, J.M., Cihlar, J., & Chen, W. (1999). Net primary productivity distribution in the

BOREAS region from a process model using satellite and surface data. Journal of

Geophysical Research: Atmospheres, 104, 27735-27754

Liu, J., Chen, J.M., Cihlar, J., & Park, W.M. (1997). A process-based boreal ecosystem productivity

simulator using remote sensing inputs. Remote Sensing of Environment, 62, 158-175

Middleton, E.M., Huemmrich, K.F., Landis, D.R., Black, T.A., Barr, A.G., & McCaughey, J.H.

(2016). Photosynthetic efficiency of northern forest ecosystems using a MODIS-derived

Photochemical Reflectance Index (PRI). Remote Sensing of Environment, 187, 345-366

Monteith, J.L. (1972). Solar Radiation and Productivity in Tropical Ecosystems. Journal of

Applied Ecology, 9, 747-766

Zhou, Y., Ju, W., Sun, X., Hu, Z., Han, S., Black, T.A., Jassal, R.S., & Wu, X. (2014). Close

relationship between spectral vegetation indices and Vcmax in deciduous and mixed forests.

Tellus B: Chemical and Physical Meteorology, 66, 23279