the application of unmanned aerial vehicle to precision

148
Utah State University Utah State University DigitalCommons@USU DigitalCommons@USU All Graduate Theses and Dissertations Graduate Studies 5-2016 The Application of Unmanned Aerial Vehicle to Precision The Application of Unmanned Aerial Vehicle to Precision Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Estimation Estimation Manal Elarab Utah State University Follow this and additional works at: https://digitalcommons.usu.edu/etd Part of the Civil and Environmental Engineering Commons Recommended Citation Recommended Citation Elarab, Manal, "The Application of Unmanned Aerial Vehicle to Precision Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Estimation" (2016). All Graduate Theses and Dissertations. 4891. https://digitalcommons.usu.edu/etd/4891 This Dissertation is brought to you for free and open access by the Graduate Studies at DigitalCommons@USU. It has been accepted for inclusion in All Graduate Theses and Dissertations by an authorized administrator of DigitalCommons@USU. For more information, please contact [email protected].

Upload: others

Post on 01-Mar-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Application of Unmanned Aerial Vehicle to Precision

Utah State University Utah State University

DigitalCommons@USU DigitalCommons@USU

All Graduate Theses and Dissertations Graduate Studies

5-2016

The Application of Unmanned Aerial Vehicle to Precision The Application of Unmanned Aerial Vehicle to Precision

Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration

Estimation Estimation

Manal Elarab Utah State University

Follow this and additional works at: https://digitalcommons.usu.edu/etd

Part of the Civil and Environmental Engineering Commons

Recommended Citation Recommended Citation Elarab, Manal, "The Application of Unmanned Aerial Vehicle to Precision Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Estimation" (2016). All Graduate Theses and Dissertations. 4891. https://digitalcommons.usu.edu/etd/4891

This Dissertation is brought to you for free and open access by the Graduate Studies at DigitalCommons@USU. It has been accepted for inclusion in All Graduate Theses and Dissertations by an authorized administrator of DigitalCommons@USU. For more information, please contact [email protected].

Page 2: The Application of Unmanned Aerial Vehicle to Precision

THE APPLICATION OF UNMANNED AERIAL VEHICLE TO PRECISION

AGRICULTURE: CHLOROPHYLL, NITROGEN, AND

EVAPOTRANSPIRATION ESTIMATION

by

Manal Elarab

A dissertation submitted in partial fulfillment of the requirements for the degree

of

DOCTOR OF PHILOSOPHY

in

Civil and Environmental Engineering

Approved:

______________________ ____________________ Dr. Mac McKee Dr. Niel Allen Major Professor Committee Member

______________________ ____________________ Dr. David Stevens Dr. Douglas Ramsey Committee Member Committee Member

______________________ ____________________ Dr. David Rosenberg Dr. Mark McLellan Committee Member Vice President for Research and Dean of the School of Graduate Studies

UTAH STATE UNIVERSITY Logan, Utah

2016

Page 3: The Application of Unmanned Aerial Vehicle to Precision

ii

Copyright © Manal Elarab 2016

All Rights Reserved

Page 4: The Application of Unmanned Aerial Vehicle to Precision

iii

ABSTRACT

The Application of Unmanned Aerial Vehicle to Precision Agriculture:

Chlorophyll, Nitrogen, and Evapotranspiration Estimation

by

Manal Elarab, Doctor of Philosophy Utah State University, 2016

Major Professor: Dr. Mac McKee Department: Civil and Environmental Engineering

Precision agriculture (PA) is an integration of a set of technologies aiming to

improve productivity and profitability while sustaining the quality of the surrounding

environment. It is a process that vastly relies on high-resolution information to enable

greater precision in the management of inputs to production. This dissertation explored

the usage of multispectral high resolution aerial imagery acquired by an unmanned aerial

systems (UAS) platform to serve precision agriculture application. The UAS acquired

imagery in the visual, near infrared and thermal infrared spectra with a resolution of less

than a meter (15 - 60 cm). This research focused on developing two models to estimate

cm-scale chlorophyll content and leaf nitrogen. To achieve the estimations a well-

established machine learning algorithm (relevance vector machine) was used. The two

models were trained on a dataset of in situ collected leaf chlorophyll and leaf nitrogen

measurements, and the machine learning algorithm intelligently selected the most

appropriate bands and indices for building regressions with the highest prediction

accuracy. In addition, this research explored the usage of the high resolution imagery to

Page 5: The Application of Unmanned Aerial Vehicle to Precision

iv

estimate crop evapotranspiration (ET) at 15 cm resolution. A comparison was also made

between the high resolution ET and Landsat derived ET over two different crop cover

(field crops and vineyards) to assess the advantages of UAS based high resolution ET.

This research aimed to bridge the information embedded in the high resolution imagery

with ground crop parameters to provide site specific information to assist farmers

adopting precision agriculture. The framework of this dissertation consisted of three

components that provide tools to support precision agriculture operational decisions. In

general, the results for each of the methods developed were satisfactory, relevant, and

encouraging.

(148 Pages)

Page 6: The Application of Unmanned Aerial Vehicle to Precision

v

PUBLIC ABSTRACT

The Application of Unmanned Aerial Vehicle to Precision Agriculture:

Chlorophyll, Nitrogen, and Evapotranspiration Estimation

by

Manal Elarab, Doctor of Philosophy Utah State University, 2016

Major Professor: Dr. Mac McKee Department: Civil and Environmental Engineering

Precision agriculture (PA) is an integration of a set of technologies aiming to

improve productivity and profitability while sustaining the quality of the surrounding

environment. It is a process that vastly relies on high-resolution information to enable

greater precision in the management of inputs to production. This dissertation explores

the usage of multispectral high resolution aerial imagery acquired by an unmanned aerial

systems (UAS) platform to estimate three fundamental crop parameters (plant

chlorophyll, leaf nitrogen and crop water demand). The UAS acquired imagery in the

visual, near infrared and thermal infrared spectra with a resolution of less than a meter

(15 - 60 cm). This research applied a well-established machine learning algorithm

(relevance vector machine) to the five band imagery (R, G, B, NIR, and TIR) to estimate

plant chlorophyll content and leaf nitrogen respectively. In addition, this research

explored the usage of the high resolution imagery to estimate crop evapotranspiration

(ET) at 15 cm resolution. A comparison was also made between the high resolution ET

and Landsat derived ET over two different crop cover (field crops and vineyards) to

Page 7: The Application of Unmanned Aerial Vehicle to Precision

vi

assess the advantages of UAS based high resolution ET. This research aims to bridge the

information embedded in the high resolution imagery with ground crop parameters to

provide site specific information to assist farmers adopting precision agriculture.

Manal Elarab

Page 8: The Application of Unmanned Aerial Vehicle to Precision

vii

I dedicate this work to Mohammad, Fairouz, Nada and Ahmad.

Your unconditional love and support created the person I am today, You made me believe in myself when I was in doubt.

I want to make you proud always.

… In the loving memory of Musa Nimah

Love you Manal

Page 9: The Application of Unmanned Aerial Vehicle to Precision

viii

ACKNOWLEDGMENTS

I am very grateful to my adviser and mentor, Dr. Mac McKee, for giving me this

opportunity, providing me with such a great project and team to work with, and above all

teaching me how to be a better researcher and person on daily basis. I would like to

extend my gratitude to all my committee members. I would like to thank Dr. Alfonso

Torres-Rua and Dr. Andres Ticlavilca for their guidance, help and support throughout my

education at USU. I appreciate all those who helped me collect and process the data

needed in my research (Roula Bachour, Shannon Syrstad, Mark Winkelaar, Alfonso

Torres, Ian Gowing, etc.). Thank you to all the AggieAir team that helped in the Scipio

project.

I also acknowledge all the members of AggieAir over the past few years who

have contributed indirectly to this work. Most of all, I owe this work to my loving

husband for his endless sacrifices and relentless support.

Page 10: The Application of Unmanned Aerial Vehicle to Precision

ix

CONTENTS

Page

ABSTRACT……………………………………………………………………………. iii PUBLIC ABSTRACT…………………………………………………………………… v DEDICATION…………………………………………………………………………. vii ACKNOWLEDGMENTS ............................................................................................... viii LIST OF TABLES ............................................................................................................. xi LIST OF FIGURES .......................................................................................................... xii CHAPTER 1: INTRODUCTION ....................................................................................... 1 Objectives ........................................................................................................................................ 3 Dissertation Organization ................................................................................................................ 4 Contributions ................................................................................................................................... 5 References ........................................................................................................................................ 6 CHAPTER 2: ESTIMATING CHLOROPHYLL FROM THERMAL AND BROADBAND MULTISPECTRAL HIGH RESOLUTION IMAGERY FROM AN UNMANNED AERIAL SYSTEM USING RELEVANCE VECTOR MACHINES FOR PRECISION AGRICULTURE ........................................................................................... 7 Abstract ............................................................................................................................................ 7 Introduction ...................................................................................................................................... 8 Materials and Methods ................................................................................................................... 14

RVMs ................................................................................................................... 14 Study Area ........................................................................................................... 18 Instrumentation: Remote Sensing Platform AggieAir ......................................... 19 Data Collection .................................................................................................... 23 Model Potential Inputs and Performance ............................................................. 24

Result and Discussion .................................................................................................................... 28 Conclusion ..................................................................................................................................... 34 Acknowledgment ........................................................................................................................... 36 References ...................................................................................................................................... 36 CHAPTER 3: USE OF HIGH RESOLUTION MULTISPECTRAL IMAGERY TO ESTIMATE PLANT NITROGEN IN OATS (AVENA SATIVA). ................................ 46 Abstract .......................................................................................................................................... 46 Introduction .................................................................................................................................... 47

Page 11: The Application of Unmanned Aerial Vehicle to Precision

x

Material and Methods .................................................................................................................... 51 Study Area ........................................................................................................... 51 Instrumentation: Remote Sensing Platform AggieAir ......................................... 52 AggieAir Data Processing ................................................................................... 53 Ground Data Acquisition ..................................................................................... 54 Relevance Vector Machine (RVM) ..................................................................... 54 Generating the Model Inputs and Targets ............................................................ 55 Model Configuration and Performance ............................................................... 57

Results and Discussion .................................................................................................................. 59 Conclusion ..................................................................................................................................... 66 References ...................................................................................................................................... 67 CHAPTER 4: ASSESSMENT ON THE USE OF AGGIEAIR UNMANNED AERIAL SYSTEM REMOTE SENSING CAPABILITIES TO ESTIMATE CROP EVAPOTRANSPIRATION AT HIGH SPATIAL RESOLUTION................................. 73 Abstract .......................................................................................................................................... 73 Introduction .................................................................................................................................... 74 Data Collection ..................................................................................................................................

Site Description.................................................................................................... 79 UAV Description ................................................................................................. 80 Sensors Description ............................................................................................. 82 Flight Details ....................................................................................................... 82

Processing Chain of AggieAir ....................................................................................................... 83 Mission Planning ................................................................................................. 83 Image Processing and Orthorectification ............................................................. 84 Radiometric Calibration ....................................................................................... 85 Thermal Calibration ............................................................................................. 86

Models: METRIC and METRIC High Resolution (METRIC-HR) ............................................... 87 METRIC .............................................................................................................. 88 METRIC-HR ....................................................................................................... 90 Albedo .................................................................................................................. 94

Results and Discussion .................................................................................................................. 96 Hot and Cold Pixel Selection ............................................................................... 96 METRIC Results.................................................................................................. 99 METRIC-HR Results ......................................................................................... 101 Comparison in Site S ......................................................................................... 104 Comparison in Site G ......................................................................................... 107

Conclusion ................................................................................................................................... 111 References .................................................................................................................................... 112 CHAPTER 5: SUMMARY.............................................................................................. 120 APPENDIX ................................................................................................................................ 123 CURRICULUM VITAE .......................................................................................................... 130

Page 12: The Application of Unmanned Aerial Vehicle to Precision

xi

LIST OF TABLES

Table Page

2.1 AggieAir Unmanned Aerial System Specifications .............................................. 20

2.2 Statistical Description of the Dataset Used for Plant Chlorophyll Estimation ..... 25

2.3 Vegetative Indices Formulation Used for Plant Chlorophyll Estimation ............. 26

2.4 Potential “Best Scenarios” .................................................................................... 29

3.1 Vegetative Indices Formulation Used for Leaf Nitrogen Estimation ................... 56

3.2 Statistical Description of the Dataset Used for Leaf Nitrogen Estimation ........... 57

3.3 Selected Model Characteristics ............................................................................. 60

4.1 AggieAir Unmanned Aerial System Features Description ................................... 81

4.2 The Developed Correction Equations for Each Band ........................................... 94

4.3 Details on the Selected Hot and Cold Pixels from the AggieAir and Landsat Imagery ................................................................................................................. 98

Page 13: The Application of Unmanned Aerial Vehicle to Precision

xii

LIST OF FIGURES

Figure Page

2.1 The location of the study area in Utah (left), the quarter used in plant chlorophyll estimation model (right) . .................................................................. 18

2.2 AggieAir airframe layout. ..................................................................................... 19

2.3 Relative spectral response of the VIS-NIR (left) and thermal camera (right). ..... 21

2.4 Raw natural color images from the unmanned aerial system (left), accurate orthorectified mosaic image from EnsoMOSAIC (center), and radiometric calibration of visible spectrum image (right). ....................................................... 22

2.5 Measured versus predicted chlorophyll concentration in the three flights for Model 4. Vertical yellow lines separate the three flight dates. ............................. 31

2.6 Model 4: Residual plot of the three flights (left) and one-by-one plot excluding the bare soil–zero chlorophyll points and reflecting the chlorophyll meter accuracy (right). .................................................................................................... 31

2.7 True-color maps, normalized difference vegetation index (NDVI) maps, leaf area index (LAI) maps, and the estimated chlorophyll concentration (μg.cm-2) map for the three different dates representing early growth, midgrowth, and early flowering. ..................................................................................................... 32

2.8 True-color image of flight zero (left), leaf area index (LAI) map (middle), and estimated chlorophyll concentration map (μg.cm-2, right). ................................... 33

3.1 The location of the study area in Utah (left), the quarter used in plant chlorophyll estimation model (right) .................................................................... 52

3.2 Measured versus predicted leaf nitrogen in the two flights used to build the model (above) and residual plot of the selected mode (below). ........................... 62

3.3 True-color maps (left), the estimated plant leaf nitrogen estimate map. .............. 63

3.4 June 17 image in false color (left), and estimated leaf nitrogen map (right). ....... 64

3.5 The six manageable areas as defined in this study. ............................................. 65

3.6 Average leaf nitrogen per sector during the crop growth. .................................... 66

4.1 The location of the study areas in Scipio, Utah, and Lodi, California. ................. 80

4.2 Details of AggieAir airframe . .............................................................................. 81

Page 14: The Application of Unmanned Aerial Vehicle to Precision

xiii

4.3 Relative quantum response of the VIS-NIR (left) and thermal camera (right). .... 82

4.4 An example of the summary report, the number of overlapped images range from 1 to 9 per frame. The black dots represent the center of the frame collected over the four flight lines . .............................................................................................. 85

4.5 An example of a ground control point (GCP), high vegetation, needed for calibration of thermal infrared (TIR) data. A picture of the GCP is captured right, a TIR imagery from a 3 m elevation in the middle, and the Blackbody reading on the left. ................................................................................................ 87

4.6 Landsat 8 and AggieAir relative spectral response. .............................................. 92

4.7 Landsat 8 point spread function (PSF) along the blue, green, and red (BGR) bands (left); Landsat 2D PSF representation on the right. .................................... 93

4.8 1:1 plot of the albedos from AggieAir and Landsat. ............................................ 96

4.9 Site S: Landsat reflectance in false color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (EtrF) output (right). ...................................................................................................... 100

4.10 Site G: Landsat reflectance bands in true color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (right). ........................................................................................... 101

4.11 Site S: AggieAir reflectance in false color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (center), TIR (Celsius) map (right) ............................................... 102

4.12 Site G: AggieAir reflectance bands in true color (left), Mapping Evapotranspiration With Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (center), TIR (celcuis) map (right). .. 103

4.13 Site S: Mapping Evapotranspiration With Internalized Calibration high resolution (METRIC-HR) reference evapotranspiration fraction (ETrf) on left, METRIC ETrf on right, both showing the 48 managable arcs as defined in this study. ........................................................................................................ 104

4.14 Northern center pivot reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs. .................................................................... 105

4.15 Southern center pivot reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs. .................................................................... 106

Page 15: The Application of Unmanned Aerial Vehicle to Precision

xiv

4.16 Site S: Comparison between reference evapotranspiration fraction (ETrf) estimates derived from Landsat and AggieAir data in the northern and southern center pivots. ....................................................................................................... 107

4.17 Mapping Evapotranspiration With Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) on left, METRIC high resolution ETrf in middle, and 19 blocks of individual manageable areas on right. ........................ 109

4.18 The average reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs for each block. ..................................................... 110

4.19 Site G: Comparison between reference evapotranspiration fraction (ETrf) estimates derived from Landsat and AggieAir data. ........................................... 111

Page 16: The Application of Unmanned Aerial Vehicle to Precision

CHAPTER 1

INTRODUCTION

Unmanned air systems (UAS) have been around for a century, and for a long time

their sole application was in military practice. Then, UAS started branching into other

applications like archeological site assessment and mineral exploration. In 2004, only

approximately 2% of the UAS were operating merely in the civil market (Newcome,

2004). Since then, use has been increasing quickly. A report discussing the economic

impact of UAS published by the Association for Unmanned Vehicle Systems

International (2013) estimated an economic impact of $82 billion by 2025. Those

estimates are rationalized by a large, emerging application for UAS: precision

agriculture.

Precision agriculture (PA), the intelligent crop production system, is a scientific

and modern approach to agriculture production in the 21st century. Building on

traditional knowledge, this production farming system integrates new technologies. The

three key technological components are (a) a remote sensing platform like UAS that

collects data; (b) a geographic information system, where data analysis and visualization

are performed using various techniques and tools; and (c) modern precision farming tools

like the variable rate applicator that allows the implementation of site-specific

recommendation. Site-specific in PA is a term that refers to the treatment of the smallest

possible area as a single element. Precision agriculture is designed to increase long-term,

site-specific production efficiency, productivity, and profitability while avoiding the

Page 17: The Application of Unmanned Aerial Vehicle to Precision

2

undesirable effects of excess chemical loading to the environment or productivity loss

due to insufficient input application.

Precision agriculture requires remote sensing platforms with unique features that

are not found in conventional platforms like satellites and airborne platforms. These

features prevail in UAS and are the following: (a) cost-effectiveness, with a much lower

operational cost compared to satellite and manned aircraft; (b) high spatial resolution that

could reach centimeters; and (c) practicality, such that the UAS can be flown any time,

under reasonable weather condition. In precision agriculture applications, UAS collects

aerial imagery of agricultural fields to monitor the health of crops, estimate nutrient

status, quantify crop water demand, and estimate yield and many other practices. All of

these help the farmer to achieve the “three glory Ms”: maximize yield, minimize resource

utilization, and minimize adverse environmental footprint.

The capability to increase the adoption of precision agriculture among farmers

relies on five specific points, which need to be addressed by the research community:

1. Identify the management challenges that the producers encounter daily,

including being agronomical, environmental, or economical.

2. Adopt cost-efficient, high spatial resolution platforms that can collect data to

address the identified issues.

3. Enhance the data-processing procedures from automatic photogrammetric

software, to instrument calibration and atmospheric correction.

4. Build algorithms that can extract and interpret the information in the processed

data.

Page 18: The Application of Unmanned Aerial Vehicle to Precision

3

5. Communicate these findings by building decision-support tools that can

support a better informed management decision to best management practices

for operations.

This research is a step in this direction. This dissertation explores remotely sensed

data acquired from a UAS to serve precision agriculture application. The UAS, named

AggieAir, was developed in the Utah Water Research Laboratory. This battery-powered

aircraft has onboard a payload computer that controls three sensors acquiring imagery in

the visible, near-infrared and thermal infrared spectra. The three sensors are ideal because

of their small size, light weight, and low cost. The imagery produced from AggieAir is of

a spatial resolution of 5–60 cm. After processing these images, they are used to develop

models that estimate crop chlorophyll content, plant leaf nitrogen content, and crop water

use. Platform and sensor description, image processing, radiometric calibration, and

estimation models are described in this research. The information generated by this

research is directly beneficial to both small- and large-scale producers and indirectly to

the environment.

Objectives

The chapters of this research each address specific objectives of this study.

Chapter 2 was written to address the following objectives: (a) investigate the suitability

of high spatial resolution data acquired by AggieAir to estimate chlorophyll

concentration, (b) develop a model that can estimate chlorophyll concentration from

remotely sensed crop surface reflectance, and (c) determine which bands in the

reflectance spectrum are sensitive to estimation of leaf chlorophyll concentration. The

Page 19: The Application of Unmanned Aerial Vehicle to Precision

4

objectives of Chapter 3 were the following: (a) determine whether leaf nitrogen content

could be predicted from reflectance data collected by AggieAir, and (b) develop a model

that can estimate leaf nitrogen of oat crop with a 15 cm spatial resolution. The objectives

of Chapter 4 were the following: (a) explore the applicability of AggieAir 15 cm

resolution data collected in the visible, near-infrared, and thermal infrared spectra to

estimate crop evapotranspiration (ET); (b) propose modification on existing model inputs

to accommodate AggieAir data; (c) compare ET estimations derived from Landsat data to

those from AggieAir; and (d) develop a 15 cm ET estimate with recommendations on

farm manageable units.

Dissertation Organization

This dissertation’s objectives, outlined above, were met in a three-manuscript

format. In Chapter 2, the retrieval of cm-scale chlorophyll content from thermal and

multispectral optical imagery collected by a UAS is presented. A relevance vector

machine is trained on a dataset of in situ collected leaf chlorophyll measurements, and the

machine learning algorithm intelligently selects the most appropriate bands and indices

for building regressions with the highest prediction accuracy. Chapter 3 applies the same

methodology used in Chapter 2 to estimate leaf nitrogen content. The model recommends

a set of inputs that are needed to estimate the spatial distribution of nitrogen at a 15 cm

spatial resolution. Chapter 4 discusses the application of AggieAir data to map ET at a 15

cm spatial resolution. A detailed description of the pre- and post-processing procedure of

the AggieAir data is presented in this chapter. Chapter 4 also suggests adjustments done

on a surface energy balance algorithm with inputs to accommodate AggieAir imagery. A

Page 20: The Application of Unmanned Aerial Vehicle to Precision

5

comparison per the smallest manageable area is presented for better irrigation

management practices over the two study sites. Chapter 5 provides a summary of this

work, draws the major conclusions, and presents recommendations for further research.

The structure of this dissertation is based on the multiple-paper format. As a result, some

redundancies and repetition of parts of the material presented occur, especially the

description of the study area and the UAS platform. Additionally, each chapter is edited

in a stand-alone format, with acronym usage and references specific to each chapter.

Contributions

The findings of this dissertation contribute towards a greater understanding of

high spatial resolution data acquired by a UAS in precision agriculture application.

Specifically, this dissertation contributes the following:

The research demonstrates the capability of a UAS system named AggieAir to

successfully determine important agronomical parameters to be used in a

precision agriculture farming system (Chapter 2, 3, 4).

The research introduces Bayesian-based learning machine algorithms in precision

agriculture research (Chapter 2, 3).

The research identifies sensitive bands and indices for building regressions with

the highest prediction accuracy that can estimate plant chlorophyll content and

nitrogen leaf content (Chapter 2, 3).

Whereas other documents may contain some of these steps, this is the first

document to include a detailed description of the AggieAir platform (sensors,

Page 21: The Application of Unmanned Aerial Vehicle to Precision

6

thermal calibration, radiometric calibration, flight planning, and imagery

orthorectification) (Chapter 4).

The research demonstrates the potential of using data-mining learning machine

algorithms to build regressions that relate spectral information and ground

samples in a spatial context (Chapter 2, 3).

Findings demonstrate that crop water demand estimates using UAS high-

resolution data were comparable to the estimates generated from Landsat 8 data

(Chapter 4).

The research identifies the spectral difference of consumer-grade cameras and

Landsat 8 sensors and applies an adequate correction procedure (Chapter 4).

The paper presents a novel approach of comparing estimates that replaces the

traditional pixel-by-pixel comparison with a unit based on the irrigation

management system and design (Chapter 3, 4).

References

Association for Unmanned Vehicle Systems International. (2013). The economic impact of unmanned aircraft systems integration in the United States. Retrieved from http://www.auvsi.org/auvsiresources/economicreport

Newcome, L. R. (2004). Unmanned aviation: A brief history of unmanned aerial vehicles. Reston, VA: American Institute of Aeronautics and Astronautics.

Page 22: The Application of Unmanned Aerial Vehicle to Precision

7

CHAPTER 2

ESTIMATING CHLOROPHYLL FROM THERMAL AND BROADBAND

MULTISPECTRAL HIGH RESOLUTION IMAGERY

FROM AN UNMANNED AERIAL SYSTEM USING RELEVANCE

VECTOR MACHINES FOR PRECISION AGRICULTURE1

Abstract

Precision agriculture requires high-resolution information to enable greater

precision in the management of inputs to production. Actionable information about crop

and field status must be acquired at high spatial resolution and at a temporal frequency

appropriate for timely responses. In this study, high spatial resolution imagery was

obtained through the use of a small, unmanned aerial system called AggieAirTM.

Simultaneously with the AggieAir flights, intensive ground sampling for plant

chlorophyll was conducted at precisely determined locations. This study reports the

development of a relevance vector machine (RVM) coupled with cross validation and

backward elimination to a dataset composed of reflectance from high-resolution

multispectral imagery (visible/near-infrared, or VIS-NIR), thermal infrared (TIR)

imagery, and vegetative indices, in conjunction with in situ SPAD measurements from

1 Reprinted with some editorial corrections from “Estimating Chlorophyll With Thermal and Broadband Multispectral High Resolution Imagery From an Unmanned Aerial System Using Relevance Vector Machines for Precision Agriculture,” by M. Elarab, A. M. Ticlavilca, A. F. Torres-Rua, I. Maslova, and M. McKee, 2015, International Journal of Applied Earth Observation and Geoinformation, 43, 32-42. © 2015 with permission from Elsevier.

Page 23: The Application of Unmanned Aerial Vehicle to Precision

8

which chlorophyll concentrations were derived, to estimate chlorophyll concentration

from remotely sensed data at 15 cm resolution. The results indicate that an RVM with a

thin plate spline kernel type and kernel width of 5.4, having leaf area index (LAI),

normalized difference vegetation index (NDVI), thermal and red bands as the selected set

of inputs, can be used to spatially estimate chlorophyll concentration with a root mean-

squared error (RMSE) of 5.31 μg.cm-2, efficiency of 0.76, and 9 relevance vectors.

Keywords: Remote Sensing, High Spatial Resolution Imagery, Relevance Vector

Machine, Precision Agriculture, Chlorophyll Concentration

Introduction

Increasing world population levels will bring increased demand for food, water,

and agricultural inputs. Various agricultural farming strategies are being reevaluated to

determine how to improve food production, minimize environmental impact, and reduce

costs. Among many, precision agriculture has evolved as a viable system to improve

profitability and productivity (Daberkow, McBride, Robert, Rust, & Larson, 2000;

Lambert & Lowenberg-De Boer, 2000; Swinton & Lowenberg-DeBoer, 1998). Precision

agriculture is a process of finely adjusting agricultural inputs (e.g., water, nutrients) and

in-field practices (e.g., irrigation, fertilization), through the use of site-specific

information and spatial imagery, to improve measures of agricultural productivity (e.g.,

yield, net farm income; Pierce & Nowak, 1999).

Use of spatial imagery in agriculture has been the focus of many studies for the

past five decades (Bauer, 1985; Benedetti & Rossini, 1993; Franke & Menz, 2007; Idso,

Jackson, & Reginato, 1977; MacDonald & Hall, 1980; Mathur &Foody, 2008; Shanahan

Page 24: The Application of Unmanned Aerial Vehicle to Precision

9

et al., 2001; Stone et al., 1996), requiring increased investments in relevant research and

technologies (Schellberg, Hill, Gerhards, Rothmund, & Braun, 2008) that indicate remote

sensing can be a valuable tool to enhance precision agriculture (Haboudane, Miller,

Tremblay, Zarco-Tejada, & Dextraze, 2002; Lamb & Brown, 2001; Seelan, Laguette,

Casady, & Seielstad, 2003). However, remote sensing has yet to reach its full capability

in precision agriculture applications. Lack of fine spatial resolution and near real-time

data, compounded by high costs, has hindered remote sensing applications at the field

scale (Brisco, Brown, Hirose, McNairn, & Staenz, 1998; Kalluri, Gilruth, Bergman, &

Plante, 2002; Liaghat & Balasundram, 2010; M. S. Moran, Inoue, & Barnes, 1997).

Thirty years ago, Jackson (1984) envisioned an autonomous remote sensing platform that

could overcome most of the limitations; this is becoming a reality with the introduction of

affordable unmanned aerial systems (UAS). UAS, a potential substitute for satellite-based

remote sensing, are gaining attention and recognition in the scientific community as a

potential technology that can generate high spatial resolution imagery (< 1 m) and at a

temporal frequency appropriate for timely responses in the production of actionable

information about crop and field status.

One such UAS, named AggieAirTM, was developed by the Utah Water Research

Laboratory at Utah State University. AggieAir is designed to carry camera payloads to

acquire high-resolution, georeferenced aerial imagery to be used in various water, natural

resources, and agricultural applications, including precision agriculture. AggieAir holds

three sensors: Sensors 1 and 2 are consumer-grade cameras (personal point-and-click

cameras) that capture imagery, depending on flight elevation above ground, of 6–25 cm

Page 25: The Application of Unmanned Aerial Vehicle to Precision

10

resolution in the VIS (red, green, blue spectrum) and NIR spectrum, respectively. Sensor

3 is a micro bolometer thermal camera that captures images of 30–150 cm resolution in

the TIR spectrum. The three sensors are ideal because of their small size, light weight,

low cost, and high resolution. The use of high-resolution imagery (< 1 m) can potentially

improve the ability to evaluate the spatial dynamics of chlorophyll and detect its temporal

variation. In this study, the use of multispectral VIS-NIR-thermal high-resolution

imagery is investigated as a tool to estimate plant chlorophyll concentration to provide

time-critical information for precision agriculture.

Chlorophyll concentration, measured in mass per unit leaf area (μg cm-2), is an

important biophysical parameters retrievable from reflectance data. Chlorophyll is a vital

pigment primarily responsible for harvesting light energy used in photosynthesis (Evans,

1989; Niinemets & Tenhunen, 1997; Sims & Gamon, 2002) and is therefore an excellent

indicator of a crop’s overall physiological status (Evans, 1989; Yoder & Pettigrew-

Crosby, 1995), stress or disease (Chaerle & Van Der Straeten, 2000; Peñuelas & Filella,

1998; Zarco-Tejada Miller, Morales, Berjón, & Agüera, 2004), and yield predictions

(Dawson, North, Plummer, & Curran, 2003; Gitelson et al., 2006). Chlorophyll can

potentially provide an assessment of leaf nitrogen, an essential plant nutrient, due to the

close relationship between leaf chlorophyll and leaf nitrogen (Daughtry, Walthall, Kim,

De Colstoun, & McMurtrey, 2000; J. A. Moran, Mitchell, Goodmanson, & Stockburger,

2000; Wood, Tracy, Reeves, & Edmisten, 1992). Chlorophyll concentration varies with

vegetation growth, thus estimating chlorophyll across the field at different growth stages

could offer the farmer time- and location-specific critical information ideal for assisting

Page 26: The Application of Unmanned Aerial Vehicle to Precision

11

decision makers in monitoring their crops and managing farming activities to achieve

maximum production.

Several leaf scale studies have focused on estimating chlorophyll concentration

from VIS-NIR reflectance data. These studies indicated that the green and far-red regions

of the visible spectrum are sensitive to variations in chlorophyll concentrations (Datt,

1999; Demarez & Gastellu-Etchegorry, 2000; Gitelson & Merzlyak, 1994; Kim, 1994;

Zarco-Tejada Miller, Noland, Mohammed, & Sampson, 2001). Various successful

indices have been formulated to estimate chlorophyll concentration (Broge & Leblanc,

2001; Haboudane et al., 2002; le Maire, François, & Dufrêne, 2004). Some of these

indices are ratios of reflectance in individual narrow visible wavebands (Blackburn,

1998; Carter & Spiering, 2002) or ratios of reflectance in VIS and NIR (Gitelson,

Kaufman, & Merzlyak, 1996), whereas others are red-edge reflectance ratio indices

(Gitelson & Merzlyak, 1994; Kim, Daughtry, Chappelle, McMurtrey, & Walthall, 1994;

Zarco-Tejada & Miller, 1999) or first and second derivatives of reflectance spectra

(Miller, Hare, & Wu, 1990). Composites of indices have been developed (Haboudane et

al., 2002) in an attempt to correct for distortions in the reflectance data caused by soil

background effect and canopy architecture.

Detailed discussions and thorough reviews concerning appropriate optimal

wavelengths and various chlorophyll indices can be found in the literature (Broge &

Leblanc, 2001; Haboudane, Miller, Pattey, Zarco-Tejada, & Strachan, 2004). However,

most of the studies have had low spatial and coarse spectral resolution characteristics;

therefore, the applicability of those indices to high spatial resolution airborne data cannot

Page 27: The Application of Unmanned Aerial Vehicle to Precision

12

be evaluated. Regarding thermal imagery, it is mainly explored when information on

plant water status is in question, for example when screening drought-tolerance

genotypes (Blum, Mayer, & Gozlan, 1982), detecting crop water stress levels (Berni,

Zarco-Tejada, Suárez, & Ferereset, 2009), or estimating soil moisture and

evapotranspiration (Hassan Esfahani, Torres-Rua, Jensen, & McKee, 2014a, 2014b;

Jackson, Idso, Reginato, & Pinter, 1981; Wallace, Lucieer, Watson, & Turner, 2012).

However, TIR data have not been investigated in estimating chlorophyll yet. Exploring

thermal data in this study is rationalized by the close relationship between heat stress and

the photosynthetic capacity of the leaves (Raison, Roberts, & Berry, 1982; Sharkey,

2005) and consequently the chlorophyll concentration. The mechanism by which

moderate heat stress reduces photosynthetic capacity has been debated since the 1980s,

when researchers attributed the photosynthesis inhibition to different factors such as the

impairment of electron transport activity or the inactivation of Rubisco (Berry &

Bjorkman, 1980; Murakami, Tsuyama, Kobayashi, Kodama, & Iba, 2000; Salvucci &

Crafts‐ Brandner, 2004; Weis, 1981).

Estimating chlorophyll at a canopy level from optical remotely sensed data can

generally be carried out by several methodologies. The simplest methodology that is

widely accepted is the empirical method, such as those based on vegetation indices

(Johnson, Hlavka, & Peterson, 1994). Nevertheless, indices generated in this context are

inclined to unstable performance when applied to images that differ from the designed

method (Verrelst, Schaepman, Malenovský, & Clevers, 2010). Physical behavior based

methods are another approach to formulating estimates from remotely sensed data. This

Page 28: The Application of Unmanned Aerial Vehicle to Precision

13

method is based on physical laws that describe the transfer and interaction of radiation

within the atmospheric column and canopy, such as radiative transfer models (Myneni et

al., 1995). This approach has become more promising with advances in atmospheric

radiative transfer modeling. The biggest drawback for such a model is that it requires

site-specific information for proper model parameterization, which is not always

available. As a result, methods based on vegetation indices or physical models may be

either too simple or too complex to deliver accurate estimates (Baret & Buis, 2008).

Several books and published papers have reviewed these methodologies and highlighted

the advantages and disadvantages associated with the complexity of the modeling

approach selected, and the degree of general or local applicability of the methodology in

remote sensing (Baret & Buis, 2008; Zarco-Tejada et al., 2001).

Considerable research has been carried out to explore advanced computational

methods that are both accurate and robust. Machine learning regression algorithms

present a potential approach for generating adaptive; robust; and, once trained, fast

estimates (Hastie, Tibshirani, & Friedman, 2009; Knudby, LeDrew, & Brenning, 2010).

Recent studies have demonstrated successful performance of a very well-known machine

learning algorithm in estimating biophysical parameters using neural network models

(Cipollini, Corsini, Diani, & Grasso, 2001; De Martino et al., 2002; González Vilas,

Spyrakos, & Torres Palenzuela, 2011; Hassan Esfahani et al., 2014a; Verrelst, Alonso,

Camps-Valls, Delegido, & Moreno, 2012). In recent studies, neural networks are being

replaced by more advanced regression-based methods that are simpler to calibrate, like

support vector machines (SVMs; Camps-Valls, Bruzzone, Rojo-Álvarez, & Melganiet,

Page 29: The Application of Unmanned Aerial Vehicle to Precision

14

2006; Moser & Serpico, 2009; Pal & Mather, 2005) and RVMs (Camps-Valls, Gómez-

Chova et al., 2006). SVMs have been widely used in various remote sensing applications;

nevertheless, their large computational complexity is a major drawback. This complexity

of SVM models is due to their liberal use of basis functions that typically grow linearly

with the size of the training set (Tipping, 2001). Studies have shown that the behavior of

RVMs is often superior to that of SVMs (Demir & Erturk, 2007). The results given by

Tipping (2001) demonstrated that the RVM has a comparable generalization performance

to the SVM, while requiring dramatically fewer kernel functions or model terms. RVM,

in a statistical learning method proposed by Tipping, constitutes a Bayesian

approximation for solving nonlinear regression models and is often used for classification

and pattern recognition. RVMs offer excellent sparseness characteristics, are robust, and

can produce probabilistic outputs that permit the capture of uncertainty in the predictions

(Gómez-Chova, Muñoz-Marí, Laparra, Malo-López, & Camps-Valls, 2011;

Thayananthan, Navaratnam, Stenger, Torr, & Cipolla, 2008).

The main purposes of this study were to (a) introduce AggieAir as a successful

tool for use in precision agriculture; (b) explore the use of VIS, NIR, and TIR in

estimating chlorophyll concentration, and (c) use RVM algorithms to formulate spatially

distributed chlorophyll concentration estimates.

Materials and Methods

RVMs

This section presents a brief description of RVMs relevant to this study. Tipping

introduced the RVM in 2001. The RVM was developed with a Bayesian framework to

Page 30: The Application of Unmanned Aerial Vehicle to Precision

15

find sparse solutions in classification and regression studies based on acquiring relevance

vectors and weights by maximizing a marginal likelihood. In RVM regression models,

the weight of each input is governed by a set of hyperparameters that describe posterior

distribution of the weights and are estimated iteratively during the machine learning

training step (Tipping, 2001). This paper adopts the RVM introduced by Tipping (2004),

which resembles the 2001 model. The main feature in the 2004 model is that the inferred

predictors are even sparser, with relatively few relevance vectors. This model also offers

good generalization performance (Yuan, Wang, Yu, & Fang, 2007).

To build the model, input-output vector pairs {𝑥𝑖, 𝑦𝑖}𝑖=1𝑁 are sampled from a data

set of N input vectors {𝑿𝒏}𝒏=𝟏𝑵 with corresponding N output values {𝐲𝒏}𝒏=𝟏𝑵 . From these

vector paired data, we generate a training data subset from which the model learns the

dependence between inputs and the output target, with the purpose of making accurate

predictions of for previously unseen values of shown in Equation 1:

𝑦 = 𝑤𝜑(𝑥) + 𝜀 (1)

where is a vector of weight parameters, 𝜑(𝑥) = [1, 𝑓(𝑥, 𝑥1,… , 𝑓(𝑥, 𝑥𝑁)]^𝑇 is a

design matrix of 𝑁 + 1 vectors of kernel basis functions , is the error that for

algorithmic simplicity is assumed to be zero-mean Gaussian with variance .

The kernel or basis function is a method that detects embedded patterns in the

data by transforming or extending linear algorithms into nonlinear ones. Kernel methods

map the data into higher dimensional spaces to increase the computational power of the

machine (Cristianini & Shawe-Taylor, 2000; Genton, 2001; Souza, 2010; Vapnik, 2000).

Kernel functions could be linear, polynomial, and Gaussian kernel. However, choosing

y x

w

f ε

2

Page 31: The Application of Unmanned Aerial Vehicle to Precision

16

the most appropriate one highly depends on the nature of the relationship between the

inputs and outputs. Six kernel types, f, were considered: Gauss, Laplace, spline, Cauchy,

thin plate spline, and bubble (Bachour, Walker, Ticlavilca, McKee, & Maslova, 2014;

Ticlavilca, McKee, & Walker, 2013; Torres, Walker, & McKee, 2011). The process of

selecting the kernel type in this paper was conducted by trial and error.

The Gaussian likelihood of the data set can be written as in Equation 2:

2

22/2

2exp)2(),|(

wywyp NN

(2)

One of the classic approaches to estimating the parameters and in Equation

2 is using the method of maximum likelihood. However, with many parameters used as

training observations, the maximum likelihood estimation would lead to severe

overfitting (Tipping, 2004). To overcome this complexity, Tipping (2001) proposed

adding a “prior” to constrain the selection of parameters by defining an explicit zero-

mean Gaussian prior probability distribution over them as shown in Equation 3:

2exp)2()|(

2

1

2/12/ mmM

mm

M wwp

(3)

Where M is the number of independent hyperparameters TM ),...,( 1 . Each is

associated independently with every weight to moderate the strength of the prior and

provide the sparsity of the model (Tipping, 2001). How far each weight is allowed to

deviate from zero is controlled by the hyperparameter vectors (Yuan et al., 2007).

Consequently, using Bayes’s posterior inference, the posterior over W could be computed

as shown in Equation 4:

w 2

Page 32: The Application of Unmanned Aerial Vehicle to Precision

17

),|()|(),|(),,|( 2

22

ypwpwypywp (4)

Here, ),|( 2yp is the normalizing factor; ),|( 2wyp and )|( wp are both

Gaussian priors, so the posterior is also Gaussian with ),,|( 2ywp ~ ),( wN . The

posterior mean and covariance are defined as:

12 )( TA And (5)

yT 2 , (6)

where is .

A fast marginal likelihood optimization algorithm is used to obtain the optimal set

of hyperparameters, . This optimization algorithm uses an efficient sequential

addition and deletion of candidate basis functions described by Tipping and Faul (2003).

Given an unseen input vector, , the predictive distribution for the

corresponding target can be computed. This search for optimal hyperparameters is

learned using a type II maximum likelihood method coupled with iterative estimation

(Tipping, 2001) as shown in Equation 7:

)dw),(σp(w|y,α))p(y*|w,(σ)=),(σp(y*|y,α optoptoptoptopt 222 (7)

=> )σ*,N(y*|)),(σp(y*|y,α o p to p t 22 *)(

Where is the predictive mean of the output of the unseen data, and the posterior

mean weight of , TM ],...,[* **

1 ; and TM ])(,...,)[(*)( 2*2*

12 is the predictive

variance. This predictive variance is the sum of variances associated with both the noise

of the data and the uncertainty in the prediction of the weight parameters (Tipping, 2004).

A ), . . . ,( 1 Mdi ag

o p t

*x

*y

* *x

Page 33: The Application of Unmanned Aerial Vehicle to Precision

18

In this optimization process, the vectors from the training set associated with nonzero

weights are called the relevance vectors. The theory behind RVM, mathematical

formulation, likelihood maximization, and optimization procedure is discussed in detail

in Tipping (2004) and Tipping and Faul (2003).

Study Area

The field study was carried out in the summer of 2013 on privately owned

agricultural land in Scipio, Utah (39°14'N 112°6'W; see Figure 2.1). The plot, mainly

composed of loamy clay soil, was equipped with a center pivot sprinkler for irrigating

and fertigation oats (Avena sativa). The study area was restricted to the northwest quarter

of the center pivot so that samples could be collected within a close time frame relative to

the AggieAir flight. AggieAir aircraft were flown four times over the area, covering the

entire growth cycle of oats. The flights on 05/16, 06/01, 06/09, and 06/17 reflected the

four stages of growth: 10 days after planting, early growth, mid growth, and early

flowering. Oats were harvested after the fourth flight to be used as forage.

Figure 2.1. The location of the study area in Utah (left), The quarter used in plant chlorophyll estimation model (right).

Page 34: The Application of Unmanned Aerial Vehicle to Precision

19

Instrumentation: Remote Sensing Platform AggieAir

AggieAir is a UAS designed to carry camera payloads to acquire aerial imagery

for precision agriculture and other types of applications (Figure 2.2). The UAS aircraft is

battery powered and equipped with a payload system (which includes three cameras and

a computer), avionics, two inertial sensors (a GPS module and an inertial measurement

unit), radio controller, and flight control. The aircraft is propelled using an electric,

brushless motor. It does not require a runway and can be flown autonomously or

manually. In autonomous mode, the aircraft follows a preprogrammed flight plan

containing navigation waypoints defined by GPS and altitude. While operational, the

payload computer instructs the three cameras to acquire imagery in the VIS, NIR, and

thermal spectra and records the position and orientation of the aircraft when each image

is taken. Table 2.1 illustrates the UAS specification in more detail.

Figure 2.2. AggieAir airframe layout.

Page 35: The Application of Unmanned Aerial Vehicle to Precision

20

Table 2.1

AggieAir Unmanned Aerial System Specifications

Specification Range

Flight duration 45–60 min

Flight altitudes 200–1000 m

Maximum takeoff weight 6.35 kg

Visible/near-infrared resolution 6–25 cm

Thermal resolution 30–150 cm

Wing span 2.5 m

The VIS camera used in AggieAir is a Canon S-95, with a 10-megapixel charge-

coupled device (CCD) sensor and an ISO range of 80 to 3200. The radiometric resolution

of the Canon S-95 is 8-bit color, which means that the digital measurement for a

particular pixel in a given spectral band ranges from 0 to 255. The NIR camera is an

identical Canon S-95, modified by replacing the manufacturer’s optical filter with a

Wratten 87 NIR filter that allows NIR wavelengths of 750 nm. The relative spectral

responses of the VIS-NIR cameras were not provided by the manufacturers but were

obtained using the algorithm provided by Jiang, Liu, Gu, and Susstrunk (2013). The

camera VIS-NIR spectral response is shown in the left portion of Figure 2.3. AggieAir

also carries a small, low-power, microbolometer thermal camera from Infrared Cameras,

Inc. (2014). The relative spectral response of the thermal camera is shown in the right

portion of Figure 2.3.

Page 36: The Application of Unmanned Aerial Vehicle to Precision

21

Figure 2.3. Relative spectral response of the VIS-NIR (left) and thermal camera (right). VIS-NIR = visible/near-infrared; ICI = Infrared Cameras, Inc.

Following VIS and NIR image acquisition, a two-step processing phase occurs.

The first step is image mosaicking and orthorectification. This technique, achieved with

EnsoMOSAIC software (MosaicMill, 2009), combines all of the images into one large

mosaic and rectifies it into a ground coordinate system. The software generates hundreds

of tie-points between overlapping images by using photogrammetric principles in

conjunction with image GPS log file data and exterior orientation information from the

on-board cameras to refine the estimate of the position and orientation of individual

images. The resulting image is an orthorectified digital number mosaic. The second step

involves radiometric calibration: the conversion of the digital pixels into a measure of

reflectance. This conversion is based on methods adapted from the research (Crowther,

1992; Miura & Huete, 2009; Neale & Crowther, 1994). The major steps involved in this

0

0.2

0.4

0.6

0.8

1

380 480 580 680 780 880

Wavebands (nm)

VIS-NIR Relative Spectral Response

0

0.2

0.4

0.6

0.8

1

1200 6200 11200 16200 21200

Wavebands (nm)

ICI Thermal Camera Relative Spectral Response

Page 37: The Application of Unmanned Aerial Vehicle to Precision

22

methodology are the reference panel calibration and solar zenith angle calculations. This

method converts raw airborne multispectral data by calculating the ratio of linearly

interpolated reference values from the pre- and post-flight reference panel readings, as

discussed in detail by Zaman, Jensen, Clemens, and McKee (2014) and by Clemens

(2012). After completing the two-step process, images are geometrically rectified and

radiometricaly corrected to obtain a four-layer (red, green, blue, NIR) canopy surface

reflectance in a single image (Figure 2.4).

Figure 2.4. Raw natural color images from the unmanned aerial system (left), accurate orthorectified mosaic image from EnsoMOSAIC (center), and radiometric calibration of visible spectrum image (right).

Thermal imagery processing also requires an initial step of mosaicking and

orthorectification similar to the VIS and NIR images. However, the resulting thermal

mosaic is composed of brightness temperature in degrees Celsius (± 0.1 degrees) instead

of digital numbers. Compensating for external disturbance and geometric calibration are

also unique challenges associated with the thermal camera. Jensen (2014) thoroughly

explained the methodology of processing thermal maps adopted by the authors.

Page 38: The Application of Unmanned Aerial Vehicle to Precision

23

Data Collection

The collection of the ground and remotely sensed data occurred under similar

weather conditions in a 1- to 2-hour window.

Multispectral image acquisition. Four multispectral mosaics were acquired by

AggieAir during summer 2013. Acquisition dates were planned to coincide with different

development stages and with overflights of Landsat. Images were collected, following the

Landsat image acquisition protocol, close to solar noon (between 12 a.m. and 1 p.m.).

The flight time (beginning to end) ranged from 30–40 min. All four missions were

successfully performed, providing image data covering the earliest, middle, and latest

periods of the oat growth. The spatial resolution is 0.15 m for VIS and NIR images and

0.6m for the TIR images.

Ground data acquisition. Intensive ground truth sampling of plant chlorophyll

was conducted simultaneously with the AggieAir flights at precise GPS locations. The

GPS data were collected using an rtkGPS with < 1 mm precision in a 1 Hz bandwidth

(Trimble® R8, Global Navigation Satellite System, Dayton, Ohio). A SPAD-502

chlorophyll meter (Minolta Corporation, New Jersey) was used for in vivo measurement

of the ratio of light transmittance through the leaf at wavelengths of 650 and 940 nm.

Instrument readings have been shown to correlate well to laboratory measurements of

chlorophyll concentration in several species (Yadawa, 1986). On each sampling

campaign, 40 SPAD measurements were collected on average. The chlorophyll meter

readings were taken midway on the fully expanded top-of-canopy leaves. Each

measurement was characterized by the mean of six replicate measurements. The

Page 39: The Application of Unmanned Aerial Vehicle to Precision

24

chlorophyll meter measures an area of 2 x 3 mm with an accuracy of ± 1.0 SPAD unit (at

room temperature). However, the SPAD-502 meter displays the chlorophyll readings in

arbitrary units (SPAD unit) rather than the actual amounts of chlorophyll in mass per leaf

area; thus, further conversions were needed. The SPAD units were transformed to a

chlorophyll concentration index (CCI) unit using Equation 8 and then to chlorophyll in

mass per leaf area using Equation 9 (Parry, Blonquist, & Bugbee, 2014). Equation 9 was

developed for barley crops; however, literature has shown that monocots have a similar

optical/absolute chlorophyll concentration relationship.

67.2*00119.01 SPADCCI (8)

(9)

Linking on-ground measurements to airborne imagery. Ground coordinates of

sampled chlorophyll coincided precisely with the location of the plants in the geo-

rectified imagery. Ground coordinates of the samples were overlaid onto the geo-rectified

imagery, and, using the ArcGIS spatial analyst tool (Extract Multi Values to Points), an

automated process was developed to extract the pixel value representing the center of

each sampled area.

Model Potential Inputs and Performance

Three of the four flights (early growth, mid growth, and early flowering),

excluding the flight 10 days after planting, were used in the dataset to train and test the

model. The dataset contains coincident in situ SPAD measurements used to derive

chlorophyll concentration and remote sensing reflectance measurements. All the data

were collected from inside the center pivot quarter (within field data); the zeros found in

)(146132)µmol.m( 43.0-2 CCIlChlorophyl

Page 40: The Application of Unmanned Aerial Vehicle to Precision

25

the data set represent the areas of no vegetation (center pivot wheels trajectory). A

statistical description of the dataset is presented in Table 2.2. Each pair of data consists of

a target, which is the chlorophyll concentration, and a set of 8 potential inputs tabulated

in Table 2.2. The potential inputs are composed of data retrieved form the UAS imagery

(VIS, NIR, TIR); vegetative indices (green model and NDVI) that were reported to be

sensitive in estimating chlorophyll (Gitelson, Vina, Ciganda, Rundquist, & Arkebauer,

2005; Shanahan et al., 2003); and LAI, a well-known and widely used vegetation index

related to crop growth. Table 2.3 shows the indices formulations.

Table 2.2

Statistical Description of the Dataset Used for Plant Chlorophyll Estimation

Input Range M ± SD Range M ± SD

Potential inputs

AggieAir inputs

Blue 0.11–0.36 0.15 ± 0.04 NIR 0.51–0.61 0.57 ± 0.02

Green 0.20–0.49 0.26 ± 0.05 Thermal (oC) 23.11–36.16 29.88 ± 4.13

Red 0.15–0.51 0.22 ± 0.07

Indices inputs

NDVI 0.00–0.78 0.44 ± 0.12 LAI (m2/m2) 0.00–4.23 2.41 ± 0.90

GM 0.17–1.74 1.17 ± 0.34

Target/output

Chlorophyll (μg.cm-2 ) 0.00–61.64 47.01 ± 11.26

Note. GM = green meter; LAI = leaf area index; NDVI = normalized difference vegetation index; NIR = near-infrared.

These potential predictors, exert to a certain degree correlation between each

other. This is because they are derived from the same AggieAir reflectance bands

Page 41: The Application of Unmanned Aerial Vehicle to Precision

26

(statistical correlation). While in customary statistics (e.g., linear regression) using these

predictors would raise issues, the Bayesian regression machine applied in this study can

deal with this problem. The kernel or basis function projects these potential inputs into a

higher dimensional space. The way these inputs are projected in the new dimensional

space, as well as the sparse representation of the observations in the final model, help the

model deal with collinearity issues.

Table 2.3

Vegetative Indices Formulation Used for Plant Chlorophyll Estimation

Indices Formula Reference

Green model RNIR / RGreen – 1 Gitelson et al., 2005

NDVI (RNIR – Rred) / (RNIR + RRed) Rouse, Haas, Schell, Deering, & Harlan, 1974

LAIa ln [(NDVI – NDVImax) / (NDVImin – NDVImax)] / –0.54

Duchemin et al., 2006; Smith, Bourgeois, Teillet, Freemantle, & Nadeau, 2008

Note. LAI = leaf area index; NDVI = normalized difference vegetation index. aLAI was calculated empirically and not validated by field measurements.

In preliminary runs different potential inputs were explored. For example, one set

composed of only single bands, and another set composed of the ratio of the single bands.

In addition, the authors tried vegetative indices sensitive to chlorophyll estimations

(transformed chlorophyll absorption reflectance index, modified chlorophyll absorption

ratio index, and modified triangular vegetation index) that were modified to adapt to the

spectral response of AggieAir sensors (e.g., replacing the required red edge by the NIR

band). However, details on these preliminary runs are not reported in this study because

of their low statistical performance.

Page 42: The Application of Unmanned Aerial Vehicle to Precision

27

The RVM is a well-established statistical learning algorithm that adopts a full

probabilistic framework. Its key feature is that it can yield a solution function that

depends on only a very small number of training samples (called relevance vectors). In

the RVM framework, the model is built on the few training examples whose associated

hyperparameters do not go to infinity during the training process, leading to a sparse

solution. The implemented RVM is based on the MATLAB code provided via Michael E.

Tipping’s website. The RVM model in this research was first trained and tested using K-

fold cross validation (K = 5); the cross validation technique is utilized to generalize an

independent training data set (Kohavi, 1995). In this procedure, the training set is

partitioned into K disjoint sets. The model is trained, for a chosen kernel, on all the

subsets except for one, which is left for testing. The procedure is repeated for a total of K

trials, each time using a different subset for testing. After the selection of the kernel

function and its width, the whole data set is trained using RVM based regression. The

advantage of this method over a random selection of training samples is that all

observations are used for either training (K times) or evaluation (once).

The model was developed with an input selection process (Guyon & Elisseeff,

2003) in an attempt to explain the data in the simplest way possible. Potential inputs were

examined to see which were most relevant to the target function and thus avoid degrading

the performance of a learning algorithm due to the presence of irrelevant input variables.

In each iteration, the input with the minimum efficiency was eliminated.

The RVM model was tested using six kernel types: Gauss, Laplace, spline,

Cauchy, thin plate spline, and bubble. The performance of the model was evaluated by

Page 43: The Application of Unmanned Aerial Vehicle to Precision

28

comparing the RMSE and the Nash-Sutcliffe efficiency (E); these two parameters have

been widely used to evaluate the performance of RVM models. The larger the value of E

and the smaller the value of RMSE, the greater the precision and accuracy of the model

to predict chlorophyll. The RMSE and E are computed as shown in Equations 10 and 11,

respectively:

N

yyRMSE

N

ttt

1

2)ˆ( (10)

N

t t

N

t t

yy

yyE

12

12

)(

)ˆ(1

(11)

where ŷt = predicted chlorophyll concentration; yt = measured chlorophyll concentration;

= mean of the observed chlorophyll concentration; = mean of the estimated

chlorophyll concentration; and N = total number of observations.

Result and Discussion

Each of the six kernel types was tested over a wide range of kernel widths (10-5-

105), and RMSE and E were calculated for all of the resulting models to assess their

predictive capabilities. An embedded loop in the coding model was developed to

represent the backward elimination tool. For each type of kernel and its corresponding

width, the RVM was first run using all of the 8 inputs, consequently generating all of the

needed statistical model performance estimates to assess the model. A set of defined

iterations then eliminated, in order, the input with the minimum efficiency, thus removing

y y

Page 44: The Application of Unmanned Aerial Vehicle to Precision

29

the input least relevant to the target function. After numerous computational runs, four

options presented themselves as potential “best model” scenarios (Table 2.4). All four of

these potential best model scenarios had an RMSE < 6 μg.cm-2 and an E > 0.7. In 94% of

all runs conducted across the six kernel types, the thermal band was the last input to be

dropped, suggesting that thermal imagery is an important input, at least in the case of

study area, possessing the most relevant information for estimating chlorophyll

concentration. Thermal data allowed the models to differentiate between the bare soil and

the different level of vegetation per pixel resulting in a more accurate chlorophyll

estimates. A preliminary interpretation could be the fact that oat leaves are very thin, with

minimal heat capacity, and as a result, leaves exposed to full sunlight can warm up

substantially above air temperature. This elevated temperature can help identify greener

leaves and as a result those with higher chlorophyll values. Nevertheless, additional

experiments that explore thermal imagery and its effect on chlorophyll estimations need

to be conducted.

Table 2.4

Potential “Best Scenarios”

Model Kernel type # of inputs Inputs

1 Gaussian 4 LAI, NDVI, thermal, green

2 Gaussian 3 Thermal, green, LAI

3 Laplace 4 NDVI, red, green, thermal

4 Thin plate spline 4 LAI, NDVI, thermal, red Note. LAI = leaf area index; NDVI = normalized difference vegetation index.

Page 45: The Application of Unmanned Aerial Vehicle to Precision

30

When plotting the 1-1 plot for the four best scenario candidates, the plots looked

almost identical. Since the statistical performance does not reveal an absolute best model,

visual comparison was made of the chlorophyll estimation maps of the four models, on

one hand, and the NDVI, LAI, and true-color maps, on the other hand. The chlorophyll

estimates for the early growth, mid growth, and early flowering images were developed

considering the unique characteristic of each of the four best models (kernel type, width,

and set of inputs). Models 1 and 2 showed clear overfitting when plotted over the entire

map: In each case, the resulting map was one solid color, with no variation in estimated

chlorophyll between bare soil and fully grown oat plants. Model 3 showed more variation

within the field; nevertheless, visual comparisons with Model 4 indicated that Model 4

was superior. Model 4 showed an RSME of 5.31 μg.cm-2, an E of 0.76, and 9 relevance

vectors. Figure 2.5 illustrates the measured chlorophyll concentration versus estimated

values with a one standard error confidence interval. The three flights are separated by

the yellow line in the graph. Some differences can be observed between the different

dates; the estimates for the first and third flights are more precise than the second flight.

This could be due to the stage of the crop growth or the homogeneity of the vegetation

cover.

Figure 2.6 represents regression diagnostic plots of Model 4 that address model

assumptions like linearity and equality of variances. The 1:1 plot confirms the adequacy

of the model proposed for most of the chlorophyll values lying within the boundaries of

± 1.0 SPAD unit (sensor accuracy), which corresponds to 14 μg.cm-2. The chlorophyll

maps generated from Model 4, along with NDVI and LAI, are presented in Figure 2.7.

Page 46: The Application of Unmanned Aerial Vehicle to Precision

31

Figure 2.5. Measured versus predicted chlorophyll concentration in the three flights for Model 4. Vertical yellow lines separate the three flight dates.

Figure 2.6. Model 4: Residual plot of the three flights (left) and one-by-one plot excluding the bare soil–zero chlorophyll points and reflecting the chlorophyll meter accuracy (right).

Page 47: The Application of Unmanned Aerial Vehicle to Precision

32

As shown in Figure 2.7, the predicted chlorophyll concentration maps show a

visual good agreement with the LAI and NDVI maps. In the early growth image, the field

exterior had weeds growing in it, which explains the predicted chlorophyll concentration

values. This area was not irrigated during the growing cycle, leaving the weeds to dry and

senescence, thus a near-zero chlorophyll concentration value was assigned by the model

in the following two images. Also, the wheel tracks and the access road that are located

around the center pivot had no vegetation cover, and the model successfully assigned a

near-zero chlorophyll concentration to these features.

Figure 2.7. True-color maps, normalized difference vegetation index (NDVI) maps, leaf area index (LAI) maps, and the estimated chlorophyll concentration (μg.cm-2) map for the three different dates representing early growth, mid growth, and early flowering.

Early growth

Mid growth

Early flowering

Page 48: The Application of Unmanned Aerial Vehicle to Precision

33

Another common pattern was the two thick horizontal and vertical lines that

protrude in the images. These were past ditch lines that had been used in flood irrigation

activities prior to the conversion of the field to a center pivot system. The greater water

content in those areas caused the plants growing along those two lines to be very

vigorous. This is reflected in the high chlorophyll concentration values given to the plants

in this area.

Chlorophyll concentration varies widely within the growing season, and therefore

any recommended analytical technique must perform well under unseen data. To explore

the model with unseen data, the May 16 flight was used. Now that the model was

established with a defined set of features (inputs, kernel type, kernel width), data from the

May 16 flight (10 days after planting of the oats) were entered in the model to explore the

model’s performance when subjected to totally unseen data. The predicted chlorophyll

concentration map is shown in Figure 2.8.

Figure 2.8. True-color image of flight zero (left), leaf area index (LAI) map (middle), and estimated chlorophyll concentration map (μg.cm-2, right).

Page 49: The Application of Unmanned Aerial Vehicle to Precision

34

Again, the predicted chlorophyll concentration map for the fourth flight showed

good association with the NDVI map. Areas of vigorous growth, bare soil, and low

vegetation were similar in the three maps and represented similar growth patterns. This

test reported an RSME of 8.52 μg.cm-2 and E of 0.71 for this flight. This result showed

that the model successfully performed when given unseen data.

Despite the complexity of the statistical model included in this paper, it is

anticipated that the lucid output (chlorophyll concentration maps) will help agricultural

decision makers quantify field chlorophyll and address its variability and as a result

improve input efficiency, environmental sustainability, and yield. Adoption of precision

agriculture is likely to continue into the foreseeable future. However, studies that explore

high-resolution sensors (< 1 m) with adequate frequent coverage, combined with

techniques capable of extracting information from imagery to provide near real-time

information, will be a determining factor in the adoption rate of precision agriculture.

Conclusion

This paper presented the application of imagery from AggieAir, a remote sensing

platform, combined with machine learning algorithms (RVM) to estimate chlorophyll

concentration as an important biophysical parameter to be used in precision agriculture.

The RVM modeling technique, coupled with cross validation and backward elimination,

was applied to a data set composed of reflectance from high-resolution multispectral

imagery (VIS-NIR), TIR imagery, and vegetative indices, in conjunction with in situ

chlorophyll concentrations derived from SPAD measurements. Six kernel types were

tested over a wide range of kernel widths. Model performance was evaluated by

Page 50: The Application of Unmanned Aerial Vehicle to Precision

35

comparing the RMSE and E of various models and later by visual comparison.

Chlorophyll concentration estimation was best achieved with Model 4 (kernel type: thin

plate spline; kernel width: 5.4; selected inputs: LAI, NDVI, thermal, and red band;

RSME: 5.31 μg.cm-2; E: 0.76; and 9 relevance vectors) for the three flights. Of all the

inputs, thermal band was retained last in 94% of the models, proving the significance of

thermal imagery as an input possessing the most relevant information in estimating

chlorophyll concentration.

Converting these chlorophyll estimate maps into actionable information to benefit

the end user now shows promise. Other research that estimates soil moisture, actual

evapotranspiration, and soil nutrient content using the same high-resolution aerial

platforms allows for wider adoption of precision agriculture by future farmers. Although

the results presented in this section are arguably not yet actionable, maps like these could

be used to quantify plant health, predict yield, and indicate where and how much fertilizer

to apply.

AggieAir imagery, combined with appropriate analytic tools, allows spatial

estimation of chlorophyll content. These estimates, made at such fine resolutions in space

and time, can aid farmers in assessing the heterogeneity of their fields and subsequently

implement needed actions accordingly. The high-resolution spatial information generated

from AggieAir imagery could enable far greater precision in the application of nitrogen

fertilizers and water through a center pivot irrigation system.

Page 51: The Application of Unmanned Aerial Vehicle to Precision

36

Acknowledgment

This research was supported by the Utah Water Research Laboratory, the Provo,

Utah Office of the U.S. Bureau of Reclamation, and the AggieAir Flying Circus at the

Utah Water Research Laboratory. The authors appreciate the support the Utah State

University AggieAir team that helped in the data collection procedure. Special thanks

goes to the Division of Research Computing in the Office of Research at Utah State

University. Appreciation goes to Mrs. Carri Richards for her timely help in editing the

paper.

References

Bachour, R., Walker, W., Ticlavilca, A., McKee, M., & Maslova, I. (2014). Estimation of spatially distributed evapotranspiration using remote sensing and a relevance vector machine. Journal of Irrigation and Drainage Engineering, 140(8), 04014029. doi:10.1061/(ASCE)IR.1943-4774.0000754

Baret, F., & Buis, S. (2008). Estimating canopy characteristics from remote sensing observations: Review of methods and associated problems. In S. Liang (Ed.), Advances in land remote sensing (pp. 173-201). Dordrecht, The Netherlands: Springer.

Bauer, M. E. (1985). Spectral inputs to crop identification and condition assessment. Proceedings of the IEEE, 73, 1071-1085.

Benedetti, R., & Rossini, P. (1993). On the use of NDVI profiles as a tool for agricultural statistics: The case study of wheat yield estimate and forecast in Emilia Romagna. Remote Sensing of Environment, 45, 311-326.

Berni, J., Zarco-Tejada, P. J., Suárez, L., & Fereres, E. (2009). Thermal and narrowband multispectral remote sensing for vegetation monitoring from an unmanned aerial vehicle. IEEE Transactions on Geoscience and Remote Sensing, 47, 722-738.

Berry, J., & Bjorkman, O. (1980). Photosynthetic response and adaptation to temperature in higher plants. Annual Review of Plant Physiology, 31(1), 491-543.

Page 52: The Application of Unmanned Aerial Vehicle to Precision

37

Blackburn, G. A. (1998). Quantifying chlorophylls and caroteniods at leaf and canopy scales: An evaluation of some hyperspectral approaches. Remote Sensing of Environment, 66, 273-285.

Blum, A., Mayer, J., & Gozlan, G. (1982). Infrared thermal sensing of plant canopies as a screening technique for dehydration avoidance in wheat. Field Crops Research, 5, 137-146.

Brisco, B., Brown, R. J., Hirose, T., McNairn, H., & Staenz, K. (1998). Precision agriculture and the role of remote sensing: A review. Canadian Journal of Remote Sensing, 24, 315-327.

Broge, N. H., & Leblanc, E. (2001). Comparing prediction power and stability of broadband and hyperspectral vegetation indices for estimation of green leaf area index and canopy chlorophyll density. Remote Sensing of Environment, 76, 156-172.

Camps-Valls, G., Bruzzone, L., Rojo-Álvarez, J. L., & Melgani, F. (2006). Robust support vector regression for biophysical variable estimation from remotely sensed images. Geoscience and Remote Sensing Letters, IEEE, 3(3), 339-343.

Camps-Valls, G., Gómez-Chova, L., Muñoz-Marí, J., Vila-Francés, J., Amorós-López, J., & Calpe-Maravilla, J. (2006). Retrieval of oceanic chlorophyll concentration with relevance vector machines. Remote Sensing of Environment, 105(1), 23-33.

Carter, G. A., & Spiering, B. A. (2002). Optical properties of intact leaves for estimating chlorophyll concentration. Journal of Environmental Quality, 31, 1424-1432.

Chaerle, L., & Van Der Straeten, D. (2000). Imaging techniques and the early detection of plant stress. Trends in Plant Science, 5(11), 495-501.

Cipollini, P., Corsini, G., Diani, M., & Grasso, R. (2001). Retrieval of sea water optically active parameters from hyperspectral data by means of generalized radial basis function neural networks. IEEE Transactions on Geoscience and Remote Sensing, 39, 1508-1524.

Clemens, S. R. (2012). Procedures for correcting digital camera imagery acquired by the AggieAir remote sensing platform. Thesis, Utah State University, Logan.

Cristianini, N., & Shawe-Taylor, J. (2000). An introduction to support vector machines and other kernel-based learning methods. New York, NY: Cambridge University Press.

Crowther, B. G. (1992). Radiometric calibration of multispectral video imagery. Unpublished doctoral dissertation, Utah State University, Logan.

Page 53: The Application of Unmanned Aerial Vehicle to Precision

38

Daberkow, S. G., McBride, W. D., Robert, P. C., Rust, R. H., & Larson, W. E. (2000). Adoption of precision agriculture technologies by U.S. farmers. In P. C. Robert, R. H. Rust, & W. E. Larson (Eds.), Proceedings of the 5th International Conference on Precision Agriculture, Bloomington, Minnesota, 16-19 July, 2000 (pp. 1-12). Bloomington, MN: American Society of Agronomy.

Datt, B. (1999). A new reflectance index for remote sensing of chlorophyll content in higher plants: Tests using eucalyptus leaves. Journal of Plant Physiology, 154(1), 30-36.

Daughtry, C. S. T., Walthall, C. L., Kim, M. S., De Colstoun, E. B., & McMurtrey, J. E., III. (2000). Estimating corn leaf chlorophyll concentration from leaf and canopy reflectance. Remote Sensing of Environment, 74, 229-239.

Dawson, T. P., North, P. R. J., Plummer, S. E., & Curran, P. J. (2003). Forest ecosystem chlorophyll content: Implications for remotely sensed estimates of net primary productivity. International Journal of Remote Sensing, 24, 611-617.

De Martino, M., Mantero, P., Serpico, S. B., Carta, E., Corsini, G., & Grasso, R. (2002). Water quality estimation by neural networks based on remotely sensed data analysis. In Proceedings of the International Workshop on Geo-Spatial Knowledge Processing for Natural Resource Management (pp. 54-58).

Demarez, V., & Gastellu-Etchegorry, J. P. (2000). A modeling approach for studying forest chlorophyll content. Remote Sensing of Environment, 71, 226-238.

Demir, B., & Erturk, S. (2007). Hyperspectral image classification using relevance vector machines. Geoscience and Remote Sensing Letters, IEEE, 4(4), 586-590.

Duchemin, B., Hadria, R., Erraki, S., Boulet, G., Maisongrande, P., Chehbouni, A., . . . Simonneaux, V. (2006). Monitoring wheat phenology and irrigation in Central Morocco: On the use of relationships between evapotranspiration, crops coefficients, leaf area index and remotely-sensed vegetation indices. Agricultural Water Management, 79(1), 1-27.

Evans, J. R. (1989). Photosynthesis and nitrogen relationships in leaves of C3 plants. Oecologia, 78(1), 9-19.

Franke, J., & Menz, G. (2007). Multi-temporal wheat disease detection by multi-spectral remote sensing. Precision Agriculture, 8(3), 161-172.

Genton, M. G. (2002). Classes of kernels for machine learning: A statistics perspective. The Journal of Machine Learning Research, 2, 299-312.

Page 54: The Application of Unmanned Aerial Vehicle to Precision

39

Gitelson, A. A., Kaufman, Y. J., & Merzlyak, M. N. (1996). Use of a green channel in remote sensing of global vegetation from EOS-MODIS. Remote Sensing of Environment, 58, 289-298.

Gitelson, A., & Merzlyak, M. N. (1994). Quantitative estimation of chlorophyll-a using reflectance spectra: Experiments with autumn chestnut and maple leaves. Journal of Photochemistry and Photobiology B: Biology, 22, 247-252.

Gitelson, A. A., Viña, A., Ciganda, V., Rundquist, D. C., & Arkebauer, T. J. (2005). Remote estimation of canopy chlorophyll content in crops. Geophysical Research Letters, 32(8). doi:10.1029/2005GL022688

Gitelson, A. A., Viña, A., Verma, S. B., Rundquist, D. C., Arkebauer, T. J., Keydan, G., . . . Suyker, A. E. (2006). Relationship between gross primary production and chlorophyll content in crops: Implications for the synoptic monitoring of vegetation productivity. Journal of Geophysical Research: Atmospheres (1984–2012), 111(D8).

Gómez-Chova, L., Muñoz-Marí, J., Laparra, V., Malo-López, J., & Camps-Valls, G. (2011). A review of kernel methods in remote sensing data analysis. In S. Prasad, L. M. Bruce, & J. Chanussot (Eds.), Optical remote sensing (pp. 171-206). Berlin, Germany: Springer-Verlag Berlin Heidelberg.

González Vilas, L., Spyrakos, E., & Torres Palenzuela, J. M. (2011). Neural network estimation of chlorophyll a from MERIS full resolution data for the coastal waters of Galician rias (NW Spain). Remote Sensing of Environment, 115, 524-535.

Guyon, I., & Elisseeff, A. (2003). An introduction to variable and feature selection. The Journal of Machine Learning Research, 3, 1157-1182.

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.

Haboudane, D., Miller, J. R., Tremblay, N., Zarco-Tejada, P. J., & Dextraze, L. (2002). Integrated narrow-band vegetation indices for prediction of crop chlorophyll content for application to precision agriculture. Remote Sensing of Environment, 81, 416-426.

Hassan Esfahani, L., Torres-Rua, A., Jensen A., & McKee, M. (2014a). Assessment of surface soil moisture using high-resolution multi-spectral imagery and artificial neural networks. Remote Sensing, 7, 2627-2646.

Hassan Esfahani, L., Torres-Rua, A., Jensen A., & McKee, M. (2014b). Top soil moisture estimation for precision agriculture using unmmaned aerial vehicle multispectral

Page 55: The Application of Unmanned Aerial Vehicle to Precision

40

imagery. In Geoscience and Remote Sensing Symposium (IGARSS), 2014 IEEE International (pp. 3263-3266). doi:10.1109/IGARSS.2014.6947175

Hastie, T., Tibshirani, R., & Friedman, J. (Eds.). (2009). The elements of statistical learning: Data mining, inference, and prediction (2nd ed.). New York, NY: Springer.

Idso, S. B., Jackson, R. D., & Reginato, R. J. (1977). Remote-sensing of crop yields. Science, 196(4285), 19-25.

Infrared Cameras, Incorporated. (2014). Home page. Retrieved from http://www .infraredcamerasinc.com

Jackson, R. D. (1984). Remote sensing of vegetation characteristics for farm management. In 1984 Technical Symposium East (pp. 81-97). Bellingham, WA: International Society for Optics and Photonics.

Jackson, R. D., Idso, S. B., Reginato, R. J., & Pinter, P. J. (1981). Canopy temperature as a crop water stress indicator. Water Resources Research, 17, 1133-1138

Jensen, A. M. (2014). Innovative payloads for small unmanned aerial system-based personal remote sensing and applications. Doctoral dissertation, Utah State University, Logan.

Jiang, J., Liu, D., Gu, J., & Susstrunk, S. (2013). What is the space of spectral sensitivity functions for digital color cameras? In 2013 IEEE Workshop on Applications of Computer Vision (WACV) (pp. 168-179). doi:10.1109/WACV.2013.6475015

Johnson, L. F., Hlavka, C. A., & Peterson, D. L. (1994). Multivariate analysis of AVIRIS data for canopy biochemical estimation along the Oregon transect. Remote Sensing of Environment, 47, 216-230.

Kalluri, S., Gilruth, P., Bergman, R., & Plante, R. (2002). Impacts of NASA’s remote sensing data on policy and decision making at state and local agencies in the United States. In Geoscience and Remote Sensing Symposium, 2002. IGARSS'02. 2002 IEEE International (Vol. 3, pp. 1691-1693). doi:10.1109/IGARSS.2002 .1026223

Kim, M. S. (1994). The use of narrow spectral bands for improving remote sensing estimation of fractionally absorbed photosynthetically active radiation (fAPAR). Master’s thesis, University of Maryland, College Park.

Kim, M. S., Daughtry, C. S. T., Chappelle, E. W., McMurtrey, J. E., III, & Walthall, C. L. (1994). The use of high spectral resolution bands for estimating absorbed photosynthetically active radiation (APAR). In Proceedings of the 6th Symposium

Page 56: The Application of Unmanned Aerial Vehicle to Precision

41

on Physical Measurements and Signatures in Remote Sensing, January 17–21, 1994, Val D’Isere, France (pp. 299–306).

Knudby, A., LeDrew, E., & Brenning, A. (2010). Predictive mapping of reef fish species richness, diversity and biomass in Zanzibar using IKONOS imagery and machine-learning techniques. Remote Sensing of Environment, 114, 1230-1241.

Kohavi, R. (1995). A study of cross-validation and bootstrap for accuracy estimation and model selection. In IJCAI’95: Proceedings of the 14th International Joint Conference on Artificial Intelligence (Vol. 2, pp. 1137-1145). San Francisco, CA: Kaufmann.

Lamb, D. W., & Brown, R. B. (2001). PA—Precision agriculture: Remote-sensing and mapping of weeds in crops. Journal of Agricultural Engineering Research, 78(2), 117-125.

Lambert, D., & Lowenberg-De Boer, J. (2000). Precision agriculture profitability review. Retrieved from Purdue University: http://agriculture.purdue.edu/SSMC/Frames /newsoilsX.pdf

le Maire, G., François, C., & Dufrêne, E. (2004). Towards universal broad leaf chlorophyll indices using PROSPECT simulated database and hyperspectral reflectance measurements. Remote Sensing of Environment, 89, 1-28.

Liaghat, S., & Balasundram, S. K. (2010). A review: The role of remote sensing in precision agriculture. American Journal of Agricultural and Biological Sciences, 5(1), 50-55.

MacDonald, R. B., & Hall, F. G. (1980). Global crop forecasting. Science, 208(4445), 670-679.

Mathur, A., & Foody, G. M. (2008). Crop classification by support vector machine with intelligently selected training data for an operational application. International Journal of Remote Sensing, 29, 2227-2240.

Miller, J. R., Hare, E. W., & Wu, J. (1990). Quantitative characterization of the vegetation red edge reflectance 1. An inverted-Gaussian reflectance model. Remote Sensing, 11, 1755-1773.

Miura, T., & Huete, A. R. (2009). Performance of three reflectance calibration methods for airborne hyperspectral spectrometer data. Sensors, 9, 794-813.

Moran, J. A., Mitchell, A. K., Goodmanson, G., & Stockburger, K. A. (2000). Differentiation among effects of nitrogen fertilization treatments on conifer seedlings by foliar reflectance: A comparison of methods. Tree Physiology, 20, 1113-1120.

Page 57: The Application of Unmanned Aerial Vehicle to Precision

42

Moran, M. S., Inoue, Y., & Barnes, E. M. (1997). Opportunities and limitations for image-based remote sensing in precision crop management. Remote Sensing of Environment, 61, 319-346.

MosaicMill. (2009). EnsoMOSAIC image processing user’s guide. Version 7.3. Vantaa, Finland: Author.

Moser, G., & Serpico, S. B. (2009). Automatic parameter optimization for support vector regression for land and sea surface temperature estimation from remote sensing data. IEEE Transactions on Geoscience and Remote Sensing, 47(3), 909-921.

Murakami, Y., Tsuyama, M., Kobayashi, Y., Kodama, H., & Iba, K. (2000). Trienoic fatty acids and plant tolerance of high temperature. Science, 287(5452), 476-479.

Myneni, R. B., Maggion, S., Iaquinta, J., Privette, J. L., Gobron, N., Pinty, B., . . . Williams, D. L. (1995). Optical remote sensing of vegetation: Modeling, caveats, and algorithms. Remote Sensing of Environment, 51, 169-188.

Neale, C. M., & Crowther, B. G. (1994). An airborne multispectral video/radiometer remote sensing system: Development and calibration. Remote Sensing of Environment, 49, 187-194.

Niinemets, Ü. & Tenhunen, J. D. (1997). A model separating leaf structural and physiological effects on carbon gain along light gradients for the shade‐ tolerant species Acer saccharum. Plant, Cell & Environment, 20, 845-866.

Pal, M., & Mather, P. M. (2005). Support vector machines for classification in remote sensing. International Journal of Remote Sensing, 26, 1007-1011.

Parry, C., Blonquist, J., & Bugbee, B. (2014). In situ measurement of leaf chlorophyll concentration: Analysis of the optical/absolute relationship. Plant, Cell & Environment, 37, 2508-2520.

Peñuelas, J., & Filella, I. (1998). Visible and near-infrared reflectance techniques for diagnosing plant physiological status. Trends in Plant Science, 3(4), 151-156.

Pierce, F. J., & Nowak, P. (1999). Aspects of precision agriculture. Advances in Agronomy, 67, 1-85.

Raison, J. K., Roberts, J. K., & Berry, J. A. (1982). Correlations between the thermal stability of chloroplast (thylakoid) membranes and the composition and fluidity of their polar lipids upon acclimation of the higher plant, Nerium oleander, to growth temperature. Biochimica et Biophysica Acta (BBA)—Biomembranes, 688(1), 218-228.

Page 58: The Application of Unmanned Aerial Vehicle to Precision

43

Rouse, J. W., Jr., Haas, R. H., Schell, J. A., Deering, D. W., & Harlan, J. C. (1974). Monitoring the vernal advancements and retrogradation (green wave effect) of natural vegetation. Greenbelt, MD: NASA.

Salvucci, M. E., & Crafts‐ Brandner, S. J. (2004). Inhibition of photosynthesis by heat stress: The activation state of Rubisco as a limiting factor in photosynthesis. Physiologia Plantarum, 120, 179-186.

Schellberg, J., Hill, M. J., Gerhards, R., Rothmund, M., & Braun, M. (2008). Precision agriculture on grassland: Applications, perspectives and constraints. European Journal of Agronomy, 29(2), 59-71.

Seelan, S. K., Laguette, S., Casady, G. M., & Seielstad, G. A. (2003). Remote sensing applications for precision agriculture: A learning community approach. Remote Sensing of Environment, 88, 157-169.

Shanahan, J. F., Holland, K. H., Schepers, J. S., Francis, D. D., Schlemmer, M. R., & Caldwell, R. (2003). Use of a crop canopy reflectance sensor to assess corn leaf chlorophyll content. In Digital imaging and spectral techniques: Applications to precision agriculture and crop physiology (pp. 135-150). Madison, WI: American Society of Agronomy.

Shanahan, J. F., Schepers, J. S., Francis, D. D., Varvel, G. E., Wilhelm, W. W., Tringe, J. M., . . . Major, D. J. (2001). Use of remote-sensing imagery to estimate corn grain yield. Agronomy Journal, 93, 583-589.

Sharkey, T. D. (2005). Effects of moderate heat stress on photosynthesis: Importance of thylakoid reactions, Rubisco deactivation, reactive oxygen species, and thermos tolerance provided by isoprene. Plant, Cell & Environment, 28, 269-277.

Sims, D. A., & Gamon, J. A. (2002). Relationships between leaf pigment content and spectral reflectance across a wide range of species, leaf structures and developmental stages. Remote Sensing of Environment, 81, 337-354.

Smith, A. M., Bourgeois, G., Teillet, P. M., Freemantle, J., & Nadeau, C. (2008). A comparison of NDVI and MTVI2 for estimating LAI using CHRIS imagery: a case study in wheat. Canadian Journal of Remote Sensing, 34, 539-548.

Souza, C. R. (2010). Kernel functions for machine learning applications. Retrieved from http://crsouza.blogspot.com/2010/03/kernel-functions-for-machine-learning.html

Stone, M. L., Solie, J. B., Raun, W. R., Whitney, R. W., Taylor, S. L., & Ringer, J. D. (1996). Use of spectral radiance for correcting in-season fertilizer nitrogen deficiencies in winter wheat. Transactions of the ASAE, 39, 1623-1631.

Page 59: The Application of Unmanned Aerial Vehicle to Precision

44

Swinton, S. M., & Lowenberg-DeBoer, J. (1998). Evaluating the profitability of site-specific farming. Journal of Production Agriculture, 11, 439-446.

Thayananthan, A., Navaratnam, R., Stenger, B., Torr, P. H., & Cipolla, R. (2008). Pose estimation and tracking using multivariate regression. Pattern Recognition Letters, 29, 1302-1310.

Ticlavilca, A. M., McKee, M., & Walker, W. R. (2013). Real-time forecasting of short-term irrigation canal demands using a robust multivariate Bayesian learning model. Irrigation Science, 31(2), 151-167.

Tipping, M. E. (2001). Sparse Bayesian learning and the relevance vector machine. The Journal of Machine Learning Research, 1, 211-244.

Tipping, M. E. (2004). Bayesian inference: An introduction to principles and practice in machine learning. In O. Bousquet, U. von Luxburg, & G. Ratsch (Eds.), Advanced lectures on machine learning (pp. 41-62). Berlin, Germany: Springer Berlin Heidelberg.

Tipping, M. E., & Faul, A. C. (2003). Fast marginal likelihood maximisation for sparse Bayesian models. In C. M. Bishop & B. J. Frey (Eds.), Proceedings of the Ninth International Workshop on Artificial Intelligence and Statistics (Vol. 1, No. 3). Society for Artificial Intelligence and Statistics.

Torres, A. F., Walker, W. R., & McKee, M. (2011). Forecasting daily potential evapotranspiration using machine learning and limited climatic data. Agricultural Water Management, 98, 553-562.

Vapnik, V. (2000). The nature of statistical learning theory. Dordrecht, The Netherlands: Springer Science & Business Media.

Verrelst, J., Alonso, L., Camps-Valls, G., Delegido, J., & Moreno, J. (2012). Retrieval of vegetation biophysical parameters using Gaussian process techniques. IEEE Transactions on Geoscience and Remote Sensing, 50, 1832-1843.

Verrelst, J., Schaepman, M. E., Malenovský, Z., & Clevers, J. G. (2010). Effects of woody elements on simulated canopy reflectance: Implications for forest chlorophyll content retrieval. Remote Sensing of Environment, 114, 647-656.

Wallace, L., Lucieer, A., Watson, C., & Turner, D. (2012). Development of a UAV-LiDAR system with application to forest inventory. Remote Sensing, 4, 1519-1543.

Weis, E. (1981). Reversible heat-inactivation of the Calvin cycle: A possible mechanism of the temperature regulation of photosynthesis. Planta, 151(1), 33-39.

Page 60: The Application of Unmanned Aerial Vehicle to Precision

45

Wood, C. W., Tracy, P. W., Reeves, D. W., & Edmisten, K. L. (1992). Determination of cotton nitrogen status with a handheld chlorophyll meter. Journal of Plant Nutrition, 15, 1435-1448.

Yadawa, U. L. (1986). A rapid and nondestructive method to determine chlorophyll in intact leaves. HortScience, 21, 1449-1450.

Yoder, B. J., & Pettigrew-Crosby, R. E. (1995). Predicting nitrogen and chlorophyll content and concentrations from reflectance spectra (400–2500 nm) at leaf and canopy scales. Remote Sensing of Environment, 53, 199-211.

Yuan, J., Wang, K., Yu, T., & Fang, M. (2007). Integrating relevance vector machines and genetic algorithms for optimization of seed-separating process. Engineering Applications of Artificial Intelligence, 20, 970-979.

Zaman, B., Jensen, A., Clemens, S. R., & McKee, M. (2014). Retrieval of spectral reflectance of high resolution multispectral imagery acquired with an autonomous unmanned aerial vehicle. Photogrammetric Engineering & Remote Sensing, 80, 1139-1150.

Zarco‐ Tejada, P. J., & Miller, J. R. (1999). Land cover mapping at BOREAS using red edge spectral parameters from CASI imagery. Journal of Geophysical Research: Atmospheres (1984–2012), 104(D22), 27921-27933.

Zarco-Tejada, P. J., Miller, J. R., Morales, A., Berjón, A., & Agüera, J. (2004). Hyperspectral indices and model simulation for chlorophyll estimation in open-canopy tree crops. Remote Sensing of Environment, 90, 463-476.

Zarco-Tejada, P. J., Miller, J. R., Noland, T. L., Mohammed, G. H., & Sampson, P. H. (2001). Scaling-up and model inversion methods with narrowband optical indices for chlorophyll content estimation in closed forest canopies with hyperspectral data. IEEE Transactions on Geoscience and Remote Sensing, 39, 1491-1507.

Page 61: The Application of Unmanned Aerial Vehicle to Precision

46

CHAPTER 3

USE OF HIGH RESOLUTION MULTISPECTRAL IMAGERY TO ESTIMATE

PLANT NITROGEN IN OATS (AVENA SATIVA).

Abstract

Remote sensing applications for precision agriculture depend on acquiring

actionable information at high spatial resolution and at a temporal frequency appropriate

for timely responses. This study was conducted to determine if high resolution canopy

reflectance could be used to evaluate leaf Nitrogen (N) status in oats (Avena sativa) for

precision agriculture. An unmanned aerial system platform named AggieAirTM, was used

to acquire high-resolution imagery in the visual, near infrared, and thermal infrared

spectra. This study reports the development of a relevance vector machine (RVM)

coupled with forward input selection process to a dataset composed of reflectance from

high-resolution multispectral imagery (visible, near-infrared, thermal infrared imagery),

and vegetative indices, in conjunction with in situ leaf N sampled and analyzed in the

laboratory, to estimate leaf N content from remotely sensed data at 15 cm resolution. The

results indicate that an RVM with a Gaussian kernel type and kernel width of 0.28,

having Near-infrared, green ratio vegetation index, red/green ratio index, and simple ratio

as the selected set of inputs, can be used to spatially estimate Leaf Nitrogen with a root

mean-squared error (RMSE) of 0.48 mg/100 mg dry tissue (DT), efficiency of 0.76, and 6

relevance vectors.

Page 62: The Application of Unmanned Aerial Vehicle to Precision

47

Keywords: Remote Sensing, High Spatial Resolution Imagery, Relevance Vector

Machine, Precision Agriculture, Leaf Nitrogen

Introduction

Monitoring crops and assessing their nutrient status throughout the vegetation

growth is a fundamental practice in precision agriculture. An essential nutrient for plant

growth is nitrogen (N), and when absent a significant decrease in both the photosynthetic

and CO2 assimilation capacity in crops is reported (Prsa, Stampar, Vodnik, & Veberic,

2007; Tracy, Hefner, Wood, & Edmisten, 1992). Farmers must balance the competing

goals of supplying enough N to their crops to reach their production goals while

minimizing the loss of N to the environment (Daughtry et al., 2000). When applied in

excess, N represents a threat to water quality because it can increase the chance of nitrate

contamination of surface and groundwater (Errebhi, Rosen, Gupta, & Birong, 1998; Ju,

Kou, Zhang, & Christie, 2006). Over-application of N fertilizer can also be an economic

loss to a farmer. Precise plant N evaluation has the potential to aid farmers in creating

this balance, by early detection of N stress and matching N supply with crop N

requirement at the correct rate, place, and time.

Traditionally, many methods have been utilized for determining crop N status,

including tissue analysis performed on leaves sampled from the field (Roth, Fox, &

Marshall, 1989). This method is known to be destructive, sample intensive, expensive,

and time consuming, with a significant lag between collecting the samples, performing

the laboratory analysis, and generating the final recommendation. This delay can result in

Page 63: The Application of Unmanned Aerial Vehicle to Precision

48

a missed opportunity in providing the crop with the needed nutrient in a timely fashion

and consequently influence the crop productivity. Recently, portable optical instruments

like chlorophyll meters are used to assess N status in crops (Filella, Serrano, Peñuelas,

1995; Han, Hedrickson, & Ni, 2001). These meters are less invasive, nondestructive, and

relatively faster in providing results (Gianquinto, Sambo, & Pimpini, 2003; Schepers,

Francis, Vigil, & Below, 1992; Turner & Jund, 1991). However, they are only practical at

leaf level and limited to evaluating plant status in small areas. This technique measures

the transmittance of radiation through a leaf in two wavelength bands centered near 650

nm and 940 nm (Peterson, Blackmer, Francis, & Schepers, 1993). The meter is actually

measuring leaf chlorophyll, and since the majority of leaf N is contained in chlorophyll

molecules, N content is deducted indirectly from those readings. However, some

researchers challenged the general polynomial equation commonly used to relate N to

measured chlorophyll and proposed an exponential equation that forces a more

appropriate fit (Markwell, Osterman, & Mitchell, 1995). Others commented that at high

N levels, N could be found in the NO3- N form and not only in chlorophyll molecules;

thus, the relationship between leaf chlorophyll and leaf N concentration may be nonlinear

(Wood, Reeves, & Himelrick, 1993). The successful usage of these meters when

assessing N content is also affected by the crop type, growth stages, measurement

positions on leaves, and environmental conditions (Chapman & Barreto, 1997; Ramesh et

al., 2002; Schepers et al., 1992; Turner & Jund, 1991). Both of these methods can

potentially lead to over- or underassessment of N status in the crops and as a result

inaccurate recommendations.

Page 64: The Application of Unmanned Aerial Vehicle to Precision

49

The importance and interest in leaf N and its spatial variation suggest a need to

analyze it in a spatially explicit and extensive manner. For instance, a leaf N content map

displaying spatial variations at canopy level could facilitate diagnosis of nutrient

availability and limitation and help to eliminate the need for intensive field sampling or

nutrient addition. Obtaining such data for spatial representation could be achieved using

remote sensing platforms. Remote sensing provides rapid, quantitative information about

crops and, above all, does so nondestructively for diagnosing the spatial variability of

crop field properties. Remotely sensed data infer canopy leaf N status based on leaf

reflectance and transmittance measurements.

An early study by Thomas and Oerther (1972) found that reflectance in the green

and red regions of the electromagnetic spectrum were highly correlated to leaf N content

determined using the Kjeldahl method. Additional research reported by Martin, Shenk,

and Barton (1989) established the relationship between crop N content and canopy

reflectance at the visible (VIS, 400–700 nm) and near-infrared (NIR, 700–1100 nm)

regions. In addition, light reflectance at wavelengths near 550 nm from individual leaves

was found to be a good indicator of N stress in corn crops (Blackmer & Schepers, 1995).

Researchers also developed various combinations of bands knows as vegetative indices

that are linearly related to N concentration, such as the Red-edge Index, R740/R720

(Mokhele & Ahmed, 2010). Canopy reflectance in the red and NIR regions have reported

success at determining crop N content, due to the ability to detect changes associated

with chlorophyll content (Guyot, 1991). Researches have studied leaf N reflectance

response on corn (Blackmer, Schepers, & Varvel, 1994: Lee, Searcy, & Kataoka, 1999),

Page 65: The Application of Unmanned Aerial Vehicle to Precision

50

sweet pepper (Thomas & Oerther, 1972), sugarcane (Miphokasap, Honda, Vaiphasa,

Souris, & Nagai, 2012), rice (Stroppiana, Boschetti, Brivio, & Bocchi, 2009; Yi, Huang,

Wang, Wang, & Liu, 2007; Zhu, Zhou, Yao, Tian, & Cao, 2007), and many others.

With airborne and satellite multispectral sensors as the providers of remotely

sensed data, various shortcomings emerged (Merlin et al., 2010; Mulla, 2013). The ability

to deliver information for N management at high spatial resolution as dictated by

precision farming practices was the biggest limitation. In addition, the high cost of

images from aircraft, the infrequency of satellite overpasses, the interference of the

weather conditions with the imagery, and delays between image capture and availability

of usable data were major concerns for farmers. To implement precision N management

successfully, efficient technologies to diagnose crop N status are needed to determine in-

season, site-specific crop N requirements (Li et al., 2010). Unmanned aerial systems

(UAS) have been proposed for precision agriculture applications (Berni, Zarco-Tejada,

Suárez, & Fereres, 2009; Kooistra et al., 2013; Zhang & Kovacs, 2012). UAS can provide

imagery with a higher spatial resolution, allow for more flexible acquisition times, and

can be flown more frequently compared to satellite imagery (Zhang & Kovacs, 2012).

The current research explores the high spatial resolution imagery collected by a

UAS system to estimate leaf N content. The imagery acquires reflectance information

from the VIS, NIR, and thermal infrared (TIR) spectra with a spatial resolution of 15–60

cm. The estimated leaf N content developed by the proposed model was validated against

the ground truthing data collected. The generated high-resolution leaf N map could help

Page 66: The Application of Unmanned Aerial Vehicle to Precision

51

farmers in customizing their fertilization practices to suit the variability and needs of the

crop.

Our objective was to use canopy reflectance collected by AggieAir for estimating

the crop N status within fields. Moreover, this research also would identify which inputs

(e.g., individual reflectance, vegetative indices) are most sensitive for detection of crop N

status differences. In conducting the work, we evaluated a relatively inexpensive UAS

application in precision agriculture.

Material and Methods

Study Area

The study area is on a privately owned agricultural land in Scipio, Utah (39°14'N

112°6'W; see Figure 3.1). The field is equipped with a modern center pivot irrigation

sprinkler system to supply water for oat crops. The study reported here was carried out in

the summer of 2013. Four flights were flown over the study site. The flights on May 16,

June 1, June 9, and June 17 reflected the four stages of growth: 10 days after planting,

early growth, midgrowth, and early flowering. The study area was restricted to the

northwest quarter of the center pivot so that samples could be collected within a close

time frame relative to the AggieAir flight.

Page 67: The Application of Unmanned Aerial Vehicle to Precision

52

Figure 3.1. The location of the study area in Utah (left), The quarter used in leaf Nitrogen estimation model (right).

Instrumentation: Remote Sensing Platform AggieAir

AggieAir is a UAS designed to carry a payload to acquire aerial imagery in the

VIS, NIR, and TIR spectra. The unmanned aerial vehicle (UAV) navigates over an area

of interest according to a preprogrammed flight plan. While operational, the payload

computer instructs the three cameras to acquire imagery and records the position and

orientation of the aircraft when each image is taken. The UAS aircraft is battery powered

with a wingspan of 2.5 m and weight of 6.35 kg. AggieAir can fly in the air for up to 60

minutes and within elevation of 200–1000m.

The VIS camera used in AggieAir is a Canon S-95, with a 10-megapixel CCD

sensor and an ISO range of 80 to 3200. The NIR camera is an identical Canon S-95,

modified by replacing the manufacturer’s optical filter with a Wratten 87 NIR filter that

allows NIR wavelengths of 750 nm. AggieAir also carries a small, low-power,

microbolometer thermal camera from Infrared Cameras, Inc. (2014). The VIS-NIR

imagery is of a spatial resolution of 0.15 m; the TIR imagery is of 0.6 m spatial

resolution. The relative spectral responses of the VIS-NIR cameras were not provided by

Study site

Page 68: The Application of Unmanned Aerial Vehicle to Precision

53

the manufacturers but were obtained using the algorithm provided by Jiang, Liu, Gu, and

Susstrunk (2013). The wavelength range peaks around 420, 500, 600, and 800 nm,

respectively, for blue, green, red, and NIR sensors. Please refer to Chapters 2 and 4 of

this dissertation for more details about the UAV and the sensors.

AggieAir Data Processing

Following image acquisition, the imagery is imported into EnsoMOSAIC

(MosaicMill, 2012). EnsoMOSAIC is a photogrammetry software for aerial image

processing. The software generates hundreds of tie-points between overlapping images

by using photogrammetric principles in conjunction with image GPS log file data and

exterior orientation information from the on-board cameras to refine the estimate of the

position and orientation of individual images. The resulting image is an orthorectified

mosaic in 8-bit digital format. This mosaic is later converted to a mosaic of reflectance

values by applying a modified reflectance mode method (Clemens, 2012; Zaman, Jensen,

Clemens, & McKee, 2014). This radiometric normalization is the ratio of the digital

number from the mosaic to the digital number from a spectralon white reflectance panel

with known reflectance coefficients, multiplied by the reflectance factor, which accounts

for the zenith angle of the sun at the time, date, and location of the photos. The end

product is a four-layered image (blue, green, red, NIR) that is geometrically rectified and

radiometricaly corrected. As for the TIR imagery processing, the collected imagery is

also processed in EnsoMOSAIC, and the resulting thermal mosaic is composed of

brightness temperature in degrees Celsius. These mosaics are later converted to

radiometric temperatures following the procedure of Jensen (2014).

Page 69: The Application of Unmanned Aerial Vehicle to Precision

54

Ground Data Acquisition

Intensive ground sampling of plant N was conducted simultaneously with the

AggieAir flights at precise GPS locations. The GPS data were collected using an rtkGPS

with < 1 mm precision in a 1 Hz bandwidth (Trimble R8, Global Navigation Satellite

System, Dayton, Ohio). Leaves from the fully expanded top-of-canopy area were

collected. The samples were placed in paper bags inside an iced cooler and taken for

laboratory analysis. In the laboratory, plant samples were analyzed for leaf N content

(percent). Each plant sample was dried in an oven at 70°C, dry weights measured and

converted into dry biomass (kg/m2), and analyzed for N content.

Relevance Vector Machine (RVM)

The RVM was first introduced in 2000 and comprehensively detailed by Tipping

(2001). The RVM is a sparse Bayesian learning algorithm that relies on probabilistic

Bayesian principles to produce accurate predictions with a good generalization

performance (Tipping, 2001). The original learning algorithm was improved by Tipping

and Faul (2003), providing faster learning.

The RVM is based on a linear model where a prediction, y, is given as a weighted

sum of x basis functions. The model “learns” the dependence between input and output

target with the purpose of making accurate predictions of the target vector for

previously unseen values of as shown in Equation 1:

)(xwy (1)

Where a vector of weight parameters is, TNxxfxxfx )],(),...,,(,1[)( 1 is a design

matrix of vectors of kernel basis functions , and is the error that for

y

x

w

1N f ε

Page 70: The Application of Unmanned Aerial Vehicle to Precision

55

algorithmic simplicity is assumed to be zero-mean Gaussian with variance .

Through training of the RVM, the optimal weights are determined. The few

nonzero weights correspond to the so-called relevance vectors that are the sparse core of

the RVM model (Tipping & Faul 2003). The optimal parameters are then used to obtain

the optimal weight matrix with optimal covariance and mean. Model complexity,

overfitting, and computational expenses are controlled by setting weights to zero to

induce sparsity (Bachour, Walker, Ticlavilca, McKee, & Maslova, 2014). The

mathematical formulation, likelihood maximization, and optimization procedure of the

RVM are discussed in detail in Chapter 2. For further reading please refer to Tipping

(2001); Tipping and Faul (2003); and Thayananthan, Navaratnam, Stenger, Torr, and

Cipolla (2008).

Generating the Model Inputs and Targets

Three of the four flights (10 days after planting May 16, early growth June 1, and

midgrowth June 9), were used to generate the model training input data. The dataset

contains coincident in situ leaf N content and remote sensing reflectance measurements.

Each GPSed location where leaves were collected and tested for N content was paired by

its corresponding reflectance value. This was achieved by using the ArcGIS 10.2 ArcMap

software spatial analyst tool (Extract Multi Values to Points). In addition to the

reflectance in the VIS-NIR spectra, additional vegetative indices were generated. The

vegetative indices selected for this study were reported to be sensitive in estimating N

(Gitelson, Viña, Ciganda, Rundquist, & Arkebauer, 2005; Shanahan et al., 2003). Table

3.1 shows the indices formulations. In preliminary runs, different potential inputs were

2

Page 71: The Application of Unmanned Aerial Vehicle to Precision

56

explored. However, details on these preliminary runs are not reported in this study

because of their low statistical performance. A statistical description of the dataset is

presented in Table 3.2.

Table 3.1

Vegetative Indices Formulation Used for Leaf Nitrogen Estimation

Index Formula Reference

Normalized difference vegetation index (NDVI)

(RNIR – Rred) / (RNIR +RRed) Rouse, Haas, Schell, Deering, & Harlan, 1974

Leaf area indexa ln [(NDVI – NDVI max) / (NDVI min –NDVI max)] / -0.54

Duchemin et al., 2006; Smith, Bourgeouis, Teillet, Freemantle, & Nadeau, 2008

Green NDVI (RNIR – Rgreen) / (RNIR + Rgreen) Buschmann & Nagel, 1993

Ratio vegetation index NIR/red Jordan, 1969

Green ratio vegetation index (GRVI)

NIR/green Sripada, Heiniger, White, & Meijer, 2006

Red/green ratio index (RGRI)

Red/green Yang, Willis, & Mueller, 2008

Simple ratios (SR) Blue/green Birth & McVey, 1968 Note. NIR = near-infrared. aLAI was calculated empirically and not validated by field measurements.

Page 72: The Application of Unmanned Aerial Vehicle to Precision

57

Table 3.2

Statistical Description of the Dataset Used for Leaf Nitrogen Estimation

Index Range Mean ± SD

Potential inputs

AggieAir inputs

Blue 0.11–0.36 0.15 ± 0.04

Green 0.20–0.49 0.26 ± 0.05

Red 0.15–0.51 0.22 ± 0.07

Near-infrared 0.51–0.61 0.57 ± 0.02

Thermal (oC) 23.11–36.16 29.88 ± 4.13

Indices inputs

Green NDVI 0.00–0.46 0.35 ± 0.10

Ratio vegetation index 1.03–3.78 2.70 ± 0.60

Green ratio vegetation index (GRVI) 1.17–2.74 2.16 ± 0.33

Red/green ratio index (RGRI) 0.72–1.13 0.81 ± 0.08

Normalized difference vegetation index (NDVI) 0.00–0.78 0.44 ± 0.12

Simple ratio (SR) 0.47–0.73 0.56 ± 0.03

Leaf area index (m2/m2) 0.00–4.23 2.41 ± 0.90

Target output

Leaf nitrogen (mg/100 mg DT ) 0.00–5.36 3.905 ± 0.91

Model Configuration and Performance

Three of the images collected over the study area were used to build the model

(May 16, June 1, June 9). The remaining imagery (June 17) was later used to test the

model as unseen data scenarios. The RVM code used in this study is based on the

MATLAB code provided via Michael E. Tipping’s website. A key feature of this

statistical learning algorithm is that it can yield a solution function that depends on only a

Page 73: The Application of Unmanned Aerial Vehicle to Precision

58

very small number of training samples. These are called relevance vectors and are those

samples from the training set that have nonzero weights. In the RVM regression model,

the weight of each input is governed by a set of hyperparameters that describe the

posterior distribution of these weights. They are estimated iteratively during the machine

learning training step. The model “learns” the dependence between input and output

target with the purpose of making accurate predictions of the target vector, which was the

leaf N content in this study.

The model was developed with forward input selection process, in an attempt to

only include the input variables of most significance (Guyon & Elisseeff, 2003). In each

iteration, inputs with the best statistical performance were added to the set of inputs until

no new predictors could be added. The RVM model was tested using six kernel types:

Gauss, Laplace, spline, Cauchy, thin plate spline, and bubble. The performance of the

model was evaluated by comparing the root mean squared error (RMSE) and the Nash-

Sutcliffe efficiency (E); these two parameters have been widely used to evaluate the

performance of RVM models. The larger the value of E and the smaller the value of

RMSE, the greater the precision and accuracy of the model to predict leaf N. The optimal

values of these parameters were selected by trial and error procedure to obtain the best

RMSE and E values. The RMSE and E are computed as shown in Equations 2 and 3,

respectively:

N

yyRMSE

N

ttt

1

2)ˆ( (2)

Page 74: The Application of Unmanned Aerial Vehicle to Precision

59

N

t t

N

t t

yy

yyE

12

12

)(

)ˆ(1 (3)

where, ŷt = predicted nitrogen content ; yt = measured nitrogen content; = mean of the

observed nitrogen content; = mean of the estimated nitrogen content; and N = total

number of observations.

Results and Discussion

Each of the six kernel types was tested over a wide range of kernel widths (10-5–

105), and RMSE and E were calculated for all of the resulting models to assess their

predictive capabilities. An embedded loop in the coding model was developed to

represent the forward selection tool. For each type of kernel and its corresponding width,

the RVM was first run using all of the 8 inputs, consequently generating all of the needed

statistical model performance estimates to assess the model. A set of defined iterations

then added, in order, the input with the second best set of statistics, thus including the

input the most relevant to the target function. After numerous computational runs, the

model with the highest prediction accuracy was selected. The optimal kernel width,

kernel type, selected inputs, and statistical performance are presented in Table 3.3.

Of all runs conducted across the six kernel types, the NIR band followed by the

GRVI band were the two inputs selected first and second. This finding suggested these

two bands in the sensor used possess the most relevant information for estimating leaf N.

Canopy reflectance in the NIR regions have reported success at determining crop N

y

y

Page 75: The Application of Unmanned Aerial Vehicle to Precision

60

content, due to the ability to detect changes associated with chlorophyll content and

decreasing cell layers (Guyot, 1991; Thomas & Oerther, 1977).

Table 3.3

Selected Model Characteristics

Characteristic Statistic

Kernel type Gaussian

Kernel width 0.28

Inputs Near-infrared (NIR), green ratio vegetation index (GRVI), red/green ratio index (RGRI), simple ratio (SR)

Root mean square error (RMSE) 0.48 mg/100 mg DT

Nash-Sutcliffe efficiency (E) 0.76

Relevance vectors 6

Reflectance in the NIR spectrum has been reported to be sensitive to N estimate

over potato (Fava et al., 2009). In addition, when N total biomass is measured in the

laboratory by a spectroscopy an NIR band is used for its detection.

As for the GRVI index, a study that was monitoring leaf nitrogen status in rice

with canopy spectral reflectance (Xue et al., 2004) reported that the GRVI index

(NIR/green ratio) was especially linearly related to total leaf N, independent of N level

and growth stage. The study recommended the index to be adopted for nondestructive

monitoring of N status in rice plants. As for the RGRI index, Gamon and Surfus (as cited

in Gitelson, Gritz, & Merzlyak, 2003) stated that this index is particularly influential in

the separation of nonvegetation pixels from vegetation pixels, allowing for easy

discrimination between soil (larger red/green ratio) and vegetation (smaller red/green

Page 76: The Application of Unmanned Aerial Vehicle to Precision

61

ratio). To the best knowledge of the authors, the SR index (blue/green) has not been

reported to be influential in any attempts of N estimation from canopy reflectance.

The leaf N estimates for the two images used to build the model was developed

based the unique characteristic (kernel type, width, and set of inputs) of the selected

model. Figure 3.2 illustrates the measured leaf N versus estimated values with a one

standard error confidence interval and the residual plot. It is evident from Figure 3.2 that

the measured and RVM predicted leaf nitrogen values are in agreement, where most of

the values lie within the 95% confidence interval. It is important to mention that zero leaf

nitrogen values representing bare soil pixels were successfully assigned a zero value by

the RVM predicted model.

The lower plot in Figure 3.2 presents the residual error over the value from the

three flights. It is noticeable that the residual errors of the first flight (May 16) are higher

than the other two flights (June 1st and 9th). This could be explained by the drastic

variation in the percentage of crop cover in the first flight compared to the second and

third. In flight one, vegetation was scares and bare soil was more prominent in the

imagery. The predictive RVM model performance is superior in full crop cover compared

to less homogenous covers.

The leaf N estimate maps for May 16, June 1, and June 9 generated from the

model, along with the corresponding true color image, are presented in Figure 3.3.

Page 77: The Application of Unmanned Aerial Vehicle to Precision

62

Figure 3.2. Measured versus RVM predicted leaf nitrogen in the two flights used to build the model (above) and residual plot of the selected mode (below).

As shown in Figure 3.3, the predicted leaf N maps show a visual good agreement

with the true-color maps. Two visual comments can be drawn from the imagery in Figure

3.3:

1. The wheel tracks and the access road that are located around the center pivot

had no vegetation cover, and the model successfully assigned a near-zero N

content to these features.

Page 78: The Application of Unmanned Aerial Vehicle to Precision

63

2. Two high vegetation horizontal and vertical lines protrude in the images.

Those were past ditch lines that had been used in flood irrigation activities

prior to the conversion of the field to a center pivot system

Figure 3.3. True-color maps (left), the estimated plant leaf nitrogen estimate map (right).

The greater water content in those areas caused the plants growing along those two lines

to be very vigorous. The model estimated higher leaf N content values given to the

Page 79: The Application of Unmanned Aerial Vehicle to Precision

64

plants in these areas.

To explore the behavior of the model with previously unseen data, the June 17

flight was used. Now that the model was established with a defined set of features

(inputs, kernel type, kernel width), June 17 data (early flowering stage) were entered in

the model to explore the model’s performance when subjected to totally unseen data. The

predicted leaf N map is shown in Figure 3.4.

Figure 3.4. June 17 image in false color (left), and estimated leaf nitrogen map (right).

Again, the predicted leaf N map for the fourth flight showed good association

with the 4:2:3 map. Areas of vigorous growth, bare soil, and low vegetation were similar

in the maps and represented similar growth patterns. This test reported an RSME of 0.68

mg/100 mg DT and E of 0.61 for this flight. This result showed that the model

successfully performed when given unseen data.

This quarter of the field is irrigated by a modern center pivot sprinkler irrigation

system with a capacity of 610 GPM. A local upstream reservoir feeds this pivot regularly.

Page 80: The Application of Unmanned Aerial Vehicle to Precision

65

The pivot is automated to spray water every 15 degrees of rotation, creating six distinct

manageable areas in this quarter as shown in Figure 3.5.

Figure 3.5. The six individually manageable areas as defined in this study.

The estimated leaf N is averaged over the six sectors for the four flown flights.

Figure 3.6 is showing the average change in each sector over the course of the cropping

season. The highest leaf N are seen in the last imagery (June 17th) coinciding with the

early flowering stage in oats. The lowest Leaf N per sector were prominent in the first

flight (May 16th) where sparse vegetation was present in sectors three to six.

The information presented in the bar graph in Figure 3.6 aids the farmer in

quantifying the leaf nitrogen content among the different sectors during the growing

season. This data would be helpful when compared with the fertilization application

schedule to assess its efficiency within the different sectors.

Page 81: The Application of Unmanned Aerial Vehicle to Precision

66

Figure 3.6. Average leaf nitrogen per sector during the crop growth.

Averaging estimates over manageable areas allows the decision maker to consider

applying different amounts of fertilizer per sector to better accommodate for the need and

specification of the sector. Formatting the output in a form that the farmer can use is

another beneficial step in this new field. Providing site-specific, practical

recommendations to farmers is the next step in increasing the adoption rate of precision

agriculture.

Conclusion

This paper presented the application of imagery from AggieAir, a remote sensing

platform, combined with machine learning algorithms (RVM) to estimate leaf N content

concentration as an important biophysical parameter to be used in precision agriculture.

The RVM modeling technique, coupled with forward input selection, was applied to a

data set composed of reflectance from high-resolution multispectral imagery (VIS-NIR),

0

1

2

3

4

5

6

1 2 3 4 5 6

Leaf

nit

roge

n m

g/1

00

mg

DT

Sectors

16-May 1-Jun 9-Jun 17-Jun

Page 82: The Application of Unmanned Aerial Vehicle to Precision

67

TIR imagery, and vegetative indices, in conjunction with in situ leaf N measurements.

Six kernel types were tested over a wide range of kernel widths. Model performance was

evaluated by comparing the RMSE and E of various models and later by visual

comparison. Leaf N estimation was best achieved with a model of kernel type: Gaussian

kernel width: 0.28; selected inputs: NIR, GRVI, RGRI, SR; RSME: 0.48 mg/100 mg DT;

E: 0.76; and six relevance vectors. Across all of the model scenarios the NIR band

followed by the GRVI band were the two inputs selected, agreeing with the literature.

These two bands possess the most relevant information for estimating leaf N. While the

model developed was trained and built to estimate leaf N in oats, a similar methodology

can be applied to train a predictive model of a different crop type. With the foreseeable

high availability and fast development occurring in the high resolution imagery

platforms, these estimation models could become more common and further embedded in

farmer management tools.

References

Bachour, R., Walker, W., Ticlavilca, A., McKee, M., & Maslova, I. (2014). Estimation of spatially distributed evapotranspiration using remote sensing and a relevance vector machine. Journal of Irrigation and Drainage Engineering, 140(8), 04014029. doi:10.1061/(ASCE)IR.1943-4774.0000754

Berni, J., Zarco-Tejada, P. J., Suárez, L., & Fereres, E. (2009). Thermal and narrowband multispectral remote sensing for vegetation monitoring from an unmanned aerial vehicle. IEEE Transactions on Geoscience and Remote Sensing, 47, 722-738.

Birth, G. S., & McVey, G. R. (1968). Measuring the color of growing turf with a reflectance spectrophotometer. Agronomy Journal, 60(6), 640-643.

Blackmer, T. M., & Schepers, J. S. (1995). Use of chlorophyll meter to monitor nitrogen status and schedule fertigation for corn. Journal of Production Agriculture, 8, 55-60.

Page 83: The Application of Unmanned Aerial Vehicle to Precision

68

Blackmer, T. M., Schepers, J. S., & Varvel, G. E. (1994). Light reflectance compared with other nitrogen stress measurements in corn leaves. Agronomy Journal, 86, 934-938.

Buschmann, C., & Nagel, E. (1993). In vivo spectroscopy and internal optics of leaves as basis for remote sensing of vegetation. International Journal of Remote Sensing, 14, 711-722.

Chapman, S. C., & Barreto, H. J. (1997). Using a chlorophyll meter to estimate specific leaf nitrogen of tropical maize during vegetative growth. Agronomy Journal, 89, 557-562.

Clemens, S. R. (2012). Procedures for correcting digital camera imagery acquired by the AggieAir remote sensing platform (Doctoral dissertation). Available from http:// digitalcommons.usu.edu/gradreports/186

Daughtry, C. S. T., Walthall, C. L., Kim, M. S., De Colstoun, E. B., & McMurtrey, J. E., III. (2000). Estimating corn leaf chlorophyll concentration from leaf and canopy reflectance. Remote Sensing of Environment, 74, 229-239.

Duchemin, B., Hadria, R., Erraki, S., Boulet, G., Maisongrande, P., Chehbouni, A., . . . Simonneaux, V. (2006). Monitoring wheat phenology and irrigation in Central Morocco: On the use of relationships between evapotranspiration, crops coefficients, leaf area index and remotely-sensed vegetation indices. Agricultural Water Management, 79(1), 1-27.

Errebhi, M., Rosen, C. J., Gupta, S. C., & Birong, D. E. (1998). Potato yield response and nitrate leaching as influenced by nitrogen management. Agronomy Journal, 90, 10-15.

Fava, F., Colombo, R., Bocchi, S., Meroni, M., Sitzia, M., Fois, N., & Zucca, C. (2009). Identification of hyperspectral vegetation indices for Mediterranean pasture characterization. International Journal of Applied Earth Observation and Geoinformation, 11, 233-243.

Filella, I., Serrano, L., & Peñuelas, J. (1995). Evaluating wheat nitrogen status with canopy reflectance índices and discriminat analysis. Crop Science, 35, 1400-1405.

Gianquinto, G., Sambo, P., & Pimpini, F. (2003). The use of SPAD-502 chlorophyll meter for dynamically optimizing the nitrogen supply in potato crop: First results. Acta Horticulturae, 627, 225-230.

Gitelson, A. A., Gritz, Y., & Merzlyak, M. N. (2003). Relationships between leaf chlorophyll content and spectral reflectance and algorithms for non-destructive chlorophyll assessment in higher plant leaves. Journal of Plant Physiology, 160, 271-282.

Page 84: The Application of Unmanned Aerial Vehicle to Precision

69

Gitelson, A. A., Viña, A., Ciganda, V., Rundquist, D. C., & Arkebauer, T. J. (2005). Remote estimation of canopy chlorophyll content in crops. Geophysical Research Letters, 32(8). doi:10.1029/2005GL022688

Guyon, I., & Elisseeff, A. (2003). An introduction to variable and feature selection. The Journal of Machine Learning Research, 3, 1157-1182.

Guyot, G. (1991). Optical properties of vegetation canopies. In M. D. Steven & J. A. Clark (Eds.), Application of remote sensing in agriculture (pp. 19-43). London, England: Butterworths.

Han, S., Hedrickson, L., & Ni, B. (2001). Comparison of satellite remote sensing and aerial photography for ability to detect in-season nitrogen stress in corn. St. Joseph, MI: American Society of Agricultural and Biological Engineers.

Infrared Cameras, Inc. (2014). Home page. Retrieved from http://www .infraredcamerasinc.com

Jensen, A. M. (2014). Innovative payloads for small unmanned aerial system-based personal remote sensing and applications. Doctoral dissertation, Utah State University, Logan.

Jiang, J., Liu, D., Gu, J., & Susstrunk, S. (2013). What is the space of spectral sensitivity functions for digital color cameras? In 2013 IEEE Workshop on Applications of Computer Vision (WACV) (pp. 168-179). doi:10.1109/WACV.2013.6475015

Jordan, C. F. (1969). Derivation of leaf area index from quality of light on the forest floor. Ecology, 50, 663-666.

Ju, X. T., Kou, C. L., Zhang, F. S., & Christie, P. (2006). Nitrogen balance and groundwater nitrate contamination: Comparison among three intensive cropping systems on the North China Plain. Environmental Pollution, 143, 117-125.

Kooistra, L., Suomalainen, J., Iqbal, S., Franke, J., Wenting, P., & Bartholomeus, H. (2013, September). Crop monitoring using a light-weight hyperspectral mapping system for unmanned aerial vehicles : First results for the 2013 season. Paper presented at the Workshop on UAV-based Remote Sensing Methods for Monitoring Vegetation, Cologne, Germany.

Lee, W., Searcy, S. W., & Kataoka, T. (1999, July). Assessing nitrogen stress in corn varieties of varying color. Paper presented at the ASAE Annual International Meeting, Toronto, Ontario, Canada.

Li, F., Miao, Y., Hennig, S. D., Gnyp, M. L., Chen, X., Jia, L., & Bareth, G. (2010). Evaluating hyperspectral vegetation indices for estimating nitrogen concentration of winter wheat at different growth stages. Precision Agriculture, 11, 335-357.

Page 85: The Application of Unmanned Aerial Vehicle to Precision

70

Markwell, J., Osterman, J. C., & Mitchell, J. L. (1995). Calibration of the Minolta SPAD-502 leaf chlorophyll meter. Photosynthesis Research, 46, 467-472.

Martin, G. C., Shenk, J. S., & Barton, F. E., II. (1989). Near infrared reflectance spectroscopy (NIRS): Analysis of forage quality (Agriculture Handbook No. 643). Washington, DC: U.S. Department of Agriculture.

Merlin, O., Duchemin, B., Hagolle, O., Jacob, F., Coudert, B., Chehbouni, G., . . . & Kerr, Y. (2010). Disaggregation of MODIS surface temperature over an agricultural area using a time series of Formosat-2 images. Remote Sensing of Environment, 114, 2500-2512. doi:10.1016/j.rse.2010.05.025

Miphokasap, P., Honda, K., Vaiphasa, C., Souris, M., & Nagai, M. (2012). Estimating canopy nitrogen concentration in sugarcane using field imaging spectroscopy. Remote Sensing, 4, 1651-1670.

Mokhele, T. A., & Ahmed, F. B. (2010). Estimation of leaf nitrogen and silicon using hyperspectral remote sensing. Journal of Applied Remote Sensing, 4(1), 043560-043560.

MosaicMill. (2009). EnsoMOSAIC image processing user’s guide. Version 7.3. Vantaa, Finland: Author.

Mulla, D. J. (2013). Twenty five years of remote sensing in precision agriculture: Key advances and remaining knowledge gaps. Biosystems Engineering, 114, 358-371. doi:10.1016/j.biosystemseng.2012.08.009

Peterson, T. A., Blackmer, T. M., Francis, D. D., & Schepers, J. S. (1993). Using a chlorophyll meter to improve N management (Nebguide G93–1171A). Lincoln: University of Nebraska.

Prsa, I., Stampar, F., Vodnik, D., & Veberic, R. (2007). Influence of nitrogen on leaf chlorophyll content and photosynthesis of “Golden Delicious” apple. Acta Agriculturae Scandinavica Section B—Soil and Plant Science, 57, 283-289.

Ramesh, K., Chandrasekaran, B., Balasubramanian, T. N., Bangarusamy, U., Sivasamy, R., & Sankaran, N. (2002). Chlorophyll dynamics in rice (Oryza sativa) before and after flowering based on SPAD (chlorophyll) meter monitoring and its relation with grain yield. Journal of Agronomy and Crop Science, 188, 102-105.

Roth, G. W., Fox, R. H., & Marshall, H. G. (1989). Plant tissue tests for predicting nitrogen fertilizer requirements of winter wheat. Agronomy Journal, 81, 502-507.

Rouse, J. W., Jr., Haas, R. H., Schell, J. A., Deering, D. W., & Harlan, J. C. (1974). Monitoring the vernal advancements and retrogradation (green wave effect) of natural vegetation. Greenbelt, MD: NASA.

Page 86: The Application of Unmanned Aerial Vehicle to Precision

71

Schepers, J. S., Francis, D. D., Vigil, M., & Below, F. E. (1992). Comparison of corn leaf nitrogen concentration and chlorophyll meter readings. Communications in Soil Science & Plant Analysis, 23, 2173-2187.

Shanahan, J. F., Holland, K. H., Schepers, J. S., Francis, D. D., Schlemmer, M. R., & Caldwell, R. (2003). Use of a crop canopy reflectance sensor to assess corn leaf chlorophyll content. In Digital imaging and spectral techniques: Applications to precision agriculture and crop physiology (pp. 135-150). Madison, WI: American Society of Agronomy.

Smith, A. M., Bourgeois, G., Teillet, P. M., Freemantle, J., & Nadeau, C. (2008). A comparison of NDVI and MTVI2 for estimating LAI using CHRIS imagery: A case study in wheat. Canadian Journal of Remote Sensing, 34, 539-548.

Sripada, R. P., Heiniger, R. W., White, J. G., & Meijer, A. D. (2006). Aerial color infrared photography for determining early in-season nitrogen requirements in corn. Agronomy Journal, 98, 968-977.

Stroppiana, D., Boschetti, M., Brivio, P. A., & Bocchi, S. (2009). Plant nitrogen concentration in paddy rice from field canopy hyperspectral radiometry. Field Crops Research, 111(1), 119-129.

Thayananthan, A., Navaratnam, R., Stenger, B., Torr, P. H., & Cipolla, R. (2008). Pose estimation and tracking using multivariate regression. Pattern Recognition Letters, 29, 1302-1310.

Thomas, J. R., & Oerther, G. F. (1972). Estimating nitrogen content of sweet pepper leaves by reflectance measurements. Agronomy Journal, 64(1), 11-13.

Thomas, J. R., & Oerther, G. F., Jr. (1977). Estimation of crop conditions and sugar cane yields using aerial photography. Proceedings of the American Society of Sugar Cane Technologies, 6, 93-99.

Tipping, M. E. (2001). Sparse Bayesian learning and the relevance vector machine. The Journal of Machine Learning Research, 1, 211-244.

Tipping, M. E., & Faul, A. C. (2003). Fast marginal likelihood maximisation for sparse Bayesian models. In C. M. Bishop & B. J. Frey (Eds.), Proceedings of the Ninth International Workshop on Artificial Intelligence and Statistics (Vol. 1, No. 3). Society for Artificial Intelligence and Statistics.

Tracy, P. W., Hefner, S. G., Wood, C. W., & Edmisten, K. L. (1992). Theory behind the use of instantaneous leaf chlorophyll measurement for determining mid-season cotton nitrogen recommendations. In 1992 Beltwide Cotton Conference Proceedings (439-443). Memphis, TN: National Cotton Council.

Page 87: The Application of Unmanned Aerial Vehicle to Precision

72

Turner, F. T., & Jund, M. F. (1991). Chlorophyll meter to predict nitrogen topdress requirement for semidwarf rice. Agronomy Journal, 83, 926-928.

Wood, C. W., Reeves, D. W., and Himelrick, D. G. (1993), Relationships between chlorophyll meter readings and leaf chlorophyll concentration, N status, and crop yield: A review. Proceedings of the Agronomy Society of New Zealand, 23, 1-9.

Xue, L., Cao, W., Luo, W., Dai, T., & Zhu, Y. (2004). Monitoring leaf nitrogen status in rice with canopy spectral reflectance. Agronomy Journal, 96(1), 135-142.

Yang, Z., Willis, P., & Mueller, R. (2008, November). Impact of band-ratio enhanced AWIFS image to crop classification accuracy. Paper presented at Pecora 17—The Future of Land Imagine: Going Operational, Denver, CO.

Yi, Q. X., Huang, J. F., Wang, F. M., Wang, X. Z., & Liu, Z. Y. (2007). Monitoring rice nitrogen status using hyperspectral reflectance and artificial neural network. Environmental Science & Technology, 41, 6770-6775.

Zaman, B., Jensen, A., Clemens, S. R., & McKee, M. (2014). Retrieval of spectral reflectance of high resolution multispectral imagery acquired with an autonomous unmanned aerial vehicle. Photogrammetric Engineering & Remote Sensing, 80, 1139-1150.

Zhang, C., & Kovacs, J. (2012). The application of small unmanned aerial systems for precision agriculture: A review. Precision Agriculture, 13, 693-712. doi:10.1007 /s11119-012-9274-5

Zhu, Y., Zhou, D., Yao, X., Tian, Y., & Cao, W. (2007). Quantitative relationships of leaf nitrogen status to canopy spectral reflectance in rice. Crop and Pasture Science, 58, 1077-1085.

Page 88: The Application of Unmanned Aerial Vehicle to Precision

73

CHAPTER 4

ASSESSMENT ON THE USE OF AGGIEAIR UNMANNED AERIAL SYSTEM

REMOTE SENSING CAPABILITIES TO ESTIMATE CROP

EVAPOTRANSPIRATION AT HIGH SPATIAL RESOLUTION

Abstract

Estimation of spatial distribution of evapotranspiration (ET) based on remotely

sensed imagery has become of critical importance for managing water in irrigated

agriculture at fine spatial scales. Currently, data acquired by conventional satellites

(Landsat, Advanced Space borne Thermal Emission and Reflection Radiometer

[ASTER], etc.) lack the needed spatial resolution to capture within-farm variability to

support crop water-consumption estimates. Newly available remote-sensing technologies,

called unmanned aerial systems (UAS), are proving to be able to deliver time-flexible,

cost-effective, tailored spatial information for decision-making purposes. In this study, a

UAS called AggieAirTM was used to acquire high-resolution imagery in the visible (VIS),

near-infrared (NIR, 0.15 m resolution), and thermal infrared (TIR) spectra (0.6 m

resolution) for estimation of crop water use or actual ET. AggieAir flew over two study

sites in Central Utah and the California Central Valley. The imageries were used as input

to an extensively used surface energy balance model designed for Landsat called

Mapping Evapotranspiration With Internalized Calibration (METRIC™), and results at

AggieAir and Landsat scales were compared. The discussion highlights the difference in

ET estimates and the implications of high-resolution ET map availability.

Page 89: The Application of Unmanned Aerial Vehicle to Precision

74

Keywords: Evapotranspiration, Remote Sensing, Unmanned Aerial Systems, High

Resolution Imagery, Precision Agriculture, METRIC

Introduction

Water scarcity is becoming a major concern in agriculture, and the fear that future

water availability might be threatened has been a driving force to improve existing crop

water demand assessments. Failing to meet crop water requirements negatively affects

vegetative growth, yield, and various product-quality attributes. Water requirements are

typically described by the term evapotranspiration (ET). ET is the largest outgoing water

flux from agricultural surfaces and the most difficult to measure directly.

ET is the combination of two processes that occur simultaneously: evaporation

and transpiration, whereby water is lost to the atmosphere from soil and vegetation

respectively. The ET rate is expressed in depth (e.g., mm, inches) per unit time (hour,

day). Estimated ET has been widely used in research related to crop water use (Penman,

1948; Thevs, Peng, Rozi, Zerbe, & Abdusalih, 2015), soil water availability prediction

(Oki & Kanae, 2006), flood and drought forecasting (Bouilloud et al., 2010; Sheffield &

Wood, 2008), and desertification (Zhou, Zhu, & Sun, 2002). Above all, the calculation of

the ET rate has become vital in the planning and management of irrigation practices. In

agriculture, key water-management decisions focus on knowing when to begin irrigation

for the growing season, how often to irrigate, and how much water to apply. In precision

agriculture particularly, these critical decisions are often complicated not only by micro

weather and climate variability, but also by intrafield variations in soil texture, terrain,

and soil fertility. Precision irrigation, a common practice of precision agriculture, has

Page 90: The Application of Unmanned Aerial Vehicle to Precision

75

developed in the last few years practices to assess intra- and interfield spatial variability

and to implement site-specific management systems (Arnó, Martínez Casanovas, Ribes

Dasi, & Rosell, 2009; Barnes et al., 2000; Bramley, 2010; Sadler, Evans, Stone, & Camp,

2005). These management practices are designed to supply the crop with the needed

amount of water at the smallest manageable scale to obtain optimum response. However,

ET estimates obtained by various current methodologies are incapable of supporting

precision irrigation decisions due to the limitations discussed below.

Traditionally, ET is measured at the either point or farm scale (e.g., lysimeter–

point level, and eddy covariance system–farm level). These measurements are

problematic because they are average or single values, time consuming, expensive, only

applicable to small homogenous surfaces, and incapable of accounting for the dynamic

nature of fluxes when dealing with large spatial scales (Kaheil, Rosero, Gill, McKee, &

Bastidas, 2008). Therefore, extrapolating ET rates from a single value to a large area

decreases the accuracy of the estimation (Mauser & Schädlich, 1998; Petropoulos,

Carlson, Wooster, & Islam, 2009). For operational applications, many water managers

adopted the methodology of measuring ET published in the Guidelines for Computing

Crop Water Requirements (Allen, Pereira, Raes, & Smith, 1998). This method consists of

estimating crop ET and land-surface fluxes for a crop canopy using a weather station

reference ET and a crop coefficient (Kc), where reference ET is retrieved using the

Penman-Monteith method. The Penman-Monteith equation used in this methodology

calculates the reference ET over grass, assuming optimum soil moisture conditions and a

constant value of surface canopy resistance (Allen et al., 1998). In reality, soil moisture

Page 91: The Application of Unmanned Aerial Vehicle to Precision

76

conditions or surface canopy resistance is neither optimal nor constant. Therefore, it is

fundamental to have a spatial distribution of the land-surface fluxes to achieve more

accurate ET estimates.

In 1973 Brown and Rosenberg (as cited in Nouri, Beecham, Kazemi, Hassanli, &

Anderson, 2013) first used TIR remotely sensed temperature for predicting

evapotranspiration; since then, quantitative estimation of ET based on remotely sensed

data has evolved rapidly. Imagery collected from different platforms over various

temporal and spatial scales in conjunction with meteorological data from ground stations

became the most efficient and economic technology in ET estimation (Nouri et al., 2013).

Remote sensing technology collects spatially distributed observations of the land-surface

fluxes as radiances and temperatures. Radiances and temperatures measured are

converted into land-surface characteristics such as albedo, leaf area index (LAI),

vegetation indices, surface emissivity and surface temperature. Better estimating these

inputs individually results in more reliable ET estimates. All remote sensing based ET

estimates make use of the information collected in the TIR, VIS bands (blue, green, and

red [BGR]), and NIR spectrum. The TIR band is the most important variable in ET

estimates because it plays a crucial role in sensible heat flux, ground heat flux, and the

balance of long-wave radiation.

Using remotely sensed satellite imagery (e.g., Landsat, moderate resolution

imaging spectroradiometer [MODIS], geostationary operational environmental satellite

[GOES], ASTER), algorithms have been developed to estimate ET with various inputs

and model-structure complexity. The literature is rich with review papers that categorize,

Page 92: The Application of Unmanned Aerial Vehicle to Precision

77

explain, compare, and discuss the advantages and shortcomings of the different remote

sensing based ET estimate methods (Allen, Pereira, Howell, & Jensen, 2011a; Calcagno,

Mendicino, Monacelli, Senatore, & Versace, 2007; Contreras, Jobbagy, Villagra,

Nosetto, & Puiddefabregas, 2011; Courault, Seguin, & Olioso, 2005; Kustas & Norman,

1996; Li et al., 2009; Liou & Kar, 2014; Van Der Tol & Parodi, 2012). Many of the

remotely sensed data techniques are based on the surface energy balance and have been

considered the most accurate methods for estimating ET over spatially varying areas. Of

the energy balance methods, the Surface Energy Balance Algorithms for Land

(Bastiaanssen, Menenti, Feddes, & Holtslag, 1998) and METRIC (Allen, Tasumi, &

Trezza, 2002, 2007) have been extensively used and tested for operational accuracy in the

western United States and other areas in the world (Allen, Tasumi, Morse, et al., 2007;

Allen, Tasumi, & Trezza, 2007; Bastiaanssen et al, 2005). METRIC, developed by Allen

et al. (2002), is an image processing model for mapping ET as a residual of the energy

balance. METRIC uses a self-calibration procedure that involves ground-based hourly

reference ET measurements and the selection of extreme (hot and cold) pixels within

agricultural surfaces (Gowda et al., 2007). Performance of the METRIC model has been

tested by Gowda et al. (2007); Santos, Lorite, Tasumi, Allen, and Ferreres (2008); and

Tasumi, Trezza, Allen, and Wright (2005). These studies showed that ET estimates

derived from METRIC were comparable with data derived from soil moisture budget and

lysimeter.

Despite the advantages of using remote sensing techniques to measure ET, several

disadvantages have been reported (Allen, Pereira, Howell, & Jensen, 2011a, 2011b;

Page 93: The Application of Unmanned Aerial Vehicle to Precision

78

Boegh et al., 2009; Chen, Chen, Ju, & Geng, 2005; Courault et al., 2005; Jiang et al.,

2009; McCabe & Wood 2006; Min & Lin, 2006; Mutiga, Su, & Woldai, 2010).

Challenges are common for researchers using various remote sensing platforms like

Landsat (Allen et al., 2002; Moran, Jackson, Raymond, Gay, & Slater, 1989), MODIS

(Zhang & Wegehenkel, 2006), and aircraft-based remote sensing (Chávez, Neale, Hipps,

Prueger, & Kustas, 2005; Gómez, Olioso, Sobrino, & Jacob, 2005; Jacob et al., 2002a;

Neale, Jayanthi, & Wright, 2003). These limitations include (a) dealing with

heterogeneous pixels, which leads to uncertainties of land-surface variables; (b) the

recurrence of satellite passage, which results in missed opportunities for addressing plant

needs in a timely fashion; (c) and the high costs associated with obtaining high-resolution

images, particularly airborne images (Allen et al., 2011b; Boegh et al., 2009; Chen et al.,

2005; Courault et al., 2005; Jiang et al., 2009; McCabe & Wood, 2006; Min & Lin, 2006;

Mutiga et al., 2010; Rana & Katerji, 2000; Stisen, Sandholt, Nørgaard, Fensholt, &

Jensen, 2008; Wu, Cheng, Lo, & Chen, 2010). These limitations are most significant

when the estimates are intended for use in precision agriculture.

A potential remote sensing platform that could overcome some of these

limitations is the unmanned aerial vehicle (UAV) system capable of acquiring high-

resolution imagery in the VIS, NIR and TIR bands. Some of the most attractive features

of UAVs include high spatial resolution and lower operating costs (Berni et al., 2009).

These characteristics make this platform particularly suitable for agricultural application,

where multitemporal flights are required for management applications and high

resolution is crucial for recognizing the inherent spatial variability associated with crops.

Page 94: The Application of Unmanned Aerial Vehicle to Precision

79

UAV use is currently increasing due to the advantages mentioned (Baluja et al., 2012;

Chiabrando, Lingua, & Piras, 2013) and because it has been successfully used to research

other agricultural topics (Berni, Zarco-Tejada, Suárez, González-Dugo, & Fereres, 2009;

Demarez, Duthoit, Baret, Weiss, & Dedieu, 2008; Johnson et al., 2003; Pinter et al.,

2003; Zarco-Tejada, Berni, Suárez, & Fereres, 2008). Other studies (Jacob et al., 2002a;

Kustas et al., 2003; Kustas, Jackson, Prueger, MacPherson, & Wolde, 2004; McCabe &

Wood, 2006) evaluated the scale influences on the estimation using remotely sensed data

using conventional satellite platforms (e.g., Landsat 7, MODIS, ASTER, airborn),

Nevertheless, remotely sensed data acquired by UAVs has yet to be explored in ET

estimation applications.

The main objective of this study was to investigate the compatibility of using

high-resolution data (15 cm) acquired by a UAV system called AggieAir in conjunction

with a modified version of METRIC. The developed ET estimate was compared to the

Landsat 8 derived ET. The paper describes the UAV platform as well as the processing

chain from designing the flight plan to orthorectification and lastly radiometric

calibration.

Data Collection

Site Description

The study was conducted in the summer of 2013 over Site S and the summer of

2014 over Site G (see Figure 4.1). Site S is 80 hectares of privately owned agricultural

land in Scipio, Utah (39°14'N 112°6'W at 1628 m). The plot, mainly composed of loamy

clay soil, is equipped with a center pivot sprinkler for irrigating and planted with oats

Page 95: The Application of Unmanned Aerial Vehicle to Precision

80

(Avena sativa) and alfalfa (Medicago Sativa). Site G is a 70-hectare commercial vineyard

located in Lodi, California (43°38'N, 116W at 40 m) in a wine-producing region, under

Mediterranean–continental climate. Site G is planted with pinot noir (Vitis vinfiera)

vines, grafted on 110R rootstock, and trained to the Geneva Double Curtain trellis. Vine

spacing was 2 m between rows and 1.5 m in the row. Irrigation was applied during the

growing season through drip irrigation. Standard management practices were applied in

this vineyard. Global radiation, wind speed, air temperature, and humidity were acquired

by a meteorological station located on each site.

Figure 4.1. The location of the study areas in Scipio, Utah, and Lodi, California.

UAV Description

AggieAir is a UAS designed to carry camera payloads to acquire aerial imagery.

The UAS aircraft is battery powered and equipped with a payload system (which includes

three cameras and a computer), avionics, two inertial sensors (a GPS module and an

inertial measurement unit), radio controller, and flight control (see Figure 4.2). The

aircraft is propelled using an electric, brushless motor. It does not require a runway and

Site G Site S

Page 96: The Application of Unmanned Aerial Vehicle to Precision

81

can be flown autonomously or manually. In autonomous mode, the aircraft follows a

preprogrammed flight plan containing navigation waypoints defined by GPS and altitude.

While operational, the payload computer instructs the three cameras to acquire imagery

in the VIS, NIR, and TIR spectra and records the position and orientation of the aircraft

when each image is taken. See Table 4.1 for specifications.

Figure 4.2. Details of AggieAir airframe.

Table 4.1

AggieAir Unmanned Aerial System Features Description

Specification Range

Flight duration 45–60 min

Flight altitudes 200–1000 m

Maximum takeoff weight 6.35 kg

Visible/near-infrared spectrum resolution 6–25 cm

Thermal resolution 30–150 cm

Wing span 2.5 m

Page 97: The Application of Unmanned Aerial Vehicle to Precision

82

Sensors Description

The VIS camera used in AggieAir is a Canon S-95, with a 10-megapixel CCD

sensor and an ISO range of 80 to 3200. The NIR camera is an identical Canon S-95,

modified by replacing the manufacturer’s optical filter with a Wratten 87 NIR filter that

allows NIR wavelengths of 750 nm. AggieAir also carries a small, low-power,

microbolometer thermal camera from Infrared Cameras, Inc. (2014). The corresponding

relative spectral responses is shown in Figure 4.3.

Figure 4.3. Relative quantum response of the VIS-NIR (left) and thermal camera (right). VIS-NIR = visible/near-infrared; ICI = Infrared Cameras, Inc.

Flight Details

On June 9, 2013 and August 9, 2014 Site S and G, respectively, were flown by

AggieAir. Flights occurred following the Landsat image acquisition protocol, close to

0

0.2

0.4

0.6

0.8

1

380 480 580 680 780 880

Wavebands (nm)

VIS-NIR Relative Spectral Response

0

0.2

0.4

0.6

0.8

1

1200 6200 11200 16200 21200

Wavebands (nm)

ICI Thermal Camera Relative Spectral Response

Page 98: The Application of Unmanned Aerial Vehicle to Precision

83

solar noon (between 11 a.m. and 1 p.m.). The flight time (beginning to end) ranged from

30–40 minutes. Sensors used were Canon S95 cameras and an ICI 9000 TIR camera.

Images were collected in the VIS-NIR spectrum at a spatial resolution of 0.15 cm and at

0.60 cm for the TIR. The UAV was flown at an altitude of 450 m acquiring a total of 165

images from Site S and 240 images from Site G.

Processing Chain of AggieAir

The process chain in AggieAir mapping consists of the following steps: mission

planning, image processing and orthorectification, reflectance calibration, and thermal

data calibration. Each is described in detail.

Mission Planning

The two flight missions were planned with AggieAir Mission Complete, a tool

built by the AggieAir team that uses the open source World Wind SDK from NASA

software. The flight routes were designed over the two sites by accounting for three

major parameters: flight area (site dimension), camera specifications (focal length), and

the desired overlap in the imagery. A longitudinal overlap of 70% and a lateral overlap of

60% have been adopted while flying at 450 m altitude. Once all this information and the

flight altitude were introduced in the module, it automatically generated the flight route

and estimated the flight duration according to the total number of images planned. The

route file is exported to a memory card embedded in the UAV via a standard serial link.

Page 99: The Application of Unmanned Aerial Vehicle to Precision

84

Image Processing and Orthorectification

Following image acquisition, all images are imported into AggieAir Mission

Complete for further cleaning of the flight lines. In this process all the images collected

outside the area of interest, captured before reaching the desired altitude, or captured in

an oblique manner are deleted. The remaining images are stretched and made ready to be

combined into a single image. The process of stitching the imagery and geometrically

correcting it to align with geospatial coordinates is called orthorectification. A

photogrammetric software named Agisoft PhotoScan was employed for this task. In

Agisoft, the software generates hundreds of tie-points between overlapping images by

using photogrammetric principles in conjunction with image GPS log file data to

automatically align and orthorectify the imagery. To further refine the estimate of the

image position and orientation of individual-image ground control points (GCPs), control

points located within the area of interest were surveyed with an RTK GPS and added to

the image for further accuracy in the geo-referencing procedure. The outputs from

Agisoft are a four-band (BGR, NIR) orthorectified mosaic with digital numbers, a TIR

orthorectified mosaic with brightness temperature, a digital terrain model derived by

creating a mesh from a simplified and filtered version of the point cloud, digital surface

model, a high-density point cloud that makes details easily recognizable, and a detailed

summary report. An example of the summary report is shown in Figure 4.4, which

demonstrates camera positioning and overlapping in addition to other details related to

the software processing algorithm.

Page 100: The Application of Unmanned Aerial Vehicle to Precision

85

Figure 4.4. An example of the summary report, the number of overlapped images range from 1 to 9 per frame. The black dots represent the center of the frame collected over the four flight lines.

Radiometric Calibration

Radiometric calibration is a process of converting digital numbers into a measure

of reflectance. The radiometric quality of the images is critical in order to enable the

application of quantitative remote sensing methodologies for a successful estimation of

biophysical parameters from remote sensing imagery (Berni et al., 2009). AggieAir

follows the reflectance mode method conversion, which is based on methods adapted

from the research (Crowther, 1992; Miura & Huete, 2009; Neale & Crowther, 1994) and

discussed in detail by Clemens (2012). In this method the ratio of digital number mosaic

Page 101: The Application of Unmanned Aerial Vehicle to Precision

86

to the digital number from a Spectralon white reflectance panel is multiplied by the panel

reflectance factor, as shown in Equation 1. After learning that the new Labsphere is a

perfect Lambertian surface that does not need to account for sun angles or time of day,

weighted averages over the range of each band in the sensor were calculated using

Labsphere’s reflectance factor report. An updated reflectance factor was calculated, and

the product of this method was an orthorectified mosaic in reflectance values.

(1)

where: DNT(t): digital numbers in mosaic at the time t; DNR(t): digital number from

panel at the time t; RR is the reflectance factor of the reference panel with respect to a

Lambertian surface of unit reflectance

Thermal Calibration

TIR data are measured as brightness temperature by the sensor embedded on the

AggieAir payload. After data collecting the imagery, these measurements are corrected

and calibrated to account for various errors. Errors are introduced in the measurement

from the convection across the lens during the course of the flight. To develop the

correction equation, GCPs were collected simultaneously with the flights over the two

sites. The field data collection procedure was designed to acquire three GCP temperature

measurements over various sampling surfaces (e.g., high vegetative crops, inter-row

crops, bare soil, top of the vines canopy, etc.). For every GCP frame (a) TIR imagery is

collected by AggieAir at an elevation of 450 m on average, (b) TIR imagery is collected

by an identical sensor mounted on a moving vehicle from an elevation of 3 m, and (c) an

RR

TT R

tDNtDNR)()(

Page 102: The Application of Unmanned Aerial Vehicle to Precision

87

ambient absolute temperature is measured by a Blackbody instrument placed on the GCP.

Figure 4.5 is an example of two measurements collected.

Figure 4.5. An example of a ground control point (GCP), high vegetation, needed for calibration of thermal infrared (TIR) data. A picture of the GCP is captured right, a TIR imagery from a 3 m elevation in the middle, and the Blackbody reading on the left.

Since the Blackbody emissivity is 0.95 and not 1.0 like the TIR sensors,

adjustment to the Blackbody reading was necessary. Equation 2 describes the correction

done on the Blackbody readings.

Tbb_corr = Tbb*(1/BB emissivity) (1/4) (2)

Where: Tbb_corr = corrected temperature of the blackbody; Tbb = reading of the blackbody;

BB emissivity = Blackbody emissivity. The collected data were used to develop a

correction equation used to produce radiometric temperature.

Models: METRIC and METRIC High Resolution (METRIC-HR)

METRIC was selected as the base model to investigate the capability of AggieAir

data in estimating ET because it has been tested with operational applications in Idaho,

California, and Colorado and because of the researcher’s knowledge and expertise in this

model. A brief description of the original METRIC algorithm and the modified version

Page 103: The Application of Unmanned Aerial Vehicle to Precision

88

called METRIC-HR is provided below. A more detailed description of METRIC can be

found in the applications manual (Allen, Tasumi, Trezza, & Kjaersgaard, 2010). The

version of METRIC used in this study is METRIC v 2.0.8 (developed March 2012)

METRIC

METRIC algorithm is a satellite image processing model whose approach is

based on the rationale that ET is a change of the state of water using available energy in

the environment for vaporization (Su, McCabe, Wood, Su, & Prueger, 2005). This energy

is partitioned into net incoming radiation, ground heat flux, sensible heat flux, and latent

heat flux. METRIC estimates spatially distributed values of actual ET as the residual of

the energy balance (Allen et al., 2007):

LE =Rn − G − H (3)

where

LE = heat flux density (W m-2);

Rn = incoming radiation flux density (W m-2) and is calculated by solving the

radiation balance as described by (Allen, Tasumi, & Trezza, 2002)

G = soil heat flux density (W m-2) and is estimated as a function of sensible

heat flux (H), net incoming radiation Rn, surface temperature, and the

normalized difference vegetation index (NDVI) and LAI.

H = sensible heat flux density (W m-2) and is estimated as a function of air

density (kg m-3); air specific heat (J kg-1 K-1); temperature difference between

two heights; the aerodynamic resistance to heat transport (s m-1); and

Page 104: The Application of Unmanned Aerial Vehicle to Precision

89

temperature gradient between two near-surface air temperatures (K), generally

approximated at 0.1 m and 2 m above the canopy (Bastiaanssen et al., 1998).

The sensible heat flux is considered to be the most difficult term in the energy

balance equation to calculate using remote sensing; thus, computing H might require

multiple iterations. A strong linear relationship exists between dT and the radiometric

surface temperature (Allen, Tasumi, & Trezza, 2007; Basstiaannssen et al., 1998, 2005;

Jacob et al., 2002a). The relationship can be expressed in Equation 4:

𝑑𝑇 = 𝑎 + 𝑏𝑇𝑠 (4)

where a and b are empirically derived parameters based on two extreme conditions

pixels, termed hot and cold pixels. These anchor pixels define the upper and lower

bounds of the sensible heat flux in the study area and thus the gradient. The cold pixels

should represent well-watered and fully vegetated areas of the image, where no water

stress is present such that H is assumed to be minimal and ET maximum or near

maximum. As for the hot pixel, it should be located in a dry and bare agricultural field

where the evaporative flux is almost zero, thus H dominating the turbulent fluxes. Once

surface temperature, Ts, and dT are calculated corresponding to hot and cold conditions,

the linear relationship as indicated in Equation 4 is defined. To effectively select

reasonable cold and hot anchor pixels, the user must be skillful and understand the

principles associated with the energy balance. For further details, refer to (Allen, Tasumi,

Trezza, & Kjaersgaard, 2010).

Page 105: The Application of Unmanned Aerial Vehicle to Precision

90

METRIC-HR

Since METRIC is designed to use Landsat 8 imagery as inputs, a few adjustments

were made to accommodate the high-resolution AggieAir input data. These modifications

are described below.

Digital elevation model (DEM) and DEM high resolution (DEM-HR). To

account for the various topography of a site, METRIC incorporated a DEM. This DEM is

used to adjust surface temperatures for lapse effects caused by elevation variation as well

as in estimating solar radiation on slopes. The 30 m resolution DEM used in the original

METRIC is downloadable from the U.S. Geological Survey website. In METRIC-HR a

DEM-HR of 15 cm replaced the original coarse DEM. The DEM-HR was generated with

Agisoft software extension tool. Note that the two versions of the models use DEM term

to refer to the digital terrain model.

National Land Cover Database (NLCD) and NLCD high resolution. Another

basic input file needed to run METRIC is land use maps. The NLCD maps hold land

cover information. Those maps are also used to support the estimation of aerodynamic

roughness and soil heat flux during METRIC processing. In METRIC-HR a high-

resolution NLCD of 15 cm replaced the original coarse NLCD. A supervised

classification was performed on the NDVI maps to identify vegetation, roads, urban

infrastructure, and other features present in the high-resolution imagery. The obtained

classes were used to develop the NLCD high-resolution maps.

Shortwave infrared (SWIR) bands. METRIC requires the usage of the SWIR

band acquired by Landsat 8. SWIR is used to calculate the normalized difference water

Page 106: The Application of Unmanned Aerial Vehicle to Precision

91

index primarily to identify water, for it tends to give values less than zero for water and

snow and is therefore a more reliable water indicator than NDVI. However, water and

snow do not exist in either study site, and therefore the absence of SWIR has no effect on

the processed model. To ensure the model ran smoothly a pseudo null band replaced the

SWIR in METRIC-HR. In addition, METRIC incorporated the SWIR bands in the albedo

estimations; this was neglected by METRIC-HR, for albedo was estimated solely over

the available four bands: BGR and NIR.

Thermal band resampling. The thermal band acquired by AggieAir is of 60 cm

spatial resolution. METRIC requires all bands used in the model to have identical

resolution; therefore, resampling the TIR data was necessary. The resampling was

performed using the nearest neighbor resampling method in ESRI’s ArcGIS 10.2 ArcMap

software. Nearest neighbor resampling is a technique of raster data resampling where

each cell in an output raster is computed using the values of the nearest pixel in the input

image. This particular approach was selected as it does not alter the original pixel’s value

(Duggin & Robinove, 1990; Lillesand & Kiefer, 2000).

High-resolution VIS data. Landsat 8 shortwave radiance imagery (BGR) is

replaced by high-resolution shortwave in METRIC-HR. However, AggieAir sensor

acquires the BGR reflectance’s with a consumer-grade sensor whose spectral relative

response is different from Landsat 8, as shown in Figure 4.6.

Comparing the AggieAir and Landsat reflectances in the BGR spectrum revealed

that AggieAir reflectances were relatively higher than Landsat’s with a consistent trend in

both imagery of Site S and Site G. Therefore, a correction to the AggieAir reflectance in

Page 107: The Application of Unmanned Aerial Vehicle to Precision

92

the BGR spectrum was developed as described below. This process included two steps:

(a) upscaling AggieAir BGR using Landsat 8 point spread function (PSF) and (b)

developing the individual correction equations.

Figure 4.6. Landsat 8 and AggieAir relative spectral response.

Upscaling AggieAir BGR using Landsat 8 PSF. Several studies noted that

upscaling using the PSF produces reliable upscale data (Goforth, 1998); hence, this

methodology was adopted. The PSF of an imaging system is a measure of the amount of

blurring that occurs due to all of the components that comprise the imaging system

(Wenny et al., 2015). In other words, PSF assumes that the spectral information in a pixel

does not originate solely from within its footprint; a substantial portion comes from

surrounding areas (Forster & Best, 1994; Huang, Townshend, Liang, Kalluri, & DeFries,

2002; Townshend & Tucker, 1981). PSF is computed as the square modulus of the field

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

350 450 550 650 750 850 950 1050 1150

Wavebands (nm)

Landsat 8 & AggieAir Relative Spectral Response

Blue_AggieAir

Green_AggieAir

Red_AggieAir

Landsat_ Blue

Landsat_Green

Landsat_Red

Landsat_NIR

NIR_AggieAir

Page 108: The Application of Unmanned Aerial Vehicle to Precision

93

amplitude on the focal plane; the amplitude function, in turn, is built through the

diffraction integral, derived from the wave description of electromagnetic radiation (Gai

& Cancelliere, 2007). Landsat 8 PSF was developed by the team at U.S. Geographical

Survey Earth Resources Observation Systems Data Center and NASA Goddard Space

Flight Center. Using the data they provided, Figure 4.7 represents the PSF of the three

bands (BGR).

Figure 4.7. Landsat 8 point spread function (PSF) along the blue, green, and red (BGR) bands (left); Landsat 2D PSF representation on the right.

A squared matrix of 210 x 210 m slides over the AggieAir imagery every 30 m,

multiplying each pixel by its corresponding weight according to the PSF. The upscaling

equation is described in Equation 5.

(5)

After upscaling the BGR bands to 30 m, a linear equation was identified that

depicts the linear relationship between the 30 m upscale AggieAir and Landsat 8. Table

-0.005

0

0.005

0.01

0.015

0.02

0.025

-100 -50 0 50 100

Res

pons

e

Distance (m)

PSF along BGR bands

RedbandBlueband

𝜌30𝑚 = ∑ 𝑃𝑆𝐹 ∗ 𝜌0.15𝑚

210𝑚

Landsat pixel

Page 109: The Application of Unmanned Aerial Vehicle to Precision

94

4.2 presents individual equations used to correct the 30 m upscale AggieAir. Note that the

correction was performed on the VIS (BGR) bands and not on the NIR band. Given the

nature of the long pass filter on the NIR band, no correction was performed or required.

Table 4.2

The Developed Correction Equations for Each Band

Albedo

Surface albedo is defined as the fraction of incident solar energy (diffuse and

direct components) reflected both in all directions above the surface and over the whole

solar spectrum (Jacob & Olioso, 2002; Jacob et al., 2002b; Pinty & Verstraete, 1992).

The in situ albedo is calculated as the ratio of reflected to incoming solar radiation

measurements; however, such data were not collected from either site to calculate albedo.

Estimating albedo from remote sensing imagery has its challenges, bidirectional

reflectance distribution function being one of them. This phenomenon may result in over-

or underestimating of albedo. An overestimate of albedo by 20% occurs when the sun

and sensor angles match, thus resulting in higher canopy reflectance (Liu et al., 2009).

Conversely, when solar angle is substantially different from the sensor view angle, albedo

can be less than hemispherical albedo (Allen et al., 2011b). The bidirectional reflectance

Band Site G Site S

Blue y = 0.62x - 0.06 (r2 = 0.95) y = 0.58x - 0.04 (r2 = 0.99)

Green y = 0.51x - 0.01 (r2 = 0.89) y = 0.47x + 0.01 (r2 = 0.92)

Red y = 0.39x + 0.02 (r2 = 0.90) y = 0.42x + 0.02 (r2 = 0.94)

Page 110: The Application of Unmanned Aerial Vehicle to Precision

95

distribution function is corrected for MODIS-based albedo retrievals but not for Landsat

(Salomon, Schaaf, Strahler, Gao, & Jin, 2006). Acknowledging this phenomenon,

METRIC determines the albedo as a weighted average of the shortwave bands based on

the percentage of total at-surface radiation occurring with each band (Tasumi, Allen, &

Trezza, 2008). METRIC weighting coefficients proposed are for low-haze atmospheric

conditions and are optimized for Landsat images. Equation 6 represents the weighting

coefficients associated with the Landsat bands for calculating albedo (Tasumi et al.,

2008).

𝛼𝑠h𝑜𝑟𝑡=0.254𝛼1+0.149𝛼2+0.147𝛼3+0.311𝛼4+0.103𝛼5+0.036𝛼6 (6)

where αshort is the albedo and αn is the at-surface reflectance calculated for each Landsat

shortwave band.

Knowing the importance of the albedo values in selecting the extreme pixels

needed to properly run the model, customized albedo coefficients were developed. A

linear relationship between Landsat albedo and AggieAir reflectances was established as

shown in Equation 7.

NIRredgreenblueshort 6555.0538.0744.0716.0 (7)

Equation 7 was developed using the total number of 30 m pixels in the imagery

from both sites. These pixels included vines, vegetation, dry soil, and wet soil pixels. The

linear regression showed a residual error of root mean square error of 0.007 and an R2 of

0.88. Figure 4.8 shows a 1:1 plot of the albedos from Landsat and AggieAir over the two

sites. Note that the graph has two clusters visually; the lower left cluster of pixels form

Site S, and the upper right cluster are the pixels from Site G.

Page 111: The Application of Unmanned Aerial Vehicle to Precision

96

In theory the sum of coefficients should have a value close to 1, and coefficients

should be positive. However, when comparing the derived coefficients with Landsat

Figure 4.8. 1:1 Plot of the albedo from AggieAir and Landsat.

coefficients in the green band, a change of sign was noticed. Nevertheless, a

similar negative response in the green band was also reported in the study by Jacob and

Olioso (2002) that derived albedo coefficient from airborne platforms.

Results and Discussion

Hot and Cold Pixel Selection

Identifying the hot and cold pixels requires expertise and time to select a reliable

representation of the anchor pixels. Various sensitivity analyses reported that selecting

different hot and cold pixels leads to large deviations in final ET estimates (Wang et al.,

2009). The researchers properly select the hot and cold pixels that satisfy the assumptions

Page 112: The Application of Unmanned Aerial Vehicle to Precision

97

made in METRIC, such that the linear correlation between the near surface temperature

difference and remotely sensed surface temperature holds true.

METRIC recommends for the cold pixel to be located in a homogenous well-

watered, full-cover crop in the image with an NDVI range of 0.76–0.84 and a surface

albedo range of 0.18–0.24. As for the hot pixel, METRIC recommends that the hot

anchor pixel should be selected in homogenous bare and dry or nearly dry agricultural

soil with little or no vegetation, with an NDVI range no higher than 0.2 and a surface

albedo range of 0.17–0.23. Further details on the selection of the two pixels are described

by Allen, Tasumi, Morese, et al. (2007) and Allen (2008).

All four anchor points were selected from the perimeters of a common

overlapping area. The values of the NDVI, albedo, LAI, surface temperature, and the

coefficients for Equation 4 are reported in Table 4.3. Note that the sensitivity of the

selection of the hot/cold pixels in this study wasn’t performed.

The hot pixel in Site S, planted with alfalfa mainly, was selected from the clear,

bare soil area inside the northern center pivot. As for the cold pixel, more than one

candidate pixel were considered, for the crop was homogenous with full land cover in

various locations in the imagery. The cold pixel was selected from the northern part of the

southern center pivot. Both anchor pixels lay in the METRIC recommended ranges of

NDVI and albedos.

However, Site G, dominated with vineyard canopies, presented different

challenges in the selection process. Challenges were closely linked to the specific

geometrical canopy structure and the radiation balance pattern. The vineyard’s unique

Page 113: The Application of Unmanned Aerial Vehicle to Precision

98

characteristics are standing grape canopy, variable shaded areas, wide spacing between

rows, discontinuous soil cover, and various vertical leaf thickness.

Table 4.3

Details on the Selected Hot and Cold Pixels from the AggieAir and Landsat Imagery

Source Albedo NDVI LAI Ts a b

Site S AggieAir

Hot pixel 0.264 0.254 0.149 319.983 0.14270 -39.7412 Cold pixel 0.233 0.869 5.345 308.406

Landsat

Hot pixel 0.177 0.277 0.824 327.481 0.1176 -31.9722

Cold pixel 0.229 0.868 6.000 308.489

Site G AggieAir

Hot pixel 0.320 0.370 0.444 332.395 0.19919 -63.3236 Cold pixel 0.238 0.910 6.000 300.355

Landsat

Hot pixel 0.164 0.174 0.038 322.885 0.25852 -78.2922

Cold pixel 0.226 0.838 5.338 305.076 Note. LAI = leaf area index; NDVI = normalized difference vegetation index.

All these variables create complex dynamics that generate a spatial heterogeneity

in the horizontal distribution of energy at the soil surface; thus, finding a homogenous area

that fulfills the METRIC recommendation was a challenge to the end user. The albedo and

NDVI for the hot pixel exceeded the recommendations, possibly due to the lack of a

prominent bare soil area within the vineyard. The hot pixel was chosen from the northwest

side of the vineyard, which was distinguishable by the leafless vines, part of a separate

yield study.

Page 114: The Application of Unmanned Aerial Vehicle to Precision

99

METRIC Results

Landsat 8 cloud-free satellite images were obtained from the U.S. Geological

Survey Earth Explorer site (http://earthexplorer.usgs.gov) for June 9, 2013 and August 9,

2014. The images were processed using the ERDAS Imagine 2014 software. METRIC

was applied on the corresponding Landsat imagery of Site G and Site S to obtain

instantaneous and daily reference ET fraction (ETrf) maps. Maps of reflectance of

shortwave radiation, vegetation indices (NDVI and LAI), surface temperature, net

radiation, and soil heat flux were generated as intermediate products during METRIC

processing. The final outputs from the METRIC energy balance model were images

showing instantaneous ETrf (fraction of alfalfa-based reference ET) at the satellite

overpass time. Instantaneous ETrf produced after running METRIC over Site G and Site

S are shown in Figures 4.9 and 4.10, respectively.

Page 115: The Application of Unmanned Aerial Vehicle to Precision

100

Figure 4.9. Site S: Landsat reflectance in false color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (EtrF) output (right).

KcHigh : 1.15

Low : 0

00.1

0.20.05

Miles

Page 116: The Application of Unmanned Aerial Vehicle to Precision

101

Figure 4.10. Site G: Landsat reflectance bands in true color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (right).

METRIC-HR Results

METRIC-HR was run with the AggieAir derived inputs including corrected

reflectance (BGR, NIR), TIR band, DEM-HR, NLCD high resolution, albedo, and NDVI.

The resulting instantaneous ETrf images are shown in Figures 4.11 and 4.12.

KcHigh : 1.15

Low : 0

00.1

0.20.05

Miles

Page 117: The Application of Unmanned Aerial Vehicle to Precision

102

Figure 4.11. Site S: AggieAir reflectance in false color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (center), TIR (Celsius) map (right).

In figure 4.11 it is clear that the “crop coefficient” Kc, also known as “reference

ET fraction” ETrF, follows a similar spatial pattern as its corresponding false color and

TIR imagery. The values of the crop coefficient ranged between 0 and 1.15, with patterns

of ET linked to canopy temperature and cover. Lower values of Kc corresponded to

hotter areas where bare soil is mainly found, particularly in the outer surroundings of the

two center pivots and in the unplanted quarters in the northern center pivot. Higher values

of Kc are prominent in wet areas (one can see the pivot arm in the northern plot). Areas

of cooler temperatures where crops are homogenous and crop cover is dense have higher

Kc values.

00.1

0.20.05

Miles

Page 118: The Application of Unmanned Aerial Vehicle to Precision

103

Figure 4.12. Site G: AggieAir reflectance bands in true color (left), Mapping Evapotranspiration with Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) output (center) TIR (Celsius) map (right).

Figure 4.12 represents the Kc estimates from site G derived from the AggieAir

imagery. Considering that the vines planted in site G are all of the same type (Pinot noir),

similar growth stage and irrigated in a similar fashion (drip irrigation), there still exist a

variation in crop water demand within the field. This variation could be explained by the

different soil types present in the vineyard.

00.1

0.20.05

Miles

Page 119: The Application of Unmanned Aerial Vehicle to Precision

104

Comparison in Site S

Site S is irrigated by two modern, center-pivot sprinkler irrigation systems with a

capacity of 610 GPM each. A local upstream reservoir feed these two pivots regularly.

The pivots are automated to spray water every 15 degrees of rotation, creating 24 distinct

manageable areas in each center pivot as shown in Figure 4.13. The arcs are labeled from

1–48 covering the two center pivots.

Figure 4.13. Site S: Mapping Evapotranspiration With Internalized Calibration high resolution (METRIC-HR) reference evapotranspiration fraction (ETrf) on left, METRIC ETrf on right, both showing the 48 manageable arcs as defined in this study.

9

1

7

43

56

2

8

11

13

19

23

16

20

22 10

15

17

1224

18

14

21

363738 3539 343340

41 32

42 313043

44 2945 28

2746 47 262548

9

1

7

43

56

2

8

11

13

19

23

16

20

22 10

15

17

1224

18

14

21

363738 3539 343340

41 32

42 313043

44 2945 28

2746 47 262548

Page 120: The Application of Unmanned Aerial Vehicle to Precision

105

ETrf estimates were averaged over each sector. The obtained measurements from

the northern center pivot are shown in Figure 4.14.

Figure 4.14. Northern center pivot reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs.

. METRIC-HR resulted in a higher ETtrf estimated, except for Sectors 20 and 21.

This could be partly explained by the presence of pixels of multiple vegetation growth

with significant differences in cover (2 m and 15 cm alfalfa crops); a variation of surface

roughness is most prominent in these two sectors. However, identical results were

obtained from Sectors 1 and 13. These two sectors are the wettest sectors in the northern

center pivot, where the spraying arm could be seen in the high-resolution imagery.

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Etrf

Sector

Northern Center Pivot

Landsat AggieAir

Page 121: The Application of Unmanned Aerial Vehicle to Precision

106

Sectors 1 and 13 are also surrounded with vegetation outside the center pivot, resulting in

an extended homogenous land cover. The maximum difference in estimates occurs in

sector 7 (20% difference).

A similar analysis was performed on the southern center pivot. The averaged ETrf

over the sectors were plotted from the two tested models, as shown in Figure 4.15.

Figure 4.15. Southern center pivot reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs.

In the southern center pivot a more homogenous crop cover was observed. ETrf

data derived from AggieAir imagery estimated higher ETrf across all sectors of the center

0.00

0.20

0.40

0.60

0.80

1.00

1.20

25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Etrf

Sector

Southern Center Pivot

Landsat AggieAir

Page 122: The Application of Unmanned Aerial Vehicle to Precision

107

pivot. This could be explained by the surrounding dry landscape of the center pivot; these

peripheral mixed pixels lowered the average ET estimates in the sectors. The averaging in

each sector included the field edges where pixel contamination can cause METRIC ET

estimates to deviate from the field average (Allen, Tasumi, & Trezza, 2007).

A comparison between ETrf estimates derived from Landsat and AggieAir data is

presented in Figure 4.16. These values were close and had a 0.90 coefficient of

correlation.

Figure 4.16. Site S: Comparison between reference evapotranspiration fraction (ETrf) estimates derived from Landsat and AggieAir data in the northern and southern pivots

ETrf estimates obtained from the both the high resolution imagery and the

Landsat imagery, showed a high correlation (0.9) in the two center pivots. This implies

0.50

0.60

0.70

0.80

0.90

1.00

1.10

0.50 0.60 0.70 0.80 0.90 1.00 1.10

Land

sat

ETrf

AggieAir ETrF

Landsat vs AggieAir ETrf

Northern Center Pivot Southern Center Pivot Linear (Northern Center Pivot )

Page 123: The Application of Unmanned Aerial Vehicle to Precision

108

that both the METRIC- HR and METRIC models showed similar performance

capabilities. The high resolution imagery was superior in the mixed pixel areas (bare soil

and vegetation) specifically in the pixels in the outer contour of the center pivot. The

northern center pivot in this site study had a more heterogeneous surface compared to the

southern pivot. As a result, with respect to the northern pivot, a high resolution ET

estimates would be more beneficiary compared to the southern pivot.

Comparison in Site G

Site G, a well-maintained vineyard, is irrigated by drip system. The irrigation

design divides the vineyards of groups of 30–40 rows on average, each group creating a

block. There are 19 blocks in the vineyard. Each block can be regulated independently

with a ball valve upstream of the pressure regulator. The block is considered as the

smallest manageable area in this study. The distribution of the blocks is visualized in

Figure 4.17 as well as the ETrf estimates of AggieAir and Landsat.

To test the results of the spatially distributed ETrf estimates, all the pixels lying

within the blocks were averaged. The average represented a single water demand that

needs to be met by the water applicators.

The ETrf estimates obtained from both models are shown in Figure 4.18.

Visually, the Landsat ETrf estimate has a more fuzzy appearance. Compared to AggieAir

results, Landsat was underestimating ETrf in 12 of the 19 blocks. The comparison of the

ETrf estimates generated by both METRIC and METRIC-HR showed the largest

differences in Block 19. All the vegetative features in Block 19 are masked by the large

Landsat pixel footprint; visually it is not recognizable as a vegetative area.

Page 124: The Application of Unmanned Aerial Vehicle to Precision

109

Figure 4.17. Mapping Evapotranspiration With Internalized Calibration (METRIC) reference evapotranspiration fraction (ETrf) on left, METRIC high resolution ETrf in middle, and 19 blocks of individual manageable areas on right.

The high-resolution AggieAir imagery enabled the METRIC-HR to sense

evaporation occurring from the smallest block (Block 19), where METRIC could not.

19

1

2

3

4

5

6

7

8

9

10

11

12 13 14

15

16

17

18

Page 125: The Application of Unmanned Aerial Vehicle to Precision

110

The two estimates showed a correlation coefficient of .73. Figure 4.19 shows the two

estimates plotted against each other.

Figure 4.18. The average reference evapotranspiration fraction (ETrf) estimates from Landsat and AggieAir inputs for each block.

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

1.00

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

ETrf

Blocks

Sits G blocks

AggieAir

Landsat

Page 126: The Application of Unmanned Aerial Vehicle to Precision

111

Figure 4.19. Site G: Comparison between reference evapotranspiration fraction (ETrf) estimates derived from Landsat and AggieAir data.

Conclusion

The objective of this study was to assess the use of AggieAir, an unmanned aerial

system, to estimate crop evapotranspiration at high spatial resolution. A high resolution

METRIC (METRIC-HR) was derived from the well-established METRIC algorithm.

High resolution inputs (RGB, NIR, TIR, DEM, NLCD) and weather data were used to

develop an ET estimates at high resolution (0.15 m). A significant amount of spatial

information was retrieved and detailed ET estimates was established over two sites S and

G. This work demonstrated that it is possible to generate quantitative remote sensing

products by means of a UAV equipped with consumer-grade cameras. However, these

consumer-grade cameras require extensive calibration to relate spectral response to

another scientific sensor (such as Landsat). The high spatial resolution provided make

0.6

0.65

0.7

0.75

0.8

0.85

0.9

0.6 0.65 0.7 0.75 0.8 0.85 0.9

Land

sat E

Trf

AggieAir ETrf

Landsat vs AggieAir ETrf

Page 127: The Application of Unmanned Aerial Vehicle to Precision

112

this platform particularly suitable for precision agriculture and irrigation scheduling,

where site specific critical management is required. The High resolution ET showed is a

useful tool to monitor crop growth, and crop water demands when managing

heterogeneous surfaces. This model lays the ground for the estimation of ET at high

spatial resolution to be used in precision agriculture. The high resolution spatial

distribution of ET helped evaluating the efficiency of irrigation applications per the

smallest manageable area.

References

Allen, R. G. (2008). REF-ET: Reference evapotranspiration calculation software for FAO and ASCE standardized equations. Kimberly, ID: University of Idaho.

Allen, R. G., Pereira, L. S., Howell, T. A., & Jensen, M. E. (2011a). Evapotranspiration information reporting: I. Factors governing measurement accuracy. Agricultural Water Management, 98, 899-920. doi:10.1016/j.agwat.2010.12.015, 2011a.

Allen, R. G., Pereira, L. S., Howell, T. A., & Jensen, M. E. (2011). Evapotranspiration information reporting: II. Recommended documentation. Agricultural Water Management, 98, 921-929. doi:10.1016/j.agwat.2010.12.016, 2011b

Allen, R. G., Pereira, L. S., Raes, D., & Smith, M. (1998). Crop evapotranspiration—Guidelines for computing crop water requirements (FAO irrigation and drainage paper No. 56). Rome, Italy: Food and Agriculture Organization of the United Nations.

Allen, R. G., Tasumi, M., Morse, A., Trezza, R., Wright, J. L., Bastiaanssen, W., . . . & Robison, C. W. (2007). Satellite-based energy balance for Mapping Evapotranspiration With Internalized Calibration (METRIC)—Applications. Journal of Irrigation and Drainage Engineering, 133, 395-406. http://dx.doi.org /10.1061/(ASCE)0733-9437(2007)133:4(395)

Allen, R. G., Tasumi, M., & Trezza, R. (2002). METRICTM: mapping evapotranspiration at high resolution—Application manual for Landsat satellite imagery. Kimberly, ID: University of Idaho.

Page 128: The Application of Unmanned Aerial Vehicle to Precision

113

Allen, R. G., Tasumi, M., & Trezza, R. (2007). Satellite-based energy balance for Mapping Evapotranspiration With Internalized Calibration (METRIC)—Model. Journal of Irrigation and Drainage Engineering, 133, 380-394. doi:10.1061 /(ASCE)0733-9437(2007)133:4(380)

Allen, R. G., Tasumi, M., Trezza, R., Kjaersgaard, J. H. (2010). METRIC: Mapping evapotranspiration at high resolution. Applications manual, V. 2.0.4. Kimberly: University of Idaho.

Arnó, J., Martínez Casanovas, J., Ribes Dasi, M., & Rosell, J. R. (2009). Review. Precision viticulture. Research topics, challenges and opportunities in site-specific vineyard management. Spanish Journal of Agricultural Research, 7, 779-790.

Baluja, J., Diago, M. P., Balda, P., Zorer, R., Meggio, F., Morales, F., & Tardaguila, J. (2012). Assessment of vineyard water status variability by thermal and multispectral imagery using an unmanned aerial vehicle (UAV). Irrigation Science, 30, 511-522.

Barnes, E. M., Pinter, P. J., Jr., Kimball, B. A., Hunsaker, D. J., Wall, G. W., & LaMorte, R. L. (2000). Precision irrigation management using modeling and remote sensing approaches. In National Irrigation Symposium, Proceedings of the Fourth Decennial Symposium, Phoenix, AZ, ASAE (pp. 332-337).

Bastiaanssen, W. G. M., Menenti, M., Feddes, R. A., & Holtslag, A. A. M. (1998). A remote sensing Surface Energy Balance Algorithm for Land (SEBAL). 1. Formulation. Journal of Hydrology, 212-213, 198–212

Bastiaanssen, W. G. M., Noordman, E. J. M., Pelgrum, H., Davids, G., Thoreson, B. P., & Allen, R. G. (2005). SEBAL model with remotely sensed data to improve water-resources management under actual field conditions. Journal of Irrigation and Drainage Engineering, 131, 85-93.

Berni, J. A. J., Zarco-Tejada, P. J., Suárez, L., González-Dugo, V., & Fereres, E. (2009). Remote sensing of vegetation from UAV platforms using lightweight multispectral and thermal imaging sensors. International Archive of the Photogrammetry, Remote Sensing and Spatial Information Sciences, 38(6).

Boegh, E., Poulsen, R. N., Butts, M., Abrahamsen, P., Dellwik, E., Hansen, S., . . . & Søgaard, H. (2009). Remote sensing based evapotranspiration and runoff modeling of agricultural, forest and urban flux sites in Denmark: From field to macro-scale. Journal of Hydrology, 377, 300-316.

Bouilloud, L., Chancibault, K., Vincendon, B., Ducrocq, V., Habets, F., Saulnier, G. M., . . . & Noilhan, J. (2010). Coupling the ISBA land surface model and the TOPMODEL hydrological model for Mediterranean flash-flood forecasting:

Page 129: The Application of Unmanned Aerial Vehicle to Precision

114

description, calibration, and validation. Journal of Hydrometeorology, 11, 315-333.

Bramley, R. G. V. (2010). Precision viticulture: Managing vineyard variability for improved quality outcomes. In A. G. Reynolds (Ed.), Understanding and managing wine quality and safety (pp. 445-480). Cambridge, England: Woodhead.

Calcagno, G., Mendicino, G., Monacelli, G., Senatore, A., & Versace, P. (2007). Distributed estimation of actual evapotranspiration through remote sensing techniques. In G. Rossi, T. Vega, & B. Bonaccorso (Eds.), Methods and tools for drought analysis and management (pp. 124-147). Dordrecht, The Netherlands: Springer.

Chávez, J., Neale, C. M., Hipps, L. E., Prueger, J. H., & Kustas, W. P. (2005). Comparing aircraft-based remotely sensed energy balance fluxes with eddy covariance tower data using heat flux source area functions. Journal of Hydrometeorology, 6, 923-940.

Chen, J. M., Chen, X., Ju, W., & Geng, X. (2005). Distributed hydrological model for mapping evapotranspiration using remote sensing inputs. Journal of Hydrology, 305, 15-39.

Chiabrando, F., Lingua, A., & Piras, M. (2013). Direct photogrammetry using UAV: Tests and first results. International Archive of the Photogrammetry, Remote Sensing and Spatial Information Sciences, 41, 81-86. doi:10.5194/isprsarchives -XL-1-W2-81-2013

Clemens, S. R. (2012). Procedures for correcting digital camera imagery acquired by the AggieAir remote sensing platform (Doctoral dissertation). Retrieved from http:// digitalcommons.usu.edu/gradreports/186

Contreras, S., Jobbágy, E. G., Villagra, P. E., Nosetto, M. D., & Puigdefábregas, J. (2011). Remote sensing estimates of supplementary water consumption by arid ecosystems of central Argentina. Journal of Hydrology, 397, 10-22. doi:10.1016/j .jhydrol.2010.11.014, 2011

Courault, D., Seguin, B., & Olioso, A. (2005). Review on estimation of evapotranspiration from remote sensing data: From empirical to numerical modeling approaches. Irrigation and Drainage Systems, 19, 223-249.

Crowther, B. G. (1992). Radiometric calibration of multispectral video imagery. Doctoral dissertation, Utah State University, Logan.

Page 130: The Application of Unmanned Aerial Vehicle to Precision

115

Demarez, V., Duthoit, S., Baret, F., Weiss, M., & Dedieu, G. (2008). Estimation of leaf area and clumping indexes of crops with hemispherical photographs. Agricultural and Forest Meteorology, 148, 644-655.

Duggin, M. J., & Robinove, C. J. (1990). Assumptions implicit in remote sensing data acquisition and analysis. Remote Sensing, 11, 1669-1694.

Forster, B. C., & Best, P. (1994). Estimation of SPOT P-mode point spread function and derivation of a deconvolution filter. ISPRS Journal of Photogrammetry and Remote Sensing, 49, 32-42.

Gai, M., & Cancelliere, R. (2007). An efficient point spread function construction method. Monthly Notices of the Royal Astronomical Society, 377, 1337-1342.

Goforth, M. A. (1998). Fusion of differing resolution imagery using multiresolution analysis. In Proceedings of the Fusion International Conference (pp. 419-426).

Gómez, M., Olioso, A., Sobrino, J. A., & Jacob, F. (2005). Retrieval of evapotranspiration over the Alpille/ReSeDA experimental site using airborne POLDER sensor and a thermal camera. Remote Sensing of Environment, 96, 399-408.

Gowda, P. H., Chavez, J. L., Colaizzi, P. D., Evett, S. R., Howell, T. A., Tolk, J. A. (2007). ET mapping for agricultural water management: Present status and challenges. Irrigation Science, 26, 223-237.

Huang, C., Townshend, J. R., Liang, S., Kalluri, S. N., & DeFries, R. S. (2002). Impact of sensor’s point spread function on land cover characterization: Assessment and deconvolution. Remote Sensing of Environment, 80, 203-212.

Infrared Cameras, Inc. (2014). Home page. Retrieved from http://www .infraredcamerasinc.com

Jacob, F., & Olioso, A. (2002). Assessing recent algorithms devoted to albedo estimation using the airborne ReSeDa/PolDER database. In J. A. Sobrino (Ed.), Proceedings of the First International Symposium on Recent Advances in Quantitative Remote Sensing (pp. 191-198). València, Spain: Universitat de València.

Jacob, F., Olioso, A., Gu, X. F., Su, Z., & Seguin, B. (2002a). Mapping surface fluxes using airborne visible, near infrared, thermal infrared remote sensing data and a spatialized surface energy balance model. Agronomie, 22, 669-680.

Jacob, F., Olioso, A., Weiss, M., Baret, F., & Hautecoeur, O. (2002b). Mapping short-wave albedo of agricultural surfaces using airborne PolDER data. Remote Sensing of Environment, 80, 36-46.

Page 131: The Application of Unmanned Aerial Vehicle to Precision

116

Jiang, L., Islam, S., Guo, W., Jutla, A. S., Senarath, S. U., Ramsay, B. H., & Eltahir, E. (2009). A satellite-based daily actual evapotranspiration estimation algorithm over South Florida. Global and Planetary Change, 67(1), 62-77.

Johnson, L. F., Herwitz, S., Dunagan, S., Lobitz, B., Sullivan, D., & Slye, R. (2003). Collection of ultra-high spatial and spectral resolution image data over California vineyards with a small UAV. In Proceedings of the 30th International Symposium on Remote Sensing of Environment (Vol. 20, No. 845-849).

Kaheil, Y. H., Rosero, E., Gill, M. K., McKee, M., & Bastidas, L. A. (2008). Downscaling and forecasting of evapotranspiration using a synthetic model of wavelets and support vector machines. IEEE Transactions on Geoscience and Remote Sensing, 46, 2692-2707.

Kustas, W.P., French, A. N., Hatfield, J. L., Jackson, T. J., Moran, M. S., Rango, A., . . . & Schumgge, T. J. (2003). Remote sensing research in hydrometeorology. Photogrammetric Engineering and Remote Sensing, 69, 631-646.

Kustas, W. P., Li, F., Jackson, T. J., Prueger, J. H., MacPherson, J. I., & Wolde, M. (2004). Effects of remote sensing pixel resolution on modeled energy flux variability of croplands in Iowa. Remote Sensing of Environment, 92, 535-547.

Kustas, W. P., & Norman, J. M. (1996). Use of remote sensing for evapotranspiration monitoring over land surfaces. Hydrological Sciences Journal, 41, 495-516. doi:10.1080/02626669609491522

Li, Z.-L., Tang, R., Wan, Z., Bi, Y., Zhou, C., Tang, B., . . . & Zhang, X. (2009). A review of 20 current methodologies for regional evapotranspiration estimation from remotely sensed data. Sensors, 9, 3801-3853.

Lillesand, T. M., & Kiefer, R. W. (2000). Remote sensing and image interpretation. New York, NY: Wiley.

Liou, Y.-A., & Kar, S. K. (2014). Evapotranspiration estimation with remote sensing and various surface energy balance algorithms—A review. Energies, 7, 2821-2849. doi:10.3390/ en7052821

Liu, J., Schaaf, C., Strahler, A., Jiao, Z., Shuai, Y., Zhang, Q., … & Dutton, E. G. (2009). Validation of Moderate Resolution Imaging Spectroradiometer (MODIS) albedo retrieval algorithm: Dependence of albedo on solar zenith angle. Journal of Geophysical Research: Atmospheres (1984–2012), 114(D1).

Mauser, W., & Schädlich, S. (1998). Modelling the spatial distribution of evapotranspiration on different scales using remote sensing data. Journal of Hydrology, 212, 250-267.

Page 132: The Application of Unmanned Aerial Vehicle to Precision

117

McCabe, M. F., & Wood, E. F. (2006). Scale influences on the remote estimation of evapotranspiration using multiple satellite sensors. Remote Sensing of Environment, 105, 271-285.

Min, Q., & Lin, B. (2006). Remote sensing of evapotranspiration and carbon uptake at Harvard Forest. Remote Sensing of Environment, 100, 379-387.

Miura, T., & Huete, A. R. (2009). Performance of three reflectance calibration methods for airborne hyperspectral spectrometer data. Sensors, 9, 794-813.

Moran, M. S., Jackson, R. D., Raymond, L. H., Gay, L. W., & Slater, P. N. (1989). Mapping surface energy balance components by combining Landsat Thematic Mapper and ground-based meteorological data. Remote Sensing of Environment, 30, 77-87.

Mutiga, J. K., Su, Z., & Woldai, T. (2010). Using satellite remote sensing to assess evapotranspiration: Case study of the upper Ewaso Ng’iro North Basin, Kenya. International Journal of Applied Earth Observation and Geoinformation, 12, S100-S108. doi:10.1016/j.jag.2009.09.012

Neale, C. M., & Crowther, B. G. (1994). An airborne multispectral video/radiometer remote sensing system: Development and calibration. Remote Sensing of Environment, 49, 187-194.

Neale, C. M., Jayanthi, H., & Wright, J. L. (2003, September). Crop and irrigation water management using high resolution airborne remote sensing. Paper presented at the International Workshop on Use of Remote Sensing of Crop Evapotranspiration for Large Regions, 54th IEC Meeting of the International Commission on Irrigation and Drainage, Montpellier, France.

Nouri, H., Beecham, S., Kazemi, F., Hassanli, A. M., & Anderson, S. (2013). Remote sensing techniques for predicting evapotranspiration from mixed vegetated surfaces. Hydrology and Earth System Sciences Discussions, 10, 3897-3925.

Oki, T., & Kanae, S. (2006). Global hydrological cycles and world water resources. Science, 313(5790), 1068-1072.

Penman, H. L. (1948). Natural evaporation from open water, bare soil and grass. In Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences (Vol. 193, No. 1032, pp. 120-145). London, England: The Royal Society.

Petropoulos, G., Carlson, T. N., Wooster, M. J. & Islam, S. (2009). A review of Ts/VI remote sensing based methods for the retrieval of land surface energy fluxes and soil surface moisture. Progress in Physical Geography, 33, 224-250.

Page 133: The Application of Unmanned Aerial Vehicle to Precision

118

Pinter, J., Hatfield, J., Schepers, J. S., Barnes, E. M., Moran, M. S., Daughtry, C. S. T., & Upchurch, D. R. (2003). Remote sensing for crop management. Photogrammetric Engineering & Remote Sensing, 69, 647-664.

Pinty, B., & Verstraete, M. (1992). On the design and validation of surface bidirectional reflectance and albedo model. Remote Sensing of Environment, 41, 155-167.

Rana, G., & Katerji, N. (2000). Measurement and estimation of actual evapotranspiration in the field under Mediterranean climate: A review. European Journal of Agronomy, 13, 125-153. doi:10.1016/s1161-5 0301(00)00070-8

Sadler, E. J., Evans, R., Stone, K. C., & Camp, C. R. (2005). Opportunities for conservation with precision irrigation. Journal of Soil and Water Conservation, 60, 371-378.

Salomon, J. G., Schaaf, C. B., Strahler, A. H., Gao, F., & Jin, Y. (2006). Validation of the MODIS bidirectional reflectance distribution function and albedo retrievals using combined observations from the aqua and terra platforms. IEEE Transactions on Geoscience and Remote Sensing, 44, 1555-1565.

Santos, C., Lorite, I. J., Tasumi, M., Allen, R. G., & Fereres, E. (2008). Integrating satellite-based evapotranspiration with simulation models for irrigation management at the scheme level. Irrigation Science, 26, 277-288.

Sheffield, J., & Wood, E. F. (2008). Projected changes in drought occurrence under future global warming from multi-model, multi-scenario, IPCC AR4 simulations. Climate Dynamics, 31(1), 79-105.

Stisen, S., Sandholt, I., Nørgaard, A., Fensholt, R., & Jensen, K. H. (2008). Combining the triangle method with thermal inertia to estimate regional evapotranspiration—Applied to MSG-SEVIRI data in the Senegal River basin. Remote Sensing of Environment, 112, 1242-1255. doi:10.1016/j.rse.2007.08.013

Su, H., McCabe, M. F., Wood, E. F., Su, Z., & Prueger, J. H. (2005). Modeling evapotranspiration during SMACEX: Comparing two approaches for local- and regional-scale prediction. Journal of Hydrometerology, 6, 910-922.

Tasumi, M., Allen, R. G., & Trezza, R. (2008). At-surface reflectance and albedo from satellite for operational calculation of land surface energy balance. Journal of Hydrologic Engineering, 13, 51-63.

Tasumi, M., Trezza, R., Allen, R. G., & Wright, J. L. (2005). Operational aspects of satellite-based energy balance models for irrigated crops in the semi-arid U.S. Irrigation and Drainage Systems, 19, 355-376.

Page 134: The Application of Unmanned Aerial Vehicle to Precision

119

Thevs, N., Peng, H., Rozi, A., Zerbe, S., & Abdusalih, N. (2015). Water allocation and water consumption of irrigated agriculture and natural vegetation in the Aksu-Tarim river basin, Xinjiang, China. Journal of Arid Environments, 112, 87-97.

Townshend, J. R. G., & Tucker, C. J. (1981). Utility of AVHRR NOAA-6 and -7 for vegetation mapping. In Proceedings of the International Conference on Matching Remote Sensing Technologies and Their Applications (London/Nottingham: Remote Sensing Society (pp. 97-109).

Van der Tol, C., & Parodi, G. N. (2012). Guidelines for remote sensing of evapotranspiration. In A. Irmak (Ed.), Evapotranspiration—Remote sensing and modeling (Ch. 11). Rijeka, Croatia: InTech Open Access. doi:10.5772/18582

Wang, J., Sammis, T. W., Gutschick, V., Gebremichael, M., & Miller, D. R. (2009). Sensitivity analysis of the Surface Energy Balance Algorithm for Land (SEBAL). Transactions of the ASABE, 52, 801-811.

Wenny, B. N., Helder, D., Hong, J., Leigh, L., Thome, K. J., & Reuter, D. (2015). Pre-and post-launch spatial quality of the Landsat 8 thermal infrared sensor. Remote Sensing, 7, 1962-1980.

Wu, C., Cheng, C., Lo, H., & Chen, Y. (2010). Study on estimating the evapotranspiration cover 25 coefficients for stream flow simulation through remote sensing techniques. International Journal of Applied Earth Observation and Geoinformation, 12, 225-232. doi: 10.1016/j.jag.2010.03.001

Zarco-Tejada, P. J., Berni, J. A., Suárez, L., & Fereres, E. (2008, December 17). A new era in remote sensing of crops with unmanned robots. SPIE Newsroom. doi:10 .1117/2.1200812.1438

Zhang, Y., & Wegehenkel, M. (2006). Integration of MODIS data into a simple model for the spatial distributed simulation of soil water content and evapotranspiration. Remote Sensing of Environment, 104, 393-408.

Zhou, X. D., Zhu, Q. J., & Sun, Z. P. (2002). Preliminary study on regionalization desertification climate in China. Journal of Natural Disasters, 11, 125-131.

Page 135: The Application of Unmanned Aerial Vehicle to Precision

120

CHAPTER 5

SUMMARY

This dissertation has shown how data from consumer-grade sensors acquired by

an unmanned aerial system (UAS), AggieAir, could be used in precision agriculture

application. The high spatial, spectral, and temporal resolution makes this platform

particularly suitable for precision agriculture and irrigation scheduling, where time-

critical management is required. Chapter 2 presented the retrieval of chlorophyll content

from thermal and multispectral optical imagery. A complex statistical regression model

(Bayesian relevance vector machine) was trained on a dataset of in situ collected leaf

chlorophyll measurements, and the machine learning algorithm intelligently selected the

most appropriate bands and indices for building regressions with the highest prediction

accuracy. Chapter 3, with a similar methodology to Chapter 2, discussed the estimation of

leaf nitrogen content. The predicted estimates were averaged over the smallest

manageable unit in the center pivot, thus providing a preliminary actionable information

for nutrient management. Chapter 4 investigated the possibility of mapping

evapotranspiration using the high spatial resolution data. Appropriate configurations were

applied on the Mapping Evapotranspiration With Internalized Calibration algorithm to

match the input data. Evapotranspiration estimates from AggieAir inputs were compared

to those obtained from Landsat 8 data and showed agreement between the two. The

estimates were averaged over the smallest irrigateable unit in the two study sites to assist

in a more precise efficient irrigation scheduling. All these estimates, made at such fine

resolutions in space and time, can aid farmers in assessing the heterogeneity of their

Page 136: The Application of Unmanned Aerial Vehicle to Precision

121

fields and subsequently implement needed actions accordingly. The high-resolution

spatial information generated from AggieAir imagery could enable far greater precision

in the application of various production resources (fertilizers, irrigation water). UAS

facilitate enhanced monitoring with time-critical account of fine-scale variations in plant

health and function.

The final findings presented in this dissertation are in general promising and

encourage the development of future improvements. Plenty could be done to improve the

collection, processing, and quality of the data. Perhaps the most urgent research topics to

address are the automatic georeferencing of imagery, establishing a standardized

procedure for orthorectification, and better understanding of the sensors calibration and

performances. Also, data acquired by the UAS could be explored in different research

within precision agriculture (e.g., crop stress, yield potential, crop disease). In addition,

exploring UAS data with physical-based model approaches rather than with statistical

and engineering approaches will create more robust algorithms that can be generalized

over different kinds of crops. Perhaps closing the link between scientists who generate

information from UAS data and farmers by presenting these findings in an actionable

manner remains the biggest missing link for further adoption of precision agriculture and

a significant research topic that needs to be addressed in future studies.

This work has demonstrated that the technology needed to produce actionable

information is possible but unfortunately, the cost embedded in this technology is far

from being affordable. The cost of flying the payload, collecting enough representable

samples to train the models and processing the data exceeds the limits of the average

Page 137: The Application of Unmanned Aerial Vehicle to Precision

122

farmer and as a result restrain the technology to research purposes only. Another big

challenge this technology presents is the big amount of data it produces. The “big data”

itself has advantages and short comes. A prominent advantage of “big data” is the

knowledge that emerges from the data. Such findings can change, modify or better

inform us about the research in question. For example, exploring the TIR data and its

relationship with chlorophyll estimation is something that haven’t been widely explored

in the community and this finding opened the door to such interest. On the other hand,

these “big data” sets are large enough to require supercomputers. Storing, managing and

processing these data is beyond the ability of commonly used computers. Perhaps, the

most fundamental challenge for big data applications is to explore the large volumes of

data and extract useful information or knowledge for future actions.

Page 138: The Application of Unmanned Aerial Vehicle to Precision

123

APPENDIX

Page 139: The Application of Unmanned Aerial Vehicle to Precision

124

Permission Letters

Page 140: The Application of Unmanned Aerial Vehicle to Precision

125

From: Permissions Helpdesk <[email protected]> Date: Mon, Mar 21, 2016 at 1:33 PM Subject: RE: Permission To: Manal Alarab <[email protected]>

Dear Manal, As an Elsevier journal author, you retain the right to Include the article in a thesis or dissertation (provided that this is not to be published commercially) whether in part or in toto, subject to proper acknowledgment; see http://www.elsevier.com/about/company-information/policies/copyright/personal-use for more information. As this is a retained right, no written permission from Elsevier is necessary. As outlined in our permissions licenses, this extends to the posting to your university’s digital repository of the thesis provided that if you include the published journal article (PJA) version, it is embedded in your thesis only and not separately downloadable: 19. Thesis/Dissertation: If your license is for use in a thesis/dissertation your thesis may be submitted to your institution in either print or electronic form. Should your thesis be published commercially, please reapply for permission. These requirements include permission for the Library and Archives of Canada to supply single copies, on demand, of the complete thesis and include permission for Proquest/UMI to supply single copies, on demand, of the complete thesis. Should your thesis be published commercially, please reapply for permission. Theses and dissertations which contain embedded PJAs as part of the formal submission can be posted publicly by the awarding institution with DOI links back to the formal publications on ScienceDirect. Best of luck with your thesis and best regards, Laura Laura Stingelin Permissions Helpdesk Associate Elsevier 1600 John F. Kennedy Boulevard Suite 1800 Philadelphia, PA 19103-2899 T: (215) 239-3867 F: (215) 239-3805 E: [email protected] Questions about obtaining permission: whom to contact? What rights to request? When is permission required? Contact the Permissions Helpdesk at:

Page 141: The Application of Unmanned Aerial Vehicle to Precision

126

+1-800-523-4069 x 3808 [email protected] From: Manal Alarab [mailto:[email protected]]

Sent: Monday, March 21, 2016 2:18 AM To: Permissions Helpdesk

Subject: Permission Dear editors, I am requesting permissions to my paper in the “ International Journal of Applied Earth Observation and Geoinformation”, as a Chapter in my dissertation. DOI of my paper: doi:10.1016/j.jag.2015.03.017 Reference: Elarab, Manal, Andres M. Ticlavilca, Alfonso F. Torres-Rua, Inga Maslova, and Mac McKee. "Estimating chlorophyll with thermal and broadband multispectral high resolution imagery from an unmanned aerial system using relevance vector machines for precision agriculture." International Journal of Applied Earth Observation and Geoinformation 43 (2015): 32-42. Manal Elarab, PhD Civil and Environmental Engineering Utah Water Research Laboratory - USU Phone: +1 (435) 754-9299 From: Inga Maslova <[email protected]> Date: Mon, Mar 21, 2016 at 8:02 AM Subject: Re: Permission To: Manal Al Arab <[email protected]>

Dear Manal,

Yes, you have my permission to include the paper in your thesis.

Page 142: The Application of Unmanned Aerial Vehicle to Precision

127

Sincerely, Inga Maslova Assistant Professor Department of Mathematics and Statistics American University

On Mar 21, 2016 2:21 AM, "Manal Alarab" <[email protected]> wrote: Dear Inga, I am in the process of preparing my dissertation. And I'm sending this email to request your permission to include in my dissertation a copy of the following paper in which you appear as a coauthor: ”ESTIMATING CHLOROPHYLL FROM THERMAL AND BROADBAND MULTISPECTRAL HIGH RESOLUTION IMAGERY FROM AN UNMANNED AERIAL SYSTEM USING RELEVANCE VECTOR MACHINES FOR PRECISION AGRICULTURE" Best, Manal -- Manal Elarab, PhD Civil and Environmental Engineering Utah Water Research Laboratory - USU Phone: +1 (435) 754-9299 From: Alfonso Torres-Rua <[email protected]> Date: Mon, Mar 21, 2016 at 6:48 AM Subject: Re: Permission To: Manal Alarab <[email protected]> Dear Manal, You have my permission. Alfonso Sent from my iPhone On Mar 21, 2016, at 12:20 AM, Manal Alarab <[email protected]> wrote:

Dear Alfonso,

Page 143: The Application of Unmanned Aerial Vehicle to Precision

128

I am in the process of preparing my dissertation. And I'm sending this email to request your permission to include in my dissertation a copy of the following paper in which you appear as a coauthor: ”ESTIMATING CHLOROPHYLL FROM THERMAL AND BROADBAND MULTISPECTRAL HIGH RESOLUTION IMAGERY FROM AN UNMANNED AERIAL SYSTEM USING RELEVANCE VECTOR MACHINES FOR PRECISION AGRICULTURE" Best, Manal -- Manal Elarab, PhD Civil and Environmental Engineering Utah Water Research Laboratory - USU Phone: +1 (435) 754-9299 From: Andres Ticlavilca <[email protected]> Date: Mon, Mar 21, 2016 at 5:52 AM Subject: Re: Permission To: Manal Alarab <[email protected]>

Dear Manal You have my permission. Best regards, Andres On Mon, Mar 21, 2016 at 2:21 AM, Manal Alarab <[email protected]> wrote: Dear Andres, I am in the process of preparing my dissertation. And I'm sending this email to request your permission to include in my dissertation a copy of the following paper in which you appear as a coauthor: ”ESTIMATING CHLOROPHYLL FROM THERMAL AND BROADBAND MULTISPECTRAL HIGH RESOLUTION IMAGERY FROM AN UNMANNED

Page 144: The Application of Unmanned Aerial Vehicle to Precision

129

AERIAL SYSTEM USING RELEVANCE VECTOR MACHINES FOR PRECISION AGRICULTURE" Best, Manal -- Manal Elarab, PhD Civil and Environmental Engineering Utah Water Research Laboratory - USU Phone: +1 (435) 754-9299 -- Andres M. Ticlavilca, Ph.D. Research Engineer Utah Water Research Laboratory 8200 Old Main Hill Logan, UT 84322-8200 Phone: 435-757-0851 Email: [email protected] Web: https://sites.google.com/a/aggiemail.usu.edu/ticlavilca/ -- Manal Elarab, PhD Civil and Environmental Engineering Utah Water Research Laboratory - USU Phone: +1 (435) 754-9299

Page 145: The Application of Unmanned Aerial Vehicle to Precision

130

CURRICULUM VITAE

Manal Elarab

[email protected]| 435-754-9299 (US Citizen)

Professional Profile An Engineer scientist with extensive background in agricultural research and remote sensing geospatial analysis. Proven ability to generate innovative approaches in precision agriculture combining UAV data and machine learning algorithms. Practical experience working in a team performing research to apply in commercial applications. Excellent initiative, communication, problem solving and management skills. Areas of Expertise Precision Agriculture Computational Modeling Geospatial Data Analysis Statistical Analysis Machine Learning UAV Application Spatial Evapotranspiration Spatial Data Mining Experimental Design

Experience Graduate Research Assistant 2012-2015 Utah Water Research Laboratory, Logan, Utah

Actively involved in development and calibration of UAV imagery—sensor selection, reflectance calibration, thermal calibration, orthorectification, ground data collection, comparative analysis, upscaling/downscaling data techniques

Developed models to estimate leaf chlorophyll and plant tissue nitrogen using high resolution thermal and broadband multi-spectral UAV imagery combined with relevance vectors machine algorithm and validated by ground data

Adapted METRIC model to estimate high resolution crop evapotranspiration using UAV data and evaluated against Landsat METRIC estimates

Graduate Research Assistant 2009-2011 American University of Beirut, Lebanon

Designed and installed irrigation systems Conducted experiments on water saving in agriculture and deficit irrigation

Page 146: The Application of Unmanned Aerial Vehicle to Precision

131

Relevant Projects

Private Farm—Oats and Alfalfa, Scipio, Utah (2012 – 2013): Developed models to estimate plant chlorophyll and leaf nitrogen content using UAV high resolution data. Experiment was conducted over two center pivots.

Designed data collection procedure, location and frequency Installed ground sensors and weather stations Collected field data over two growing seasons to develop and train statistical

models Collected data to calibrate thermal and reflectance imagery Coordinated experimental activity with farmers—harvesting, irrigation

scheduling, and fertilization Gallo Commercial Vineyard, Lodi, California (2014-2015): Mapped Evapotranspiration over vineyards using UAV high resolution remote sensing data, funded by ARS-USDA.

Designed and conducted data collection to calibrate thermal and reflectance imagery

Corrected high resolution reflectance data (visual and infrared) to Landsat 8 imagery

Developed Albedo coefficients customized to the UAV Adjusted METRIC model to accommodate the high resolution UAV data to

estimate crop evapotranspiration

Education Doctor of Philosophy in Civil and Environmental Engineering Dec 2015 Utah State University, Logan, Utah

Dissertation: The Application of Unmanned Aerial Vehicle to Precision Agriculture: Chlorophyll, Nitrogen, and Evapotranspiration Estimation

Masters of Science in Irrigation Engineering Dec 2011 American University of Beirut, Lebanon Masters of Science in Plant Production May 2008 Lebanese University, Lebanon Bachelor of Science in Agricultural Engineering May 2007 Lebanese University, Lebanon Skills

Computational: WEAP, Arc Map, Matlab, R, Python, Erdas Imagine Language: Fluent Arabic and English, Intermediate French

Page 147: The Application of Unmanned Aerial Vehicle to Precision

132

Interests and Activities

Cooking, Reading, Travelling, Hiking, Languages, Bowling and Mountain Biking Peer Reviewed Publications

Elarab, M., Ticlavilca, A. M., Torres-Rua, A. F., Maslova, I., & McKee, M. (2015). Estimating chlorophyll with thermal and broadband multispectral high resolution imagery from an unmanned aerial system using relevance vector machines for precision agriculture. International Journal of Applied Earth Observation and Geoinformation, 43, 32-42. Torres-Rua, A., Al Arab, M., Hassan-Esfahani, L., Jensen, A., & McKee, M. (2015). Development of unmanned aerial systems for use in precision agriculture: The AggieAir experience. In Technologies for Sustainability (SusTech), 2015 IEEE Conference on (pp. 77-82). IEEE. M. Al-Arab, A. Torres-Rua, A. M. Ticlavilca, A M. Jensen, M. McKee (2013). Use of high-resolution multispectral imagery from and unmanned aerial vehicle in precision agriculture. IEEE International Geoscience and Remote Sensing Symposium, Melbourne, Australia. M. Al-Arab L. Meeks A. Gilchrist (2012). Delaware River basin, Wikipedia www.engr.usu.edu/wiki/index.php/Delaware_River_Basin R. Bachour, M. Al-Arab, M. Nimah., & M. McKee (2011). Research Priorities for Water and Irrigation System Management on the Orontes Benchmark Site, Lebanon. Conferences M. Elarab, A. Torres-Rua, M. McKee; “Use of UAS Remote Sensing Data (AggieAir) to Estimate Crop ET at High Spatial Resolution” American Geophysical Union – Fall Meeting 2015 – Oral Presentation. M. Elarab, McKee, Mac; Torres-Rua, Alfonso FHassan-Esfahani, Leila; Jensen, Austin; “Use of UAS to Support Management in Precision Agriculture: The AggieAir Experience” ASA, CSSA & SSSA International Annual Meeting – Meeting 2015 – Poster Presentation

M. Elarab, A. Torres-Rua, M. McKee (2014). Oral presentation, Mapping ET at High Resolution using AggieAir Airborne Multispectral Imagery. ASABE International Symposium on "Evapotranspiration: Challenges in Measurement and Modeling from Leaf to the Landscape Scale and Beyond" being held in Raleigh, North Carolina, U.S.A. on April 7-10, 2014 M. Elarab, A. Torres-Rua, M. McKee (2014). Oral presentation, Mapping ET at High Resolution using AggieAir Airborne Multispectral Imagery. ASABE International Symposium on "Evapotranspiration: Challenges in Measurement and Modeling from Leaf to the Landscape Scale and Beyond" being held in Raleigh, North Carolina, U.S.A. on April 7-10, 2014

Al-Arab, Manal; Torres-Rua, Alfonso F; Ticlavilca, Andres; Jensen, Austin; McKee, Mac, “Use of high-resolution multispectral imagery from an unmanned aerial vehicle in precision

Page 148: The Application of Unmanned Aerial Vehicle to Precision

133

agriculture,” Geoscience and Remote Sensing Symposium (IGARSS), 2013 IEEE International”, 2852-2855, 2013, IEEE. M. Al-Arab, A. M. Ticlavilca, A. Torres-Rua, I. Maslova and M. McKee (2013). Oral presentation, Use of high-resolution multispectral imagery to estimate chlorophyll and plant nitrogen in oats (Avena sativa). Presentation at 2013 Fall Meeting, AGU, San Francisco, Calif. December 9-13, 2013. M.N Nimah, M. Al-Arab 2010 (keynote paper): Solutions for Sustainable Irrigated Agriculture and Food Security 2010 presented at the 1st Kuwait - Food security Expo, Kisr- Kuwait 2010 M.N Nimah, M. Al-Arab (2010) Comparison between Irrigation with Saline Water and regular Deficit Irrigation on Potato Yield presented at MERC Potato Symposium Larnaca, Cyprus, 2010