You are on page 1of 110

1

Final Report to NASA


Decisions/05-2-0000-0167 and 0119
March 2011

National Drought Monitoring System for


Drought Early Warning Using Hydrologic
and Ecologic Observations from
NASA Satellite Data

Research Team

Jet Propulsion Laboratory: S. V. Nghiem (PI), E. G. Njoku, G.


Neumann, and S. K. Chan
U.S. Geological Survey: J. Verdin (co-PI), J. Brown, and Y. Gu
National Drought Mitigation Center: D. A. Wilhite (Co-I,
Institution Lead), M. J. Hayes, M. D. Svoboda, B. D. Wardlow,
and T. Tadesse
NOAA Physical Science Division: R. Dole (Co-I, Institution
Lead), B. Liebmann, D. Allured
Dartmouth College: G. R. Brakenridge (Co-I, now at CSDMS,
INSTAAR, University of Colorado)
NOAA Climate Prediction Center: D. LeComte (Collaborator),
M. Rosencrans
2

Section 1 (Led by Jet Propulsion Laboratory)

Microwave Remote Sensing of Soil Moisture

S. V. Nghiem1, D. Allured2, M. D. Svoboda3, B. D. Wardlow3, D. LeComte4, M. Rosencrans4, S.

K. Chan1, and G. Neumann1

1
Jet Propulsion Laboratory, MS 300-227

California Institute of Technology

4800 Oak Grove Drive, Pasadena, CA 91109, USA

2
Earth System Research Laboratory, Physical Science Division

National Oceanic and Atmospheric Administration

325 Broadway, Boulder, CO 80305, USA

3
National Drought Mitigation Center

University of Nebraska-Lincoln

3310 Holdrege Street, Lincoln, NE 68583, USA

4
Climate Prediction Center, Room 604

National Oceanic and Atmospheric Administration

5200 Auth Road, Camp Springs, MD 20746, USA


3

Abstract

This chapter focuses on soil moisture measurements using wide-swath microwave sensors

including satellite passive radiometer and active scatterometer. First, a review is presented on

the background research and state-of-the-art methods for remote sensing of soil moisture.

Emission and scattering physics is then presented to explain the underlying science of the

different approaches of passive and active estimates of soil moisture. Results are obtained at

local, regional, and continental scales to show patterns of soil moisture variations in daily,

weekly, and seasonal time scales. Practical applications to assessments of drought conditions are

illustrated in specific examples in the United States and other countries. For these applications,

soil moisture change is an important parameter to address problems of anomalous propagation in

radar meteorology and to identify issues related to dry rain or virga, which is relevant to

hydrological drought monitoring and forecasting.

1. Introduction

Soil moisture is a fundamental link between global water and carbon cycles, and has major

applications in predicting natural hazards such as floods and droughts [National Research

Council, 2007]. Precipitation data can be used to estimate soil wetness. In the United States

(US), preliminary precipitation data are based on measurements gathered from many active

stations nationwide each month, and it takes three to four months to assemble final, quality-

controlled data. In western US, some climate divisions may have no stations reporting in a

particular month or may lack a first- or second-order station altogether. A first-order station
4

reports all climate variables three times per day, and a second-order station reports maximum

and minimum temperatures and rainfall once per day and other variables twice per day. For

direct soil moisture measurements, a limited number of observations are provided from the

Oklahoma Mesonet System [Illston et al., 2004] and the Soil Climate Analysis Network [USDA,

2009a], but the data are generally too sparse spatially with different data quality and accuracy for

their assimilation to be of much use. Also, there are soil-moisture observations from enhanced

additions in the SNOTEL network [USDA, 2009b], but fully calibrated data are not yet available

routinely. Given the very limited number of stations collecting point-based, in-situ data, the use

of such information may not be representative of regional water or soil moisture amounts and

variations.

Soil moisture measurements over large spatial extent (areal data rather than point data) with

little or no missing gaps are crucial to represent land surface water distribution from regional to

continental scales. Recognizing the importance of soil moisture, especially as a key variable for

drought monitoring, the void in adequate soil moisture information in the U.S. and elsewhere

demands immediate and extensive measurements from satellite microwave remote sensing of

soil moisture with both passive and active sensors. A radiometer is used to measure the natural

emission from land surface in passive remote sensing, while a radar, including synthetic aperture

radar (SAR) and scatterometer, is used to transmit signals to a targeted surface area and measure

the scattering return in active remote sensing. From passive or active microwave data, soil

moisture can be estimated with various algorithms.

Passive methods use data from microwave radiometers such as the Scanning Multichannel

Microwave Radiometer (SMMR), Special Sensor Microwave Imager (SSM/I), the Tropical

Rainfall Measuring Mission (TRMM) Microwave Imager (TMI), the Advanced Microwave
5

Scanning Radiometer on the Earth Observing System (AMSR-E), and the Soil Moisture and

Ocean Salinity sensor (SMOS) [Wang, 1985; Owe et al.,1988; Kerr and Njoku, 1990; Teng et al.,

1993; van de Griend and Owe, 1994; Engman, 1995; Jackson, 1997; Kerr et al., 2001; Njoku et

al., 2003].

For active remote sensing, many approaches have been used for various datasets from SARs

including Seasat, Spaceborne Imaging Radar-C (SIR-C), European Remote Sensing (ERS),

RADARSAT, Environmental Satellite (Envisat), and Advanced Land Observing Satellite

(ALOS) [Blanchard and Chang, 1983; Dubois et al., 1995; Cognard et al., 1995; Shrivastava et

al., 2009; Loew et al., 2006; Takada et al., 2009], and from scatterometers such as ERS and

QuikSCAT [Wagner et al., 1999; Wagner and Scipal, 2000; Nghiem et al., 2000]. In this

chapter, we review the science principle of active and passive remote sensing of large-scale soil

moisture and illustrate the results from AMSR-E and QuikSCAT for drought applications.

2. Remote Sensing Science

The principle of microwave remote sensing of soil moisture is based on the sensitivity of soil

permittivity to the amount of liquid water. The permittivity of a medium, like moist soil,

characterizes electromagnetic wave propagation and attenuation in the medium. Both brightness

temperature (measured by a radiometer) and backscatter (measured by a radar) are dependent on

the soil permittivity. Empirical models of dielectric constant, which is the permittivity relative to

that of the free space, for different soil types as a function of volumetric moisture content (mv) at

microwave frequencies between 1.4 and 18 GHz was developed [Hallikainen et al., 1985;

Dobson et al., 1985].


6

While soil dielectric constant can be measured in-situ with a probe [Jackson, 1990], satellite

remote sensors do not directly provide soil dielectric measurements. Instead, these sensors

acquire brightness temperature or backscatter signatures, which are dependent on soil dielectric

and thus on soil moisture. Such a relationship enables the inversion of soil moisture from

brightness temperature or backscatter data, which can be complicated by natural effects such as

vegetation cover, surface roughness, rainfall, and anthropogenic effects such as radio-frequency

interference (RFI), which have different impacts on the accuracy of soil moisture retrieval at

different microwave frequencies.

2.1. Passive Remote Sensing

The retrieval of soil moisture from brightness temperature, measured by a satellite passive

radiometer, was long formulated by many researchers and has been summarized by Njoku et al.

[2003]. Here is a brief review for an isothermal vegetated soil surface, whose brightness

temperature Tbp at the physical temperature Ts is expressed as

Tbp = Ts { esp exp(c) + (1 p) [ 1 - exp(c) ] [ 1 + rsp exp(c) ] } (1)

where the soil emissivity is esp = 1 rsp for soil reflectivity rsp, which is influenced by soil

moisture through the effect of moisture on the soil dielectric constant. The emitting depth is

controlled by the near-surface moisture profile, and is smaller for higher microwave frequencies

and for wetter soils. Although microwaves can only sense soil moisture in the top surface layer,

there is a correlation to soil moisture in deeper soil at night when the soil moisture and

temperature profiles are more uniform than in the early afternoon. In (1), c and p are the

vegetation opacity and the vegetation single scattering albedo, respectively. Multiple scattering
7

in the vegetation layer is neglected, and a quasi-specular soil surface and no reflection at the air-

vegetation boundary are assumed in (1). Vegetation opacity and multiple scattering have lesser

effects at lower microwave frequencies.

For surface roughness, based on a fixed viewing angle, an empirical formulation has been

found practically useful for relating the reflectivity of a rough soil surface, rsp, to that of the

equivalent smooth surface, rop [Wang and Choudhury, 1981; Wang, 1983], which is expressed

as:

rsp = [ (1Q) rop + Q roq ] exp(h) (2)

where p and q represent either of the orthogonal polarization states (V or H), and Q and h are

roughness parameters. Q may be approximated as zero at low frequencies (e.g., L and C bands).

The simple separation of soil moisture and roughness effects through (2) is not precise, and the

parameter h has a residual moisture dependence [Li et al., 2000; Wigneron et al., 2001].

To normalize the surface temperature (Ts) dependence in (1), the polarization ratio (PR) is

obtained by:

PR = ( Tbv - Tbh ) / ( Tbv + Tbh ) (3)

which is suitable for multi-channel data taken at the same incidence angle [Kerr and Njoku,

1990]. At large incidence angles (e.g., >50o), the difference between the vertically and

horizontally polarized brightness temperatures for bare soils is large, giving rise to a large PR

signal. Nevertheless, the observation path length through the vegetation becomes longer at large

incidence angles, increasing the vegetation attenuation and thus decreasing the sensitivity to the

soil moisture.
8

While (1-3) form a general theoretical basis for soil moisture retrieval from passive

microwave data, different approaches have been used for different satellite datasets with

different correction methods for the effects of soil types, roughness, vegetation, and surface

temperature [Njoku et al, 2003] although further advances should be considered for various non-

isothermal conditions and multiple interactions between soil surface and vegetation cover at

different growth stages. In practice for data from the Advanced Microwave Scanning

Radiometer on the EOS Aqua satellite (AMSR-E), the soil moisture retrieval utilizes primarily

the frequency channels of 10.7 and 18.7 GHz to consider the effects of atmospheric and

vegetative attenuation and to minimize the requirement for ancillary data inputs. The Tropical

Rainfall Measuring Mission (TRMM) Microwave Imager (TMI) has 10.7 and 19.3-GHz

channels, which can be used to obtain PR for soil moisture applications with a better consistency

at the lower frequency [Njoku et al, 2003]. Further details of the retrieval were published [Njoku

and Li, 1999; Njoku et al, 2003; Njoku, 2004].

2.2. Active Remote Sensing

In active remote sensing, soil moisture can be derived from backscatter measured by a radar such

as a SAR at a high spatial resolution with a small and infrequent coverage, and by a

scatterometer at a low spatial resolution with a large and frequent coverage. Many theoretical

models (references given in the introduction) have been developed to characterize backscatter

signatures of vegetated soil. Here, a scattering model, based on the analytic vector wave theory

[Nghiem et al., 1993a], and a practical formulation are reviewed.

Backscatter 0 from moist soil with a vegetation cover is determined from an ensemble

average of the correlation of scattered field components E as follows:


9

(4)

where subscript 0 represents the air space above the vegetation, and subscript 1 indicates the

vegetation cover occupying volume V1 over the soil surface. The dyadic Green's function G and

the mean field F are obtained as described by Nghiem et al. [1990]. The correlation function C

characterizes the vegetation scatterers having different size, shape, and orientation angle f in

elevation and f in azimuth. For the vegetation canopy, the effective permittivity is calculated

under the strong permittivity fluctuation theory, which accounts for wave attenuation including

scattering and absorption loss [Nghiem et al., 1993a]. The analytic vector wave theory accounts

for fully polarimetric scattering, preserves the phase information, and includes multiple

reflection and transmission interactions of up-going and down-going electromagnetic waves with

the soil surface, thereby conveying the soil moisture information. This is because the soil

transmissivity and the soil reflectivity are controlled by the soil dielectric constant as a function

of volumetric soil moisture.

Rough surface scattering can be included in the contribution to the total backscatter. The

small-scale roughness of the soil surface is described with a standard deviation height and a

slope. When a large-scale roughness also exists, the overall roughness is accounted for by a joint

probability density function for both roughness scales [Nghiem et al., 1995]. The vegetation

volume scattering and the soil surface scattering are assumed to be uncorrelated due to

independent statistic representations of vegetation scatterers (e.g., leaves, twigs, branches) and

soil surface roughness. Then, the total backscatter is a sum of the vegetation volume backscatter
10

and soil surface backscatter. In the layer scattering configuration, such as a vegetation layer over

a rough soil surface, contributions from the rough surface scattering are considered with wave

interactions, differential propagation delay, and wave attenuation in the vegetation layer [Nghiem

et al., 1995], which can be effectively anisotropic when vegetation scatters have a preferential

directional structure (e.g., planophile, plagiothile, erectophile, or extremophile orientation

distribution) [Nghiem et al., 1993b]. The backscatter from rough soil surface depends strongly

on the soil dielectric constant and also on transmissivity and reflectivity due to wave interaction

with the soil boundary, and thus the surface scattering also contains soil moisture signature in

addition to the soil moisture information in the volume scattering component. Nevertheless, a

dense vegetation canopy can have a large imaginary part in its effective permittivity, which

attenuates both the soil surface scattering and soil interactions in the volume scattering, and

consequently masks the soil moisture signature. Specific mathematical details of the volume and

surface scattering in layered media can be found in earlier publications [Nghiem et al., 1990,

1993a, 1993b, 1995].

While the above formulation provides physical insights and serves as a theoretical basis for

active remote sensing of soil moisture, in practice, due to the complexities of natural

environments in different climate regimes, it is not possible to set up an inversion method of soil

moisture strictly based on theoretical modeling of electromagnetic scattering. A simple empirical

linear equation relates backscatter 0 to volumetric soil moisture mv as

0 = amv + b (5)

where coefficients a and b are dependent on incidence angle, polarization, vegetation conditions,

soil type, surface variation, and climate regime [Mo et al., 1984, Prevot et al., 1993, Shrivastava
11

et al., 1997; Shoshany et al., 2000; Hutchingson, 2003]. Particularly for Ku-band backscatter

data from the SeaWinds scatterometer aboard the QuikSCAT satellite (QSCAT), the bias term b

in (5) contains a signature of seasonal vegetation change while changes in volumetric soil

moisture mv from rainwater are detectable in backscatter variations in a time scale consistent

with the initial impulse increase in wetness from the precipitation input throughout the

subsequent drying process [Nghiem et al., 2005]. Thus, soil moisture change (SMC) can be

directly inferred from (5) using the temporal backscatter-change method, which removes most of

the background bias.

2.3. Passive and Active Blending

As presented in sections 2.1 and 2.2, passive and active sensors measure different parameters:

passive brightness temperature and active radar backscatter, which represent different responses

to vegetated soil with different sensitivities to soil moisture and vegetation cover.

In a theoretical ideal case of smooth bare soil (c = 0), (1) dictates that the brightness

temperature is directly proportional to the emissivity esp, which is determined by soil dielectric

constant and is thereby most sensitive to soil moisture. As opposed to the passive signature in

this idealized case, there is no backscatter since there is no vegetation (V1 = 0 in (4) without

vegetation) and no rough surface; hence, the soil moisture is not measurable by a radar for bare

soil without any roughness. In reality, there is some surface roughness or some vegetation cover

on moist soil, which affects the sensitivity to soil moisture differently in passive data [Njoku et

al., 2003] and in active data [Nghiem et al., 1993a] until the vegetation cover becomes

sufficiently dense to start masking the soil effects. Therefore, information conveyed in passive
12

and active signatures together may yield more comprehensive results compared to the capability

of each one separately.

As an illustration of the passive and active blending, correlation analysis results of observed,

satellite-based remote sensing signatures versus in-situ soil moisture and vegetation

measurements at a U.S. Department of Agriculture (USDA) Natural Resources Conservation

Service (NRCS) Soil Climate Analysis Network (SCAN) site in Lonoke, Arkansas (91.867oW

and 34.833oN) is presented. The vegetation cover in the Lonoke region is primarily agricultural

crops including soybeans (41%), rice (21%), and wheat (11%) [Njoku et al., 2003]. TMI passive

microwave data at 10.7 and 19.3 GHz were collected for all points with their centers located

within 25 km of the Lonoke SCAN site location. More than a year of time series data in 1999-

2000 were analyzed. Results show a wide range of sensitivity in the response of instantaneous

PR (obtained at each local overpass time) and in that of transient SMC after rain events. The

variance between measurements and linear fit values of daily PR (10.7 GHz) versus the

contemporaneous daily mv becomes large at larger values of soil moisture so much that PR can

be different by a factor of 3 at mv = 34%, or transient soil moisture can change between 6% to

34% for the same PR around 0.017. This is consistent with the conclusion that a number of

transient soil moisture events recorded in the SCAN data are not evident in the TMI data [Njoku

et al., 2003], from which the retrieved soil moisture can be inconsistent among various transient

precipitation events.

For seasonal trends, seasonal TMI PR (90-day running average) is well correlated with

seasonal soil moisture (90-day running average) measured at a depth of 5 cm at the Lonoke

SCAN site, as indicated by high values of correlation coefficient (Table 1). Plots of seasonal data

(90-day running average) for contemporaneous SCAN mv and TMI PR at both frequencies reveal
13

a hysteresis behavior (Figure 1). Theoretically, (1) suggests that the hysteresis is caused by

attenuation effects on the passive microwave signatures under different vegetation conditions in

different seasons. In Figure 1, the linear fit for all data in the whole year is used as a reference

for each frequency. The polarization ratio PR is mostly above the annual linear fit in fall and

summer seasons from fall equinox in 1999 to spring equinox in 2000, when vegetation starts to

decrease in early fall down to a minimum in winter and then increases toward the spring

equinox. In contrast, PR is mostly below the annual linear fit during spring and summer from

spring equinox in 2000 to fall equinox in 2000, when vegetation increases in spring to a peak in

summer then slightly decrease toward the fall equinox. In different seasons, PR can be larger for

less vegetation in winter and smaller for more vegetation in summer, thereby causing the

hysteresis in seasonal PR versus mv at both frequencies. Also, the vegetation attenuation effects

are less severe at the lower frequency. This is evident by two distinctive characteristics in the

plots at 10.7 and 19.3 GHz (Figure 1): steeper slope at 10.7 GHz, and larger hysteresis at 19.3

GHz. Note that the annual correlation coefficient declines significantly from 0.936 at 10.7

GHz to 0.792 at 19.3 GHz because the hysteresis is much larger, corresponding to strong

vegetation attenuation at the higher frequency. Thus far, this discussion has been theoretical

since the passive microwave analysis has not included any independent data characterizing the

vegetation cover, which needs to be addressed.

For the active microwave analysis, time-series QSCAT data are extracted within 25 km

around the same SCAN site in the same manner as done in the case for TMI data. In contrast to

the passive microwave case, daily QSCAT backscatter change correlates well with

contemporaneous SMC from rainwater. In fact, daily QSCAT data capture 91% of the rain

events recorded at the Lonoke SCAN site in 1999-2000. To illustrate the high correlation of
14

QSCAT backscatter to transient soil moisture, a regression analysis is carried out for the period

of 4 October to 19 November 1999 when two major rain events occurred as observed in daily in-

situ soil moisture measured at the SCAN site. With the linear formulation in the inverted form of

(5) such that mv = a'0 + b' for backscatter 0 in dB and mv in percent, a' = 8.921%/dB and b' =

111.11% with a high correlation coefficient of 0.907 and a small standard deviation of 3.7% for

the case of backscatter at the horizontal polarization (0HH), which well represents both the initial

impulse of soil moisture increase from rain and the subsequent soil moisture decrease in the

entailed drying process. For backscatter at the vertical polarization (0VV), the result is similar

with SMC of 8.369% for a dB change in 0VV, and thus the backscatter at the vertical polarization

is slightly less sensitive than that at the horizontal polarization to transient soil moisture. This is

consistent with (4) in which the dyadic Greens function includes soil reflection which is

stronger at the horizontal polarization compared to that at the vertical polarization due to the

Brewster effects. Also, the incidence angle at 54o for 0VV is larger than 46o for 0HH, which

means that 0VV suffers from higher attenuation effects due to the longer path length in the

vegetation cover. Nevertheless, QSCAT data can identify rain events even in peak vegetation

conditions when the rain is sufficiently heavy for the backscatter to increase above the seasonal

level of the background backscatter, explaining the high percentage of success in the QSCAT

capability for identification of transient SMC. This capability to capture transient rain events

illustrates the complimentary information that active microwave remote sensing provides in

combination with the seasonal SMC that can be estimated from passive microwave remote

sensing to provide a better representation of soil moisture conditions when both types of satellite-

based microwave data are combined or blended.


15

Regarding seasonal trends, active backscatter data primarily contain vegetation information.

To verify this, normalized difference vegetation index (NDVI) data representing seasonal

vegetation change [Justice et al. 1985; Verdin et al., 1999; Zhang et al., 2010] are used to

compare with seasonal QSCAT backscatter data (90 running average). NDVI is defined by

NDVI = (NIRVIS) / (NIR+VIS), where NIR is the near infrared band and VIS is the visible band

from a spectral sensor such as the Advanced Very High Resolution Radiometer (AVHRR)

aboard the National Oceanic and Atmospheric Administration (NOAA) operational polar-

orbiting satellites. AVHRR NDVI data are collected and averaged within 25 km around the

SCAN site so that the spatial scale of AVHRR NDVI data become compatible with that of

QSCAT data. Results from the linear regression analysis of contemporaneous QSCAT

backscatter and AVHRR NDVI around Lonoke show a very high correlation coefficient of 0.946

for linear 0VV and a lower correlation coefficient of 0.864 for linear 0HH. Therefore, seasonal

QSCAT backscatter can be used to characterize seasonal vegetation change regardless of cloud

cover, which is transparent to QSCAT at the Ku-band frequency of 13.4 GHz. This is consistent

with earlier results on the relation of Ku-band backscatter with NDVI [Moran et al., 1997], with

green leaf area index [Moran et al., 1998], and with above-ground biomass [Nghiem, 2001].

Here, QSCAT data, conveying vegetation change information independent of the passive

data, are used to substantiate the theoretical discussion of vegetation cover effects on passive

microwave presented earlier. For this purpose, seasonal running averaged QSCAT 0VV (more

sensitive to seasonal vegetation change compared to 0HH) and TMI PR at 10.7 GHz (more

sensitive to seasonal SMC compared to 19.3 GHz data) are collocated in time and partitioned in

fall-winter seasons and spring-summer seasons. The hysteresis behavior is clearly observed in

the curve of 0VV versus PR (Figure 2). In fall and winter, PR is below the annual linear fit,
16

corresponding to less vegetation cover as compared to spring-summer PR above the linear fit

with more vegetation cover. The less vegetation cover indicated by lower backscatter in fall and

winter supports the fact that PR is above the annual linear fit in the PR-mv hysteresis (Figure 1)

for less vegetation attenuation effects on PR. Similarly, the more vegetation cover characterized

by higher backscatter in spring and summer confirms the stronger vegetation attenuation forcing

PR to the lower section of the PR-mv hysteresis below the annual linear fit (Figure 1). In terms

of timing, the vegetation peak observed in 0VV occurs in summer after the seasonal soil moisture

reaches the maximum seen in PR in spring. Thus, independent information on seasonal

vegetation change in active backscatter data is complementary to the information on seasonal

SMC in passive polarization ratio data. The active and passive combination can thereby provide

different information that will be missed if only one type of data is used separately.

3. Drought Applications

3.1. Drought Monitoring Issues

In U.S., two major factors have been identified as important impediments to early detection and

monitoring of drought and its impacts at the county level where key drought-related decisions are

made: the lack of accurate and objective data source and the coarse level of spatial detail in

drought analyses and results [Nghiem et al., 2010]. These issues are even more severe in many

other countries across the world where there is a significant lack of both data and drought

monitoring systems.

The Lincoln Declaration on Drought Indices in 2009 declared a consensus that the

Standardized Precipitation Index (SPI) be used to characterize the meteorological droughts


17

around the world [World Meteorological Organization, 2009]. The U.S. Drought Monitor

(USDM) [Svoboda et al., 2001, 2002] uses SPI, which is based on preliminary precipitation data

obtained from surface observations from the NOAA Climate Prediction Center (CPC) and

National Climatic Data Center (NCDC). The use of rain gauge data from stations (point

measurements) is a hit-or-miss approach and may not be representative of regional rainfall

amounts, and rain gauge data can also be in considerable error [Story, 2009]. Rain radars may

provide better coverage for rain rate estimations; however, surface radars suffer from anomalous

propagation (AP) problems [Smith et al., 1996] resulting in inaccurate representations of

rainwater. Satellite precipitation estimates can provide nearly global coverage [Kummerow et

al., 1996; Sorooshian et al., 2000] without AP problems in surface radar meteorology; however,

the temporal and spatial resolutions are coarse.

A more fundamental issue is that, while precipitation data are useful for assessing

meteorological drought, such data may not represent rainwater that actually reaches to land

surface and accumulates in soil, especially in drought-prone regions where virga (dry rains, dry

thunderstorms, or dry lightning) are prevalent. The United Nations World Meteorological

Organization stressed the need for undertaking a comprehensive review to develop common

indices for better early drought warnings in the agricultural and water sectors [World

Meteorological Organization, 2009]. For hydrological and agricultural drought assessment and

monitoring, water on land surface and in soil is the most relevant, and thus soil moisture data

must have a key role. Nevertheless, the current in-situ station network is inadequate, and soil

moisture measurements are too sparse for effective use or even non-existent in many areas

[NIDIS, 2007].
18

For county-level monitoring, an important goal of the National Integrated Drought

Information System (NIDIS) [Western Governors Association, 2004; NIDIS 2006, 2007], the

National Weather Service (NWS) has determined that an effective Cooperative Observer

Network would require a minimum spatial density of one observing site per 1000 km2 across the

country, or a separation of about 24 to 32 km [NIDIS 2007]. The location of each in-situ sensor

must be carefully selected such that the measured soil moisture is representative of the

surrounding area. Furthermore, consistency and persistency in data collections are important in

terms of data quality and data availability across different agencies and across different nations.

3.2. Uses of Satellite Data

In view of the above issues in drought monitoring, recent efforts have enabled certain use of soil

moisture measurements derived from satellite remote sensing data for enhancing drought

monitoring systems [Nghiem et al., 2010]. Several specific results are presented in this section to

illustrate various uses of satellite data with different temporal and spatial scales.

3.2.1. Temporal Data at Local Scale

Regarding the issue of the lack of accurate and objective data from limited in-situ station

measurements, we use temporal QSCAT observations together with in-situ station measurements

in the local station area to illustrate how satellite data can help to enhance the drought monitoring

capabilities. Figure 3 presents results at the NCDC Global Summary Of the Day (GSOD)

Station 727760 in Great Falls, Montana (47.467oN, 111.383oW). The NCDC GSOD station

dataset consists of meteorological data measured by weather stations in the World

Meteorological Organization and Global Telecommunications System network, including

precipitation, temperature, humidity, dew point, pressure, and other meteorological data
19

measured by a global network of weather stations [NCDC, 2010a]. Time series of QSCAT data

together with in-situ measurements around this station are constructed with the Special Satellite-

Station Processor (SSSP) [Nghiem et al., 2003]. SSSP uses an innovative algorithm based on

pointers in computer coding, which enables rapid satellite data extraction and collocation with

in-situ data over numerous global stations represented by a global station mask allowing satellite

data selection to any radius within 60 km around each station. In Figure 3, daily QSCAT data

whose centroids within 25 km around Station 727760 are selected for the horizontal polarization

(more sensitive to soil moisture compare to the vertical polarization) from ascending orbits (~6

am local overpass, and more correlated with soil moisture compared to data from descending

orbits).

QSCAT 0HH data around Great Falls (top panel in Figure 3) clearly identify rain events

before and after the dry period between 9 July and 5 September 2000 when there was very little

rain. In August 2000, the long-term Palmer Drought Index for the region around Great Falls was

4 or below, indicating long-term extreme drought conditions. The U.S. Department of

Agriculture (USDA) issued Natural Disaster Determinations for drought for the entire state of

Montana in 2000, when severe and persistent drought caused significant losses to agriculture and

other sectors [Resource Management Service, 2004]. The summer drought period observed by

QSCAT is validated by the lack of rain in in-situ precipitation data (bottom panel in Figure 3)

from Station 727760 in the same period, when there were several heat waves indicated by spells

of high temperatures (bottom panel in Figure 3). Both before and after the drought period in the

summer of 2000, QSCAT detected a number of significant rain events which increased

backscatter by about 3 dB, which is equivalent to 26.8% in volumetric soil moisture increase per

the Lonoke rating value of a' = 8.921%/dB. Thus, rainwater from these rain events reached the
20

land surface and significantly increased the moisture in soil. However, precipitation data from

the station rain gauge corresponding to these significant rain events disparately ranged across

one order of magnitude from low values (< 0.2 cm) to high values (> 2.0 cm).

In contrast to SMC, which has a consistent relationship with backscatter change (section 2.3),

the inconsistency in in-situ precipitation data measured by a rain gauge at this station

demonstrates accuracy issues causing difficulties in assessing drought conditions in an objective

manner, especially for hydrological droughts. The inconsistency of precipitation data compared

to SMC exists not only in satellite measurements of SMC but also in in-situ data of SMC from

SCAN, which also has rain gauges to measure local precipitation. This problem can be

compounded by the measurement location where data may be affected by localized effects (e.g.,

wind effects, shading effects, equipment maintenance) and by the point-based nature of rain

gauge measurement itself where a local rainfall event may not be recorded if the rain does not

fall over the exact station location. These issues can be addressed with the satellite data to

enhance the drought monitoring in an objective and consistent manner. Moreover, similar time

series of QSCAT signature can be obtained at all grid cells for a full and continuous spatial

coverage while station data are only available at local station points, which may or may not be

representative of the surrounding area.

3.2.2. Spatial Data at Regional Scale

Satellite microwave remote sensing data, such as AMSR-E or QSCAT, can be used to monitor

drought and water resources at regional to global scales. Both have swath widths of 1400 km or

larger [Njoku et al., 2003; Tsai et al., 2000], which allows a nearly daily coverage across the

world and as many as two times per day at high latitudes. Several attributes related to water can
21

be obtained from microwave satellite data for drought monitoring from local areas to large

scales. Results are presented here to illustrate the use of satellite data for Texas (TX), an

important agricultural state. Pertaining to the regional climatic regime, drought is inevitable in

TX, where water demand will significantly increase and drought will become more critical

[Texas Water Development Board, 2007].

A relevant attribute for water resource and drought assessments is precipitation frequency,

which quantifies the recurrence of rain events in a given period [Gonzlez and Valds, 2004].

Instead of an apparent precipitation frequency (APF) derived from in-situ rain gauge data or

surface rain radar data, a different measure of precipitation frequency is derived from satellite

scatterometer data. This measure is defined as effective precipitation frequency (EPF) because it

accounts for rainwater that effectively reaches land surface and increases soil moisture as

opposed to APF that may have problems with AP, virga, or inconsistent point data. For

applications to QSCAT data, EPF = 100 (NW / NC) is defined as the percentage of the number of

wet days (NW), when soil moisture increase in the top soil layer (5 cm) is 5 % such that the

corresponding backscatter increase is 0.56 dB for 0HH or 0.60 dB for 0VV above the

background level, over the total number of satellite coverage days (NC) without counting missing

data days in a given period.

EPF is retrieved from QSCAT data across the state of TX over the period of 1 June to 31

August 2009 (left panel in Figure 4). In summer 2009, TX suffered exceptional drought over

most of the southern region of the state as observed in the USDM maps from June to August

2009 (right panels in Figure 4). By August 2009, extreme and exceptional drought conditions

(D3 and D4, respectively) remained persistent across south central TX, where the topsoil

conditions were very dry and river levels were near historic lows [NCDC, 2009]. Consistent
22

with these drought conditions, QSCAT EPF showed little to no rain event across most of

southern TX (black to magenta areas, left panel of Figure 4). In contrast, the soil in a part of the

TX Panhandle was shown to be wetted by rainwater the most often during this time period (light

blue to green and yellow areas, left panel of Figure 4), where abnormally dry (D0) depicted in

the June USDM map had improved to no drought by late August 2009.

While EPF carries information on wet precipitation frequency or how often land surface

becomes wet due to rainwater, daily SMC from QSCAT data represents the quantitative change

in soil moisture or the amount (intensity) of rainwater that accumulates on land surface each day.

Therefore, SMC is an attribute relevant to monitoring hydrological drought because it is related

to water on land rather than rain drops in the atmosphere (a meteorological parameter). Figure 5

presents maps of daily SMC compared to the semi-monthly average over the state of TX from

early September to early October 2009. Intense SMC (yellow areas in maps), which reflect large

increases in soil moisture conditions occurred across large areas of central TX on 10, 11, 13, and

14 September, as well as 4 October. The SMC results on these days are consistent with torrential

rainfall events reported across central and south TX (rainfall totals up to 20 inches were

recorded in some locations)in September 2009, causing flash flooding [NWS, 2009]. With this

new water input, drought conditions in central and south TX had significantly improved by early

October 2009 (as shown in the USDM map for October 6 in Figure 6).

Complementary to the transient change observed in the QSCAT daily SMC, AMSR-E

passive microwave data provide good measurements of seasonal soil moisture (as discussed in

section 2.3). Figure 6 (left panel) shows the difference in seasonal soil moisture between the

September-October and June-July periods in 2009. AMSR-E seasonal soil moisture results

reveal a large region of increased soil moisture in south and south central TX (blue areas). This
23

corresponds to the marked improvement in drought conditions in September compared to that in

July 2009 as seen on the USDM drought maps (right panels in Figure 6). In contrast, an area in

western TX has a substantial reduction in soil moisture by the September-October period (red-

brown areas in left panel of Figure 6) compared to more moist conditions in the June-July period.

This area had a larger EPF observed by QSCAT in the earlier months as seen in the left panel of

Figure 4 for June-August 2009. Therefore, the independent attributes derived from different

remote sensing datasets (QSCAT and AMSR-E) are consistent and thereby cross-verifying the

results. Although the USDM map (lower right panel in Figure 6) suggests a slight improvement

by 6 October 2009, this improved condition was a result of recent transient wetting events

observed in daily SMC from QSCAT (e.g., SMC map for 4 October 2009 in Figure 5). The

above results demonstrate the capability and consistency of different parameters to depict the

state of soil moisture and its transient as well as seasonal changes, which are relevant to drought

monitoring at the regional scale.

3.2.3. Spatial Data at Continental Scale

A major advantage of satellite data is its large spatial coverage across the continental scale to the

global scale compared to surface in-situ measurements from station networks. Here, the pattern

of soil moisture change, observed by QSCAT and AMSR-E satellites, is examined across the

contiguous United States (CONUS) and compared to rainfall patterns from the regional multi-

sensor precipitation analysis assembled into a national product (stage 4). Maps of stage-4 daily

precipitation (SDP) are available from the National Mosaic and Multi-Sensor QPE [NMQ, 2009].

With large swaths of measurements, QSCAT and AMSR-E can provide coverage over

CONUS on a near daily basis. Nonetheless, data gaps exist and a full coverage of the entire
24

CONUS is not possible within a day, especially when ascending and descending orbit data are

used separately. Figure 7 includes daily SMC maps in May 2009 from QSCAT ascending-orbit

data (upper panel) at about 6 am local overpass time, and from AMSR-E descending-orbit data

(lower panel) at about 1:30 am local overpass time. Both maps have data gaps, which are larger

in AMSR-E data due to its narrower swath compared to that of QSCAT at the vertical

polarization. There are also missing AMSR-E data in most of West Virginia due to invalid

conditions (e.g. vegetation cover) for soil moisture retrieval.

Overall, the patterns of daily SMC from QSCAT and AMSR-E are similar. Both reveal

precipitation water on land surface in the Midwest and the Great Lakes states extending toward

the northeastern U.S., whereas most of the western U.S. was dry. An extensive wet region is

observed across Kansas (KS) and Nebraska (NE) in both SMC maps (marked by the circles in

Figure 7). Interestingly, a well-defined dry area is detected by both QSCAT and AMSR-E just

south of Lake Michigan at the Illinois-Indiana border. Nevertheless, there are discrepancies.

First, the amount of SMC observed by QSCAT can be more than volumetric 10 % (yellow areas

in the upper panel of Figure 7) compared to AMSR-E SMC that barely exceeds 5 % (blue areas,

lower panel, Figure 7). In New York, the region east of Lake Ontario had a large positive SMC

(wet) in the QSCAT map while the AMSR-E SMC showed a slightly negative value (dry).

These differences, given the better sensitivity of QSCAT data to transient SMC, are not

surprising as discussed earlier in section 2.3.

In the case of the discrepancy between QSCAT and AMSR-E SMC in New York, it could be

hypothesized that the difference was due to the different observation times of the two

instruments (6 am for QSCAT and 1:30 am for AMSR-E). However, SDP maps indicate

significant rainfall on 27 May continuing to 28 May 2009 in New York (Figure 8). Thus, the
25

lower sensitivity in AMSR-E data to transient SMC is likely the cause of the differences in SM

results. The SDP map on 28 May 2009 (Figure 8a) also shows a large-scale overall pattern

similar to the SMC observed by both satellites with band of heavier rainfall across the upper

Midwest and Great Lakes region extending into the northeastern states. However, the 28-May

SDP map indicates no precipitation in KS and NE where both QSCAT and AMSR-E detected

rainwater on land surface resulting from the intense rainfall on the prior day (marked by circles

in Figures 7 and 8). This case illustrates that SMC can represent the rainwater accumulated from

preceding strong precipitation events so that the corresponding large amount of water can stay in

the top soil for a period of time after the rain events. As such, SMC is also an indicator of the

intensity or amount of rainwater on land surface in terms of the SMC duration.

There are discrepancies between SMC and SDP. For example, in New Mexico (NM), SDP

observed extensive precipitation across the state while SMC from both QSCAT and AMSR-E

found wetness only in some areas of NM (such as in northeastern NM). This difference suggests

either the rainwater did not fully reach land surface (virga problem) or SDP has uncertainties in

surface radar data (AP problem). Similarly, the SDP pattern was much more widespread

compared to the SMC pattern (Figures 7 and 8) in TX, where AP problems can cause significant

difficulties in precipitation mapping [Story, 2009]. These observations suggest that SMC is more

relevant to hydrological drought monitoring while SDP is for meteorological drought

monitoring.

3.2.4. Soil Moisture Products for Drought Monitoring and Forecasting

In an operational environment, science results need to be transitioned into data and image

products with appropriate formats and protocols that can be used by drought experts, such as the
26

USDM authors, in the process of making drought monitoring maps in conjunction with a suite of

attributes from multiple data sources (e.g., vegetation conditions, stream flow, and precipitation).

Here, examples of various SMC products are presented and compared with other traditional

products to identify their advantages and limitations.

Data from the QuikSCAT satellite are processed at the Jet Propulsion Laboratory in

Pasadena, California (CA), which are then automatically and routinely uploaded to the NOAA

Physical Science Division (PSD) in Boulder, Colorado (CO), where multiple SMC products are

generated. For the data link, a file-transfer-protocol (ftp) is established between JPL using an

automated code and PSD using a semi-automatic approach. In the data transfer process, human

interventions are needed to account for anomalies in satellite data collections and operations;

therefore, a fully automated system requires a full integration with the satellite operation and

data processing systems in the future.

The development of various SMC products for use in USDM is an iterative process where

many interim products are made for different attributes of SMC in different formats with various

color protocols. In this process, three SMC attributes, including daily SMC, weekly maximum

SMC, and weekly mean SMC, in five different formats (gif, jpg, png, NetCDF, and kmz) using

five different color protocols, result in 75 different products when all combinations are produced.

Several raster sizes were also considered so that images could be displayed on a drought authors

computer terminal without a missing section in each of the images. The timing of the routine

SMC product delivery is important in the operational environment so that updated results can be

used in time. This timing dictates which period of SMC data should be used to compile weekly

results. For USDM, the weekly SMC products are made ready by Monday afternoon to meet the
27

operational timing requirement for USDM authors to have the inputs by late Monday or early

Tuesday for the preparation of the drought monitor for the current week.

The various SMC products have been evaluated and specific products useful for USDM have

been identified. We present an example in Figure 9, which compares 8-day mean and 8-day

maximum SMC products with the rain-gauge precipitation (RGP) product used in USDM for the

period of 14 October 2008 and the ensuing seven days. The RGP product is made by CPC with

an Oracle database that houses observations from a multitude of sources such as surface weather

measurements from the Automated Surface Observing Systems (ASOS) and from co-operative

observers. These data are run through several quality-control steps and through Geographic

Information System (GIS) processing to obtain RGP maps. About 7000 daily reports of in-situ

rain gauge data are included in the making of the RGP product [Higgins et al., 2000]. RGP maps

are made with different time periods from 5 to 8 days for operational use in the USDM drought

assessment. In this particular example, we compare the full 8-day RGP product with the

corresponding 8-day SMC map. The mean and maximum SMC maps (Figures 9a and 9b), based

on the color protocol with yellow-to-brown for drier conditions and green-to-blue for wetter

conditions, are shown together with the corresponding USDM D-level contours from the USDM

for 14 October 2008. This color protocol is selected to enhance the visibility of the different

SMC levels against the colors of the D-level contours.

The mean SMC map (Figure 9a) reveals a significant value of soil moisture increase (light

blue area) extending from western KS toward its border with CO. In Figure 9a, noticeable

increase in soil moisture (green areas) is observed around the TX Panhandle, in southeastern TX,

central Oklahoma (OK), eastern New Mexico (NM), and in some areas of Montana (MT). The

mean SMC map also shows significant drying across several states in the upper midwest US.
28

Compared to the mean SMC, the maximum SMC map (Figure 9b) indicates much more intensive

soil moisture increase over extensive regions (blue to magenta areas) since it is the peak soil

moisture increase detected at any time in the 8-day period, representing the largest value of any

rainwater detectable on land surface on any given day including the remnant of rainwater from

previous days. In the maximum SMC map, a caveat is that low SMC values (gray and light

green) are noisy and have large uncertainties.

While the maximum SMC corresponds to the peak water accumulation on land surface, the

mean SMC provides an assessment of the persistency of the rainwater in soil because the more

the number of days when a significant amount of soil moisture increase occurs the larger the

mean SMC becomes in the given time period. Therefore, it possible to have a large peak SMC

due to an intensive rain event in a day in an area (e.g., blue area between Indiana and Ohio in

Figure 9b) where the mean SMC is low because there is no rain water accumulation in other days

during the 8-day period. Since the persistency of SMC (how long rainwater accumulates and

stays in soil) depends on soil type, infiltration, and runoff processes among other factors, the

mean SMC carries hydrological information and is thus relevant for hydrological drought

monitoring.

For benchmarking, the traditional RGP product used in USDM is included in Figure 9c to

compare with the mean and maximum SMC products. The comparison of the RGP and SMC

products in Figure 9 clearly points to the different characteristics of these measurements: RGP

consists of point data localized at each separated rain gauge station, while SMC are composed of

25-km pixel with a continuous spatial coverage within each satellite swath. Here, an important

note is that a goal of NIDIS is to resolve county-level drought conditions because many

important drought-related decisions are made at that spatial scale. The average size of a county
29

or a county-equivalent unit for CONUS is approximately 50 km in linear scale (~2,500 km2 in

area). Thus, to resolve the county scale of 50 km, the Nyquist sampling theorem requires a

spatial scale of 25 km, which is satisfied by the SMC data. However, RGP can have hourly data

while SMC is available only two times per day at most. Although SMC temporal scale is suitable

for the weekly time scale of USDM, better temporal coverage can improve the overall result,

especially in the tropics where current satellite data gaps are the largest.

While the spatial patterns in both RGP and maximum SMC maps (Figures 9b and 9c) agree

in general over the areas of extensive precipitation in the states of NM, TX,OK, KS, NE, and

Iowa (IA), there is however a large area of discrepancy between maximum SMC and RGP in

MT, Wyoming (WY), and a part of North Dakota (ND). A reason for this discrepancy is SMC

has a memory of any precipitation water as long as it remains on land surface at the time of the

satellite measurement as opposed to the instantaneous and temporally discrete characteristics of

rain gauge data at the in-situ station location. In particular for the case in Figure 9, the wetness

in the maximum SMC was observed the area of discrepancy only on the first day (14 October

2008). Another difficulty is the sparse number of stations in certain regions (e.g., south TX, MT,

northwest NE, etc.).

Regarding the mean SMC (Figure 9a), a high value requires sufficient rain water to

accumulate on land surface over a significant duration during the period under consideration. As

such, the mean SMC can truly represent the significance in both quantity and persistence of new

precipitation water in soil that are most relevant to hydrological drought monitoring, compared to

both maximum SMC and RGP. For example, maximum SMC and RGP show some precipitation

pattern in NE and IA; however, the same region appears dry in the mean SMC, indicating that the
30

transient rainwater may not be sufficient to sustain the presence of soil moisture over a

significant fraction of the 8-day period to have any significant impact on the overall condition.

For the 2009 growing season (i.e., June to early October), Figures 10 and 11 present a

comparison of QSCAT SMC and USDM results across CONUS. There is an overall consistency

between the two sets of results on a regional scale. For example, there was not much water from

precipitation detected on land surface (Figure 10) in the West and the Southwest, where USDM

maps (Figure 11) show either no improvement (e.g., CA) or more severe drought (e.g., Arizona).

For south TX, no significant wet events occurred in the first part of the growing season, which is

reflected by the severe to extreme drought conditions in the USDM, while the rainfall events in

August and September that lessened the drought conditions are represented by the positive SMC

in September and the corresponding reduction in drought classification in the USDM in

September and early October.

In the Midwest, Figure 10 reveals extensive SMC in South Dakota (SD), southern Minnesota

(MN), eastern NE, and western IA in June and July 2009. During this same time period, USDM

results consistently indicate some improvement in SD, NE, and IA (primarily change from D0 to

no drought classification), and USDM maps suggest drought levels in MN remained unimproved

or became slightly worse. In the first half of June 2009, SDP results showed extensive rain

pattern in the Midwest, which supports the existence of rain water on land seen in SMC (top left

two panels in Figure 10). Since the mean SMC represents a persistent amount of rainwater on

land surface, SMC inherently reflects information about temperature, wind, insolation, and other

parameters that affect the soil wetness. Therefore, SMC may supplement information in synergy

with other parameters currently used in USDM to enhance the results, especially regarding

hydrological drought conditions.


31

In addition, with the participation from the NIDIS team, further advances have been made in

order to demonstrate SMC products in the NIDIS operational environment. For this NIDIS

demonstration, a set of nearly three years of SMC products (2007-2009) are made in the Network

Common Data Form (NetCDF) and delivered for transition into the environment of the NIDIS

portal. The first working prototype is available for the demonstration of SMC maps together

with USDM drought overlay is at http://www.drought.gov/imageserver/NASADroughtViewer/.

Currently working primary features include zoom, pan (via the grab tool), layer selection (both

SMC and USDM and other layers are selectable separately), transparency control for each layer

(via a dialog box), locator by geographic names, etc. Figure 12 illustrates the SMC

demonstration at the national level with USDM and AHP River Gauge layers, and at the county

level in Nebraska as an example. Further work is on going to include more operational features.

Soil moisture change as measured by satellite can benefit not only drought monitoring, but

also benefits drought forecasting. Skillful forecasts of drought or soil moisture would have

significant uses for agriculture and hydrology (water planning). Recognizing the importance of

seasonal forecasts of drought, NOAA CPC has been issuing such forecasts since March 2000.

These forecasts are designed to indicate whether existing droughts will persist or improve, and

whether a new drought will form. An important first step in creating an improved forecast would

be better knowledge of existing conditions, and the SMC is an appropriate hydrological

parameter to contribute to a more accurate depiction of near-surface moisture supplies. Secondly,

improved knowledge of short-term moisture trends can contribute positively to drought forecasts.

Although there is no certainty that short-term trends will persist, forecasters need to know if

moisture conditions are deteriorating, and how fast they are deteriorating. Such trends serve to

flag situations that require additional analysis.


32

4. Concluding Remarks

As presented in this chapter, both passive and active microwave remote sensing have advantages

and limitations for soil moisture monitoring, which necessitates using them in combination with

each other, as well as in situ observations to more fully characterize soil moisture conditions.

First, in-situ measurements can be obtained many times in a day (e.g., hourly measurements)

whereas a satellite sensor typically collects data one or two times per day depending on latitudes.

Most in-situ stations have a longer time-series compared to satellite data. In particular, AMSR-E

data have been collected since 2002 to the present and QSCAT data were obtained over a decade

(1999-2009), while many rain gauge stations were established several decades ago. Regarding

spatial coverage, satellite data, such as seasonal soil moisture from AMSR-E or SMC from

QSCAT, have two important advantages: (1) areal data represent the condition over the pixel

size, and (2) continuous coverage from the local area to regional and continental scales, whereas

in-situ RGP data or soil moisture network such as SCAN consists of sparse point data separated

in space with different numbers of stations and different data quality in different areas.

Furthermore, there are differences in the characteristics of different attributes measured by in-

situ gauges and by satellite sensors as presented in the benchmark study in the previous section.

Given the advantages and limitations in both surface and remote sensing measurements, the

data should be combined in an optimal synergistic manner to improve the quality of both kinds

of data (in-situ versus remote sensing). Here, an approach is to use satellite data to assess the

representativeness of in-situ point measurement in the surrounding area. For example, soil

moisture data from SCAN can be compared or correlated in time (across months, seasons, or

years) to satellite soil moisture signatures collected over areas with different radii away from the
33

in-situ station, to determine whether and how far the different measurements are correlated. This

is valuable in the selection of station location for long-term maintenance so that the surface data

are valid over the largest area as possible (not just in a close proximity of the station) in view of

the county-scale goal of NIDIS.

Because of the differences in the different data types in time and in space, the role of data

assimilation system [Mitchell et al., 2004; Kumar et al., 2008] is important to integrate various

ground measurements and satellite observations using multiple community land surface models.

The modeling approach allows the various time scales and spatial coverage to be incorporated in

a systematic manner. Since in-situ networks are changing and improving while new satellite

data and products are being developed, land data assimilation needs to be continuously evolved

to account for changes in land surface data from all data sources and to provide enhanced

products that can be used to advance drought monitoring. Furthermore, new measurements can

allow better cross-verifications and validations among different models used in the land data

assimilation systems in an effort to produce the most accurate and high quality products.

Drought is a common climatic phenomenon throughout the world and is not stopped or

limited at geo-political boundaries between different nations, and therefore it is a global problem

requiring international efforts for drought assessment, forecast, and mitigation. In this regard,

satellite data from different nations can contribute to the overall goal. The QSCAT antenna

ceased to spin in November 2009 after its continuous operation collecting global data over a

decade since July 1999. Meanwhile, the Indian Space Agency successfully launched another

scatterometer similar to QSCAT aboard the Oceansat-2 Satellite [Jayaraman et al., 1999] in

September 2009. These satellite events, together with scatterometer data agreement signed

among the different nations, highlight the importance of international collaborations in the use of
34

satellite data to extend the observational record for various applications including drought

monitoring.

Currently, QSCAT is still measuring good backscatter data along narrow tracks at a fixed

azimuth, which are valuable to assist in the calibration and validation of OSCAT. Once

consistently calibrated with QSCAT, OSCAT can continue the QSCAT time-series in producing

SMC results, which are important for drought monitoring based on both present and long-term

data for a consistent assessment of drought conditions within its climatic perspective. In the near

future, China will launch the Haiyan-2 (HY-2) satellite carrying another scatterometer [Dong et

al., 2005]. Moreover, the development of another advanced satellite scatterometer is being

studied in U.S., stemming from the recommendation of the Decadal Survey [National Research

Council, 2007]. Regarding soil moisture measurements, the current European Soil Moisture and

Ocean Salinity (SMOS) mission [Kerr et al., 2010] and the proposed U.S. Soil Moisture Active

and Passive (SMAP) mission [Entekhabi et al., 2010] to be launched in this decade could

provide global measurements critical for drought monitoring. Collectively, these successive

satellite missions would be important to provide multi-decadal data to address the non-stationary

issue in climate change. Moreover, long-term data are necessary for the probabilistic

standardized index approach with multiple time scales of soil moisture variability to be used for

drought monitoring.

Regarding drought monitoring systems, experiences in using satellite data to enhance USDM

and NIDIS are valuable for the development and improvement of international drought

monitoring systems, such as the North American Drought Monitor (NADM) that is a cooperative

effort between drought experts in Canada, Mexico and the United States to monitor drought

across the continent on an ongoing basis [NCDC, 2010b]. In developing countries, the lack of
35

in-situ or surface measurement networks further emphasizes the value of satellite data for

drought monitoring because satellite products such as AMSR-E soil moisture or QSCAT SMC

can be retrieved from global satellite data across international boundaries. For example, Figure

13 presents SMC patterns over the African continent [Nghiem, 2009], from which it is possible to

produce multiple SMC derivatives similar to those developed from USDM (Figure 9). Such

products can contribute to the global drought monitoring as a common goal, for which the Global

Earth Observation System of Systems (GEOSS) [Lautenbacher, 2006] will be crucial as an

overall integrator.

Regarding drought forecasting, another potential contribution of satellite moisture conditions

to forecasting is the use of climatology. Given the limited skill of seasonal forecasts of

temperature and precipitation, drought forecasters place considerable weight on projecting

current conditions forward based on what has happened in the past. In this regard, careful

attention should be paid to the issue of non-stationary due to significant changes in regional

climatic trends in recent years. Once the SMC products are obtained for a suitable number of

years to capture contemporary changes, forecasters may gain knowledge of the probabilities that

soil moisture conditions will likely improve or deteriorate during the following season. One of

the goals of drought forecasting is to cast the forecasts in terms of probabilities so as to provide a

more accurate portrayal of confidence levels, and the statistics of historical soil moisture

conditions can contribute to this effort. In short, improved knowledge of initial moisture

conditions, short-term trends, and climatology has the potential to further the skill of current and

future drought forecasts for not only the United States, but globally as well.
36

Acknowledgments

The research carried out at the Jet Propulsion Laboratory, California Institute of Technology,

was supported by the National Aeronautics and Space Administration (NASA) Water Resources

Area of the NASA Applied Sciences Program. The contribution of the NIDIS team in the SMC

product demonstration in the NIDIS portal is acknowledged. Thanks to Robert Rabin from the

NOAA National Severe Storms Laboratory for his direction to Stage-4 daily precipitation

products for comparison with SMC patterns.

References

Blanchard, B. J., and A. T. C. Chang, Estimation of soil-moisture from SEASAT-SAR data,

Water Resources Bull., 19(5), 803-810, 1983.

Cognard, A. L., C. Loumagne, M. Normand, P. Olivier, C. Ottle, D. Vidalmadjar, S. Louahala,

and A. Vidal, Evaluation of the ERS-1 Synthetic Aperture Radar capacity to estimate surface

soil-moisture 2-year results over the Nazin watershed, Water Resources Res., 31(4), 975-

982, 1995.

Dobson, M. C., F. T. Ulaby, M. T. Hallikainen, and M. A. Elrayes, Microwave dielectric

behavior of wet soil, Part II: Dielectric mixing models, IEEE Trans. Geosci. Remote Sens.,

GE-23(1), 35-46, 1985.

Dong, X., K. Xu, H. Liu, and J. Jiang, The radar altimeter and scatterometer of Chinas HY-2

satellite, Proc. Geosci. Remote Sens. Symp., IGARSS, vol. 3, 1703-1706,

doi:10.1109/IGARSS.2004.1370659, 2004.
37

Dubois, P., J. J. van Zyl and T. Engman, Measuring soil moisture with imaging radars, IEEE

Trans. Geosci. Remote Sens., 33(4), 916-926, 1995.

Engman, E. T., Recent advances in remote-sensing in hydrology, Rev. Geophys., 33, Part 2,

Suppl. S, 967-975, 1995.

Entekhabi, D., E. G. Njoku, P. E. O'Neill, K. H. Kellogg, W. T. Crow, W. N. Edelstein, J. K.

Entin, S. D. Goodman, T. J. Jackson, J. Johnson, J. Kimball, J. R. Piepmeier, R. D. Koster,

N. Martin, K. C. McDonald, M. Moghaddam, S. Moran, R. Reichle, J. C. Shi, M. W.

Spencer, S. W. Thurman, L. Tsang, Leung, and J. J. Van Zyl, The Soil Moisture Active

Passive (SMAP) mission, Proc. IEEE, 98(5), 704-716, 2010.

Gonzlez, J., and J. B. Valds, The mean frequency of recurrence of in-time-multidimensional

events for drought analyses, Natural Hazards and Earth Sys. Sci., 4(1), 17-28, 2004.

Hallikainen, M. T., F. T. Ulaby, M. C. Dobson, M. A. El-Rayes, and L.-K. Wu, Microwave

dielectric behavior of wet soil, Part I: Empirical models and experimental observations, IEEE

Trans. Geosci. Remote Sens., GE-23(1), 25-34, 1985.

Higgins, R. W., W. Shi, E. Yarosh, and R. Joyce, Improved United States precipitation quality

control system and analysis, NCEP/Climate Prediction Center ATLAS No. 7, U.S.

Department of Commerce, National Weather Service, NOAA, available at

http://www.cpc.noaa.gov/products/outreach/research_papers/ncep_cpc_atlas/7/, 2000.

Hutchinson, J. M. S., Estimating near-surface soil moisture using active microwave satellite

imagery and optical sensor inputs, Trans. ASAE, 46(2), 225-236, 2003.

Illston, B. G., J. B. Basara, and K. C. Crawford, Seasonal to interannual variations of soil

moisture measured in Oklahoma, Int. J. Climatol., 24, 1883-1896, 2004.


38

Jackson, T. J., Laboratory evaluation of a field-portable dielectric soil-moisture probe, IEEE

Trans. Geosci. Remote Sens., 28(2), 241-245, 1990.

Jackson, T. J., Soil moisture estimation using SSM/I satellite data over a grassland region, Water

Resources Res., 33, 1475-1484, 1997.

Jayaraman, V., V. S. Hedge, M. Rao, and H. H. Gowda, Future Earth observation missions for

oceanographic applications: Indian perspectives, Acta Astronautica, 44(7-12), 667-674,

1999.

Justice, C. O, J. R. G. Townshend, B. N. Holben, and C. J. Tucker, Analysis of the phenology of

global vegetation using meteorological satellite data, Int. J. Remote Sens., 6(8), 1271-1318,

1985.

Kerr, Y. H. and E. G. Njoku, A semiempirical model for interpreting microwave emission from

semiarid land surfaces as seen from space, IEEE Trans. Geosci. Rem. Sens., 28, 384-393,

1990.

Kerr, Y. H., P. Waldteufel, J.-P. Wigneron, J.-M. Martinuzzi, J. Font, and M. Berger, Soil

moisture retrieval from space: The Soil Moisture and Ocean Salinity (SMOS) mission, IEEE

Trans. Geosci. Remote Sens., 39(8), 1729-1735, 2001.

Kerr, Y. H., P. Waldteufel, J. P. Wigneron, S. Delwart, F. Cabot, J. Boutin, M. J. Escorihuela, J.

Font, N. Reul, C. Gruhier, S. E. Juglea, M. R. Drinkwater, A. Hahne, M. Martin-Neira, and

S. Mecklenburg, The SMOS mission: New tool for monitoring key elements of the global

water cycle, Proc. IEEE, 98(5), 666-687, 2010.

Kumar, S. V., R. H. Reichle, C. D. Peters-Lidard, R. D. Koster, X. W., Zhan, W. T. Crow, J. B.

Eylander, and P. R. Houser, A land surface data assimilation framework using the land
39

information system: Description and applications, Advances in Water Resources, 31(11),

1419-1432, 2008.

Kummerow, C., W. S. Olson, and L. Giglio, A simplified scheme for obtaining precipitation and

vertical hydrometeor profiles from passive microwave sensors, IEEE Trans. Geosci. Remote

Sens., 34, 12131232, 1996.

Lautenbacher, C. C., The Global Earth Observation System of Systems: Science serving society,

Space Policy, 22(1), 8-11, 2006.

Li, Q., L. Tsang, J. Shi, and C. H. Chan, Application of physics-based two-grid method and

sparse matrix canonical grid method for numerical simulations of emissivities of soils with

rough surfaces at microwave frequencies, IEEE Trans. Geosci. Rem. Sens., 38, 1635-1643,

2000.

Loew, A., R. Ludwig, and W. Mauser, Derivation of surface soil moisture from ENVISAT

ASAR wide swath and image mode data in agricultural areas, IEEE Trans. Geosci. Remote

Sens., 44(4), 889-899, 2006.

Mitchell, K. E., D. Lohmann, P. R. Houser, E. F. Wood, J. C. Schaake, A. Robock, B. A.

Cosgrove, J. Sheffield, Q. Y. Duan, L. F. Luo, R. W. Higgins, R. T. Pinker, J. D. Tarpley, D.

P. Lettenmaier, C. H. Marshall, J. K. Entin, M. Pan, W. Shi, V. Koren, J. Meng, B. H.

Ramsay, and A. A. Bailey, The multi-institution North American Land Data Assimilation

System (NLDAS): Utilizing multiple GCIP products and partners in a continental distributed

hydrological modeling system. J. Geophys. Res., 109(D7), D07S90,

doi:10.1029/2003JD003823, 2004.
40

Mo, T., T. J. Schmugge, and T. J. Jackson, Calculations of radar backscattering coefficient of

vegetation covered soils, Rem. Sens. Environ., 15, 119-133, 1984.

Moran, M. S., V. Vidal, D. Troufleau, J. Qi, T. R. Clarke, P. J. Pinter, Jr., T. A. Mitchell, Y.

Inoue, and C. M. U. Neale, , Combining multifrequency microwave and optical data for crop

management, Remote Sens. Environ., 61, 96-109, 1997.

Moran, M. S., A. Vidal, D. Troufleau, Y. Inoue, and T. A. Mitchell, Ku- and C-band SAR for

discriminating agricultural crop and soil conditions, IEEE Trans. Geosci. Remote Sens., 36,

265-272, 1998.

Myneni, R. B., C. D. Keeling, C. J. Tucker, G. Asrar, and R. R. Nemani, Increased plant growth

in the northern high latitude from 1981 to 1991, Nature, 386(6626), 698-702, 1997.

National Research Council, Earth Science and Applications from Space: National Imperatives

for the Next Decade and Beyond. The National Academies Press, Washington, D.C., 2007.

NCDC, State of the Climate Drought, August 2009, National Climatic Data Center, NESDIS,

NOAA, U.S. Department of Commerce, http://www.ncdc.noaa.gov/sotc/?report=drought&

year=2009&month=8, 2009.

NCDC, Global Surface Summary of the Day GSOD, National Climatic Data Center, NESDIS,

NOAA, U.S. Department of Commerce, http://www.data.gov/geodata/g600037/, accessed

2010a.

NCDC, North American Drought Monitor, National Climatic Data Center, NESDIS, NOAA,

U.S. Department of Commerce, http://www.ncdc.noaa.gov/temp-and-

precip/drought/nadm/index.html, accessed 2010b.


41

Nghiem, S. V., M. Borgeaud, J. A. Kong, and R. T. Shin, Polarimetric remote sensing of

geophysical media with layer random medium model, Progress in Electromagnetics

Research - Polarimetric Remote Sensing, ed. by J. A. Kong, Elsevier, New York, Vol. 3,

Chapter 1, pp. 1-73, 1990.

Nghiem, S. V., T. Le Toan, J. A. Kong, H. C. Han, and M. Borgeaud, Layer model with random

spheroidal scatterers for remote sensing of vegetation canopy, J. Electromag. Waves Applic.,

7(1), 49-76, 1993a.

Nghiem, S. V., S. H. Yueh, R. Kwok, and D. T. Nguyen, Polarimetric remote sensing of

geophysical medium structures, Radio Sci., 28(6), 1111-1130, 1993b.

Nghiem, S. V., R. Kwok, S. H. Yueh, and M. R. Drinkwater, Polarimetric signatures of sea ice,

1, Theoretical model, J. Geophys. Res., 100(C7), 13665-13679, 1995.

Nghiem, S. V., J. J. Van Zyl, W.-Y. Tsai, and G. Neumann, Potential application of

scatterometry to large-scale soil moisture monitoring, JPL Doc. D-19523, Jet Propulsion

Lab., California Inst. Tech., Pasadena, California, 2000.

Nghiem, S. V., Advanced scatterometry for geophysical remote sensing, JPL Doc. D-23048, Jet

Propulsion Lab., California Inst. Tech., Pasadena, California, 2001.

Nghiem, S. V., E. G. Njoku, J. J. Van Zyl, and Y. Kim, Global energy and water cycle Soil

moisture variability pattern over continental extent observed with active and passive satellite

data, JPL Doc. D-26225, Jet Propulsion Lab., California Inst. Tech., Pasadena, California,

2003.
42

Nghiem, S. V., E. G. Njoku, G. R. Brakenridge, and Y. Kim, Land surface water cycles observed

with satellite sensors, J6.14, Proc. 19th Conference on Hydrology and 16th Conference on

Climate Variability and Change, 85th Amer. Met. Soc. Meeting, San Diego, California, 2005.

Nghiem, S. V., NASA satellite data for applications to early warning of droughts across the

world Examples for Africa, Inter-Regional Workshop on Indices and Early Warning

Systems for Drought, Lincoln, Nebraska, 8-11 December 2009.

Nghiem, S. V., J. Verdin, M. Svoboda, D. Allured, J. Brown, B. Liebmann, G. Neumann, E.

Engman, and D. Toll, Improved drought monitoring with NASA satellite data, EWRI

Currents, 12(3), 7, Environ. Water Res. Inst., Amer. Soc. Civil Eng., 2010.

NIDIS, U.S. Public Law 109-430, 109th Congress, 2nd sess., National Integrated Drought

Information System Act of 2006, 20 Dec.2006.

NIDIS, The National Integrated Drought Information System Implementation Plan A Pathway

for National Resilience, NIDIS Implementation Team, 28 pp., July 2007.

Njoku, E. G., and L. Li, Retrieval of land surface parameters using passive microwave

measurements at 6-18 GHz, IEEE Trans. Geosci. Rem. Sens., 37, 79-93, 1999

Njoku, E. G., T. J. Jackson, V. Lakshmi, T. K. Chan, and S. V. Nghiem, Soil moisture retrieval

from AMSR-E, IEEE Trans. Geosci. Remote Sens., 41(2), 215-229, 2003.

Njoku, E., AMSR-E/Aqua L2B Surface Soil Moisture, Ancillary Parms, & QC EASE-Grids

V002, http://nsidc.org/data/docs/daac/ae_land_l2b_soil_moisture.gd.html, Boulder, Colorado

USA: National Snow and Ice Data Center, 2004.

NMQ, Stage-4 24hr QPE Accumulation, National Mosaic & Multi-Sensor QPE, NOAA National

Severe Storms Laboratory, http://nmq.ou.edu/, 2009.


43

NWS, September 2009 Weather in Review, NOAA National Weather Service, Southern Region

Headquarters, http://www.srh.noaa.gov/images/ewx/wxevent/sep2009.pdf, 2009.

Owe, M., A. Chang, and R. E. Golus, Estimating surface soil moisture from satellite microwave

measurements and a satellite-derived vegetation index, Rem. Sens. Environ., 24, 131-345,

1988.

Prevot, M., M. Dechambre, O. Taconet, D. Vidal-Madjar, M. Normand, and S. Galle, Estimating

the characteristics of vegetation canopies with airborne radar measurements, Int. J. Rem.

Sens., 14, 2803-2818, 1993.

Resource Management Services, State of Montana multi-hazard mitigation plan and statewide

hazard assessment, Land and Water Consulting, Big Sky Management, 251 pp., 2004.

Shrivastava, S. K., N. Yograjan, V. Jayaraman, P. P. N. Rao, and M. G. Chandrasekhar, On the

relationship between ERS-1 SAR/backscatter and surface/sub-surface soil moisture

variations in vertisols, Acta Astronautica, 40(10), 693-699, 1997.

Shrivastava, H. S., P. Patel, Y. Sharma, and R. R. Navalgund, Large-area soil moisture

estimation using multi-incidence-angle RADARSAT-1 SAR data, IEEE Trans. Geosci.

Remote Sens., 47(8), 2528-2535, 2009.

Shoshany, M., T. Svoray, P. J. Curran, G. M. Foody, and A. Perevolotsky, The relationship

between ERS-2 SAR backscatter and soil moisture: Generalization from a humid to semi-arid

transect, Int. J. Remote Sens., 21(11), 2337-2343, 2000.

Smith, J. A., D. J. Seo, M. L. Baeck, and M. D. Hudlow, An intercomparison study of NEXRAD

precipitation estimates, Water Resources Res., 32(7), 2035-2045, 1996.


44

Sorooshian, S., K. L. Hsu, X. Gao, H. V. Gupta, B. Imam, and D. Braithwaite, Evaluation of

PERSIANN system satellite-based estimates of tropical rainfall, Bull. Amer. Meteorol. Soc.,

81(9), 2036-2046, 2000.

Story, G., The difficulty of achieving good precipitation estimates for use in real-time drought

monitoring, 6th U.S. Drought Monitor Forum, Lower Colorado River Authority, Austin, TX,

7-8 Oct. 2009.

Svoboda, M. D., M. J. Hayes, and D. A. Wilhite, The role of integrated drought monitoring in

drought mitigation planning, Ann. Arid Zone, 40(1), 1-11, 2001.

Svoboda, M., D. LeComte, M. Hayes, R. Heim, K. Gleason, J. Angel, B. Rippey, R. Tinker, M.

Palecki, D. Stooksbury, D. Miskus, and S. Stephens, The Drought Monitor, Bull. Amer.

Meteorol. Soc., 83(8),1181-1190, 2002.

Takada, M., Y. Mishima, and S. Natsume, Estimation of soil surface properties in peatland using

ALSO/PALSAR, Landscape and Ecological Engineering, 5(1), 45-58, 2009.

Texas Water Development Board, Highlights of the 2007 State Water Plan, Water for Texas

2007, 39 pp., vol. I, Doc. No. GP-8-1., 2007.

Teng, W. L., Wang, J. R., and Doriaswamy, P. C., Relationship between satellite microwave

radiometric data, antecedent precipitation index, and regional soil moisture, Int. J. Rem.

Sens., 14: 2483-2500, 1993.

Tsai, W.-Y., S. V. Nghiem, J. N. Huddleston, M. W. Spencer, B. W. Stiles, and R. D. West,

Polarimetric scatterometry: A promising technique for improving ocean surface wind

measurements, IEEE Trans. Geosci. Remote Sens., 38, 1903-1921, 2000.


45

USDA, SCAN - Soil Climate Analysis Network, SCAN Brochure, Natural Resources

Conservation Service, National Water & Climate Center National Soil Survey Center, 2009a.

USDA, SNOTEL And Snow Survey & Water Supply Forecasting, SNOTEL Brochure, Natural

Resources Conservation Service, National Water & Climate Center, NWCC Rev. 3, 2009b.

van de Griend, A. A. and M. Owe, Microwave vegetation optical depth and inverse modelling of

soil emissivity using Nimbus/SMMR satellite observations, Meteorol. Atmos. Phys., 54,

225-239, 1994.

Verdin, J., C. Funk, R. Klaver, and D. Roberts, Exploring the correlation between Southern

Africa NDVI and Pacific sea surface temperatures: results for the 1998 maize growing

season, Int. J. Remote Sens., 20(10), 2117-2124, 1999.

Wagner, W., J. Noll, M. Borgeaud, and H. Rott, Monitoring soil moisture over the Canadian

prairies with the ERS scatterometer, IEEE Trans. Geosci. Remote Sens., 37(1), 206-216,

1999.

Wagner, W., and K. Scipal, Large-scale soil moisture mapping in western Africa using the ERS

scatterometer, IEEE Trans. Geosci. Remote Sens., 38(4), Part 2, 1777-1782, 2000.

Wang, J. R., and B. J. Choudhury, Remote sensing of soil moisture content over bare field at 1.4

GHz frequency, J. Geophys. Res., 86, 5277-5282, 1981.

Wang, J. R., Passive microwave sensing of soil moisture content: the effects of soil bulk density

and surface roughness, Remote Sens. of Environ., 13, 329-344, 1983.

Wang, J. R., Effect of vegetation on soil moisture sensing observed from orbiting microwave

radiometers, Rem. Sens. Environ., 17, 141-151, 1985.


46

Western Governors Association, Creating a Drought Early Warning System for the 21st

Century, The National Integrated Drought Information System, 13 pp., Denver, Colorado,

2004.

Wigneron, J.-P., L. Laguerre, and Y. H. Kerr, A simple parameterization of the L-band

microwave emission from rough agricultural soils, IEEE Trans. Geosci. Rem. Sens., 39,

1697-1707, 2001.

World Meteorological Organization, Expert agree on a universal drought index to cope with

climate risks, Press Release, WMO No.-872, United Nations, Copenhagen, Geneva, 2009.

Zhang, X. Y., M. Goldberg, D. Tarpley, M. A. Friedl, J. Morisette, F. Kogan, and Y. Y. Yu,

Drought-induced vegetation stress in southwestern North America, Environ. Res. Lett., 5(2),

Art. 024008, 2010.


47

List of Table

Table 1. Correlation results between seasonal TMI polarization ratio PR and seasonal SCAN

volumetric soil moisture mv at 5-cm depth from linear regression analysis in the form of PR =

mv + with a correlation coefficient .

List of Figures

Figure 1. Seasonal TMI polarization ratio versus seasonal SCAN soil moisture at 5-cm depth in

an agricultural area at Lonoke, Arkansas. All data are contemporaneous (collocated in time), and

are 90-day running averages. The upper plot is for 10.7 GHz, and the lower one is for 19.3 GHz.

Figure 2. Seasonal QSCAT backscatter 0VV versus seasonal TMI polarization ratio PR at 10.7

GHz in an agricultural area within 25 km around Lonoke, Arkansas. All data are

contemporaneous (collocated in time), and are 90-day running averages.

Figure 3. Measurements around the NCDC GSOD Station 727760 (47.467oN 111.383oW) at

Great Falls in Montana. Top panel is QSCAT 0HH within 25 km around the station, middle

panel is in-situ air temperature (magenta for minimum, black for average, and red for maximum),

and bottom panel is precipitation from station rain gauge (there were missing data). Thin vertical

lines align rain events to backscatter impulses. Yellow marks the period between 7/9/2000 and

9/5/2000 when there was very little rain.


48

Figure 4. Effective precipitation frequency (%) measured by QSCAT the period of June-August

2009 (left panel), and drought levels from D0 to D4 from the U.S. Drought Monitor for weeks

ending on the marked dates in 2009 (right panels). The USDM drought levels include D0 for

abnormally dry, D1 for moderate drought, D2 for severe drought, D3 for extreme drought, and

D4 for exceptional drought [Svoboda et al., 2001, 2002].

Figure 5. Soil moisture change (SMC) measured by QSCAT with the vertical polarization along

ascending orbits in September to early October 2009. The color scale represents backscatter

change in dB, and volumetric SMC in % with the Lonoke rating.

Figure 6. Difference of AMSR-E monthly averaged soil moisture in % of mv(9/7/2009-

10/6/2009) mv(6/29/2009-7/28/2009) showing seasonal SMC (left panel), and drought condition

change between USDM drought maps in July and in September 2009 (right panels).

Figure 7. Soil moisture change (SMC) on 28 May 2009 compared to the two-week average

between 14-28 May 2009 observed by: (a) QSCAT SMC represented by backscatter change in

dB and by volumetric moisture change in % from the Lonoke rating, and (b) AMSR-E by

volumetric moisture change in % with yellow-brown for drier and cyan-blue for wetter

conditions.

Figure 8. Stage-4 24-hour precipitation measurements [NMQ, 2009] at 12:00 UTC in inches for:

(a) 28 May 2009 in the upper panel, and (b) and 27 May 2009 in the lower panel.
49

Figure 9. Comparison of QSCAT SMC with rain-gauge precipitation (RGP) for the period of 14

October 2008 and the ensuing 7 days: (a) mean SMC, (b) max SMC, and (c) RGP used in making

USDM maps.

Figure 10. Weekly QSCAT mean SMC maps for the growing season in June-October 2009.

Figure 11. Weekly USDM maps for the growing season in June-October 2009.

Figure 12. SMC product demonstration in the NIDIS portal at the national level with USDM

and AHP River Gauge layers (upper panel), and at the county level for Nebraska (lower panel).

Figure 13. Soil moisture change (SMC) across Africa [Nghiem, 2009] measured by QSCAT

with the vertical polarization along ascending orbits on 16 October 2009. The color scale

represents backscatter change in dB, and volumetric SMC in % with the Lonoke rating. Data

gaps between orbits are seen in dark bands on land. High SMC is observed in an extensive

pattern (yellow areas) across South Africa curving northward to Botswana, Namibia, Zambia,

and Mozambique, with a wet region seen in Madagascar.


50

Table 1. Correlation results between seasonal TMI polarization ratio PR and seasonal SCAN

volumetric soil moisture mv at 5-cm depth from linear regression analysis in the form of PR =

mv + with a correlation coefficient .

10.7 GHz 19.3 GHz



Fall-winter 0.00109 0.00766 0.977 0.000931 0.00587 0.953
Spring-summer 0.00131 -0.00108 0.988 0.000960 -0.000908 0.946
All year 0.00124 0.00235 0.936 0.000894 0.00320 0.792
51

(a)

(b)
Figure 1. Seasonal TMI polarization ratio versus seasonal SCAN soil moisture at 5-cm depth in
an agricultural area at Lonoke, Arkansas. All data are contemporaneous (collocated in time), and
are 90-day running averages. The upper plot is for 10.7 GHz, and the lower one is for 19.3 GHz.
52

Figure 2. Seasonal QSCAT backscatter 0VV versus seasonal TMI polarization ratio PR at 10.7
GHz in an agricultural area within 25 km around Lonoke, Arkansas. All data are
contemporaneous (collocated in time), and are 90-day running averages.
53

Figure 3. Measurements around the NCDC GSOD Station 727760 (47.467oN 111.383oW) at
Great Falls in Montana. Top panel is QSCAT 0HH within 25 km around the station, middle
panel is in-situ air temperature (magenta for minimum, black for average, and red for maximum),
and bottom panel is precipitation from station rain gauge (there were missing data). Thin vertical
lines align rain events to backscatter impulses. Yellow marks the period between 7/9/2000 and
9/5/2000 when there was very little rain.
54

Figure 4. Effective precipitation frequency (%) measured by QSCAT the period of June-August
2009 (left panel), and drought levels from D0 to D4 from the U.S. Drought Monitor for weeks
ending on the marked dates in 2009 (right panels). The USDM drought levels include D0 for
abnormally dry, D1 for moderate drought, D2 for severe drought, D3 for extreme drought, and
D4 for exceptional drought [Svoboda et al., 2001, 2002].
55

Figure 5. Soil moisture change (SMC) measured by QSCAT with the vertical polarization along
ascending orbits in September to early October 2009. The color scale represents backscatter
change in dB, and volumetric SMC in % with the Lonoke rating.
56

Figure 6. Difference of AMSR-E monthly averaged soil moisture in % of mv(9/7/2009-


10/6/2009) mv(6/29/2009-7/28/2009) showing seasonal SMC (left panel), and drought condition
change between USDM drought maps in July and in September 2009 (right panels).
57

Figure 7. Soil moisture change (SMC) on 28 May 2009 compared to the two-week average
between 14-28 May 2009 observed by: (a) QSCAT SMC represented by backscatter change in
dB and by volumetric moisture change in % from the Lonoke rating, and (b) AMSR-E by
volumetric moisture change in % with yellow-brown for drier and cyan-blue for wetter
conditions.
58

Figure 8. Stage-4 24-hour precipitation measurements [NMQ, 2009] at 12:00 UTC in inches for:
(a) 28 May 2009 in the upper panel, and (b) and 27 May 2009 in the lower panel.
59

Figure 9. Comparison of QSCAT SMC with rain-gauge precipitation for the period of 14
October 2008 and the ensuing 7 days: (a) mean SMC, (b) max SMC, and (c) precipitation used in
making USDM maps.
60

Figure 10. Weekly QSCAT mean SMC maps for the growing season in June-October 2009.
61

Figure 11. Weekly USDM maps for the growing season in June-October 2009.
62

Figure 12. SMC product demonstration in the NIDIS portal at the national level with USDM
and AHP River Gauge layers (upper panel), and at the county level for Nebraska (lower panel).
63

Figure 13. Soil moisture change (SMC) across Africa [Nghiem, 2009] measured by QSCAT
with the vertical polarization along ascending orbits on 16 October 2009. The color scale
represents backscatter change in dB, and volumetric SMC in % with the Lonoke rating. Data
gaps between orbits are seen in dark bands on land. High SMC is observed in an extensive
pattern (yellow areas) across South Africa curving northward to Botswana, Namibia, Zambia,
and Mozambique, with a wet region seen in Madagascar.
64

List of Publications (Section 1):

Journal Articles and Book Chapters

Nghiem, S. V., D. B. Wardlow, D. Allured, M. D. Svoboda, D. LeComte, M. Rosencrans, K. S.

Chan, and G. Neumann, Microwave Remote Sensing of Soil Moisture Science and

Applications, book chapter, 53 pp., in Drought and Water Crises Book Series - Remote

Sensing and Drought New and Emerging Monitoring Approaches, Taylor and Francis Pub.,

in review, 2011.

Nghiem, et al. (authors from JPL, USGS, NDMC, NOAA PSD, DFO, and others), Pattern and

Frequency of Soil Moisture Variability over the Continental United States, manuscript in

revision, 52 pp., 2011.

Nghiem, S. V., J. Verdin, M. Svoboda, D. Allured, J. Brown, B. Liebmann, G. Neumann, E.

Engman, and D. Toll, Improved Drought Monitoring with NASA Satellite Data, EWRI

Currents, 12(3), 7, Amer. Soc. Civil Eng., Summer 2010.

Nghiem, S. V., D. Balk, E. Rodriguez, G. Neumann, A. Sorichetta, C. Small, and C. D. Elvidge,

Observations of Urban and Suburban Environments with Global Satellite Scatterometer

Data, ISPRS Journal of Photogrammetry and Remote Sensing, 64, 367-380,

doi:10.1016/j.isprsjprs.2009.01.004, 2009.

Nghiem, S. V., and G. Neumann, Remote Sensing of the Global Environment with Satellite

Scatterometry, keynote paper in Microwave Remote Sensing of the Atmosphere and


65

Environment VI, ed A. Valinia, P. H. Hildebrand, and S. Uratsuka, Proc. of SPIE, Vol. 7154,

715402, doi:10.1117/12.804462, 11 pages, 2008.

Press Release

JPL Photo Journal, Rapid Dry-Up of Rainwater on Land Surface Leading to the Santa Barbara

Wildfire, Internet article http://photojournal.jpl.nasa.gov/catalog/PIA12006, Jet Propulsion

Laboratory, Pasadena, California, 8 May 2009.

Conference and Meeting Presentations

Nghiem, S. V., NASA Satellite Data for Applications to Early Warning of Droughts across the

World Examples for Africa, Inter-Regional Workshop on Indices and Early Warning

Systems for Drought, Lincoln, Nebraska, 8-11 December 2010.

Nghiem, S. V., Satellite Observation of Soil Moisture Change and Applications to Drought

Monitoring, 6th U.S. Drought Monitor Forum, Lower Colorado River Authority, Redbud

Center, Austin, Texas, 7-8 October 2009.

Nghiem, S. V., Geophysical Information from NASA Satellite Scatterometry Western U.S.

and California, ESRI International User Conference, San Diego Convention Center, San

Diego, California, 13-17 July 2009.

Nghiem, S. V., Satellite Remote Sensing of Soil Moisture for Drought Applications, invited

paper, National Integrated Drought Information System Knowledge Assessment Workshop


66

Contribution of Satellite Remote Sensing to Drought Monitoring, Boulder, Colorado, USA,

6-7 February 2008.

Nghiem, S. V., G. R. Brakenridge, and G. Neumann, Drought, wetland, and Flood Monitoring

with Satellite Scatterometer, EOS Trans, AGU, 88(23), Jt. Assem. Suppl., Abst. U53B-05,

May 2007.

Nghiem, S. V., G. R. Brakenridge, D. Cline, M. Dettinger, R. M. Dole, P. R. Houser, G.

Neumann, E. G. Njoku, D. K. Perovich, K. Steffen, M. Sturm, J. Verdin, D. A. Wilhite, S. H.

Yueh, and T. Zhang, Global Observations of Land Surface Water with Satellite Active and

Passive Microwave Sensors, Satellite Observations of the Global Water Cycle, Irvine,

California, 7-9 March 2007.

Nghiem, S. V., Drought Monitoring with NASA Satellite Data, National Integrated Drought

Information System (NIDIS) Implementation Plan, Longmont, Colorado, Sept., 2006.


67

Section 2 (Led by U.S. Geological Survey)

Vegetation Drought Response Index

J. Brown, Y. Gu, and J. Verdin

U.S. Geological Survey, EROS Data Center, Sioux Falls, SD 57198, (605) 594-6018

Summary

New soil moisture and vegetation monitoring products were developed for integration into the

weekly production of the U.S. Drought Monitor (USDM) map (http://drought.unl.edu/dm), the

recognized national reference for drought conditions in the United States. A soil moisture

change (SMC) product was developed using NASA scatterometer and radiometer data (led by

JPL, see Section 1). The Vegetation Drought Response Index (VegDRI), originally formulated

to use imagery from NOAA AVHRR instruments, was updated to use imagery from NASA

MODIS instead, allowing us to integrate superior radiometric and geometric characteristics.

Production of VegDRI with MODIS imagery was implemented on a very fast turn-around (less

than 12 hours after satellite acquisition) by taking advantage of the new NASA LANCE system

for rapid delivery of swath surface reflectance data, and the implementation of a new expedited

MODIS (eMODIS) processing chain at USGS EROS. Frequency and timing of production were

designed to meet USDM author weekly schedules, and automated procedures were implemented

to ingest SMC and VegDRI products into the GIS environment used by authors to make weekly

adjustments to drought category zones. Based on positive feedback from USDM authors, USGS
68

EROS is now investing in the transition of VegDRI production from science to operations.

Systems engineering staff are implementing a robust, automated operational VegDRI production

chain to support the U.S. drought monitoring community on into the future, beyond the life of

this project. The operational VegDRI system is expected to be actively supporting USDM

authors throughout the 2011 growing season.

VegDRI Project Background and History

Initial development of the Vegetation Drought Response Index (VegDRI) began in 2002, funded

by a seed grant from the U.S. Geological Survey (USGS). The grant funded prototype efforts, in

collaboration with the National Drought Mitigation Center (NDMC), to develop methods to

improve drought monitoring in the north central plains of the U.S. The initial study (outlined by

Brown et al. 2008) presented an approach to synoptic monitoring of drought impacts on

vegetation based on phenological indicators using data from a series of optical sensorsthe

Advanced Very High Resolution Radiometer (AVHRR). This seed money also supported

writing several research proposals. Subsequent funding awarded in 2006 by the NASA Applied

Sciences program (NASA Decisions/05-2-0000-0167) allowed for further model enhancements

and geographic expansion of the VegDRI modeling approach to the conterminous U.S.

(CONUS). The project goal was to improve drought early warning response and mitigation

through the use of timely monitoring products designed to meet the needs of the U.S. community

of drought-sensitive decision-makers.

The VegDRI methodology builds upon the monitoring traditions of both the climate and remote

sensing communities. The approach provides improvement in spatial detail, spatial coverage,
69

and the timely delivery of information in a variety of accessible formats (for example, maps,

descriptive text, and statistics), increasing the value, spatial detail, and relevancy of drought

information available to decision makers. A key project goal was to integrate VegDRI into the

weekly operational process of making the U.S. Drought Monitor (USDM) map by providing

data sets to the USDM authors.

Since its launch in 1999, the USDM (Svoboda et al. 2002) has been the state-of-the-practice

drought monitoring tool used in the United States. The USDM provides a general assessment of

weekly drought conditions (both agricultural and hydrologic) across the nation. It is produced

using a hybrid approach that involves a number of variables including short- and long-term

climate-based drought indicators, hydrologic indices, and remote sensing information. Ten years

ago, the USDM map was produced at the effective scale of climate divisions (multi-county

aggregations with similar climate), however developments in recent years, including finer

resolution point source and gridded climatic data, have led to improvement in resolution

approaching individual county scale or better when we incorporate remote sensing products

(Svoboda, pers. communication).

Climate and meteorological data have long been the primary sources for creating drought indices

for monitoring, including the USDM. Climate-based drought indices characterize the intensity

of dryness as compared to long-term average or normal condition, and are usually calculated

from one or more of the following variables: rainfall, temperature, snow pack, stream flow, soil

moisture, and other water supply indicators. The spatial coverage and detail of climate-based

drought indices is limited as these depend upon meteorological data collected at stations that are

non-uniformly distributed across the country. As a result, climate-based monitoring tools

characterize relatively broad-scale drought patterns and the level of accuracy and spatial detail in
70

the information they provide depends on the density and geographic placement of stations across

the landscape.

Time-series Vegetation Index (VI) Applied to Operational Monitoring

Satellite-based observations have proven very useful for assessing broad-scale vegetation

condition anomalies (that is, apparent changes in vegetation health), but the specific cause or

causes for the anomalies may not always be determined solely from the remotely sensed data. A

number of natural (for example, drought, flooding, fire, pest infestation, and hail damage) and

anthropogenic (for example, land cover/land use conversion) events can produce these anomalies

(Asner et al. 2000; Breshears 2005; Goetz et al. 2006; Kogan 1990; Parker et al. 2005; Peters et

al. 2000; Peters et al. 2002; Wang et al. 2003). Therefore, effective drought monitoring

approaches must consider both climate station and satellite-based information, as well as other

environmental parameters that may influence the effects of drought and its severity on

vegetation, such as soil properties or land cover type. It follows that the integration of coarser-

resolution climate data and higher-resolution satellite-based vegetation observations will provide

an improvement to monitoring and characterization of the spatial extent, intensity, and local

variability of droughts effect on vegetation conditions.

Product Requirements for Drought Decision Support

Because the effects of drought vary from region to region and season to season, there are many

approaches for early warning and monitoring that range from sub-national, to national, regional,
71

and even global in geographic scope. To be successful, monitoring approaches must deal with a

multitude of challenging drought characteristics (e.g., the difficulty in determining drought onset

and termination, the multiple, varying definitions of drought, and the existence of multiple

indicators --climatic, phenologically-related, VI, hydrologic, etc.). The drought community in

the United States, largely through the coordination and leadership of the U.S. Drought Monitor

(USDM) authors (Svoboda et al. 2002) and the National Integrated Drought Information System

(WGA 2004), has suggested product/data requirements for successful drought monitoring shown

in Table 1.

Table 1. Operational National Level Drought Monitoring Product Requirements

Product geographic coverage National synoptic (minimum 48 state coverage)

Product schedule Weekly, available Monday a.m.

Spatial scale 1-4 km2, sub-county details

Product latency <24-48 hours

Length of record ~30 years (for climate data)

Data and information should be framed using

baselines and anomalies.

The VegDRI system was designed to meet many of these requirements. The 1-km2 resolution

VegDRI maps (Figure 1) depict drought-specific information related to vegetation condition with

higher spatial detail than traditional drought monitoring tools such as the Palmer Drought

Severity Index (PDSI) and USDM. Thanks to its spatial detail and coverage, VegDRI can
72

support drought assessment at local to national levels. Because of its weekly production and

release before noon (Eastern Time) on each Monday, VegDRI supports the weekly production

schedule followed by the authors in assembling the national USDM map.

Figure 1. VegDRI map for July 12, 2009

VegDRI Overview

VegDRI characterizes the level of drought stress on vegetation by integrating two satellite-based

time-series vegetation index-based metrics, two climate-based drought indices, and five
73

biophysical characteristics of the environment. Further details on the basic input variables and

the methodology are found in (Brown et al. 2008). The VegDRI methodology consists of three

main steps:

1. Processing, summarization, and organization of the data for each input variable into a

database

2. Development of empirically derived VegDRI models by applying a classification and

regression tree (CART) analysis technique to the historical information in the database

3. Application of the seasonal models in operational processing to produce weekly 1-km2

resolution VegDRI maps

VegDRI maps portray drought severity in seven categories reflecting varying levels of drought-

induced vegetation stress that are based on the PDSI classification scheme (Wells et al. 2004;

Palmer, 1965). The maps also include three additional classes that depict areas over which

VegDRI values cannot be calculated. These additional classes include water, out of season (i.e.,

time when the vegetation is not photosynthetically active for a location), and no season (i.e.,

locations in the southwestern U.S. where there was no detectable vegetation response in the

historical VI data).

Several significant improvements to the VegDRI system were made in carrying out this NASA

project. They include:

(1) use of expedited MODIS (eMODIS) imagery instead of AVHRR to calculate the satellite

phenological indicators,

(2) an automated production flow for the near-real time geospatial model,

(3) new data inputs,

(4) more timely product delivery,


74

(5) new data product delivery (formats and websites).

Time-series Satellite VI data: eMODIS and AVHRR

The satellite VI data ingested into VegDRI are collected from two daily polar-orbiting earth

observing systems, AVHRR and MODIS. These instruments provide frequent, rapid, and

synoptic measurements of land surface conditions across large geographic regions. A more than

20-year history of AVHRR time-series data across the United States provides historical context

for monitoring drought conditions (Eidenshink, 2006). However, the newer MODIS instrument

has improved sensor characteristics specifically designed for global land surface monitoring.

The MODIS sensors carried aboard the NASA Terra and Aqua satellite platforms have 36

spectral channels and have been extensively used for land and atmospheric monitoring since

2000 (http://modis.gsfc.nasa.gov/). MODIS land products [e.g., surface reflectance and

Normalized Difference Vegetation Index (NDVI)] are similar to AVHRR products but provide

higher spatial and spectral resolutions and improved atmospheric and geometric corrections

(Townshend and Justice 2002). AVHRR data have a longer historical record, allowing more

definitive establishment of climatological or normal conditions and the identification of

anomalous conditions. Therefore, both AVHRR and MODIS NDVI data are valuable inputs for

operational monitoring.

Since standard MODIS products are often not delivered to users quickly enough to aid

operational decisions (typically 6-10 day lag times between satellite overpass and standard

product data availability have been noted in the past), USGS has designed and is now operating

an expedited MODIS, or eMODIS system that takes advantage of NASAs new Land
75

Atmosphere Near-real time Capability for EOS (LANCE). The eMODIS system at EROS

(Figure 2) provides the near-real time MODIS VI and surface reflectance data (typically less than

10 hours after satellite overpass) needed to supply VegDRI products on a schedule that meets

United States drought community requirements (Jenkerson et al., 2010).

eMODIS Historical

Data Delivery Target:


Monday 10:30 a.m.

Terra MODIS
+ 2-4hrs
LAADS USGS Drought
VegDRI
Monitoring

MODAPS
EDOS
eMODIS Expedited
+ 4hrs NIDIS Drought
MODIS L0 Data Portal

+ 3hrs

NDMC Vegetation U.S. Drought


MODIS eMODIS Expedited
Drought Monitor
LANCE eMODIS Historical
L2, L1B Data

Figure 2. Automated eMODIS and VegDRI Production Flow Diagram

Each expedited NDVI composite product from eMODIS is derived from data delivered in near

real time by LANCE. eMODIS ingests level-2 surface reflectance swaths from LANCE as soon

as they are available (3-6 hours after acquisition). The swaths are gridded per VegDRI

specifications, and NDVI is calculated. The gridded NDVI pixels are evaluated by an enhanced

maximum value compositing routine to select the best available data for each 7-day composite.
76

Expedited eMODIS NDVI is produced daily using the most recent 7 days of data. The expedited

products are later replaced in the permanent archive by precision eMODIS data generated from

standard reflectance inputs within 3-10 days as they are delivered from the Level 1 and

Atmosphere Archive and Distribution System (LAADS).

Automated VegDRI Production System

System Design Features

The VegDRI System consists of a set of software components, primarily based on C code, that

can be initiated independently and manually. For near-real time processing, the components are

run in a processing stream that is automated using scripting. Appendix 1 shows the components

of the system within the context of the typical processing stream.

Operational Status

VegDRI production is being transitioned by USGS to operational status. In the current status,

VegDRI components are run in a processing stream that is initiated at 00:01 a.m. each Monday.

Automated processing ordinarily completes by 10:30 a.m. each Monday morning, 24 hours after

Terra MODIS satellite data acquisition. When a problem is encountered with one of the required

inputs or one of the processing components, completion of processing is delayed, but can be

analyzed, completed, and/or re-run within a day. The initiation of each component for each

interim product is automated from scripts but can be done manually. Any component testing or

verification of results is done by staff, nominally during normal business hours (8:00 a.m. 4:00

p.m.).
77

Maintenance

Current VegDRI production streams [7-day (Monday a.m.) CONUS] are maintained without

interruption and fully supported by the necessary hardware and software components.

Automation and Recovery Methods

Software development completed to date is included in the VegDRI prototype operations and

maintenance environment. Processing results are logged. Successful results or problems in

production in the form of logs are emailed to staff. Software engineering staff remains

responsible for production tasks as well as for troubleshooting and catastrophic recovery actions.

Re-processing of individual components or the entire string of components of the VegDRI

system is currently a manual process because components were not developed to support a

recovery methodology. Each component was developed independently, which allows flexibility

of use, but they were not optimized with operational considerations. They are joined together

with automation scripts to create an operational work flow. Failures can occur at various steps in

processing, often without warning or notification when external inputs are unavailable. Previous

failures have involved hardware and network issues, problems with collaborator inputs, and

problems with external interfaces. The VegDRI system relies on software engineering staff to

recognize and mitigate these issues, which requires a growing knowledge base to facilitate

recognition of expected anomalies versus actionable system failures. As each individual issue

has been accommodated, processing has become more stable, but some issues are due to external

dependencies.
78

Model Recalibration

Annual recalibration of VegDRI models is desired to incorporate new time series data and new

regression tree models incorporating data inputs (eMODIS NDVI, gridded climate surfaces, and

biophysical parameters).

Extensibility

Based on increasing interest in VegDRI capabilities, we expect future improvements in models

and input data. VegDRI models and data can be extended to additional geographic windows.

Geographical expansion in North America is currently being explored in collaboration with

Agriculture and Agri-Food Canada. Finally, satellite continuity is critical and in the post-

MODIS era, extension to the Visible Infrared Imager Radiometer Suite (VIIRS) will be essential

to remain operational into the future.

System Configuration

The VegDRI system is installed on a Dell Server (purchased in 2010) running the linux operating

system with the following components: a dual-processor CPU with 32GB of memory, 2.5TB of

RAID disk storage, redundant power supplies. The server is covered by a five year warranty.

Data Management and Archive

The eMODIS expedited and historical composites ingested into VegDRI are stored for future use

in event of problems or new processing requirements. The weekly products are stored for use in

subsequent years ranking. SPI data are stored for future use when models are recalibrated.
79

Model parameter and input files are stored for comparative analysis. RAID storage is used for all

data and is estimated to provide adequate storage until the end of 2015.

Satellite Data Continuity--AVHRR-MODIS Data Translation

Data continuity from daily polar orbiting satellites is extremely important to successful drought

monitoring. Since drought reveals itself in relation to the normal climate condition at a

specific location, the continuity of repeated observations allows for the objective calculation of

both normal and anomalous conditions. Both AVHRR and MODIS sensors are incorporated

into VegDRI for drought monitoring. In the United States, regular vegetation monitoring started

in 1989 (well before the MODIS era) and continues today, facilitated by EROS direct reception

and processing of AVHRR data into national composites on weekly and 14-day time steps

(Eidenshink 2006). A data processing system produces and delivers AVHRR data to users

within 30 hours of satellite overpass. The 20+ year satellite record created by this system is used

to determine normal vegetation conditions in an operational fashion.

The long-term continuity of the AVHRR and MODIS data records (and extending into the future

with data from planned future instruments) is critical for monitoring land surface conditions. It

is a goal of the USGS to develop a linkage between AVHRR and MODIS to ensure the

continuity of a long-term record of vegetation condition. Recent research has demonstrated

development of translation functions based, in some cases, on the inter-comparison of VI values

and in other cases, on methods of inter-calibrating the pertinent surface reflectance bands. Inter-

sensor compatibility was examined and translation equations were derived to seamlessly extend

the multi-sensor data record. Multi-sensor translations were based on an overlapping period of
80

observations using geometric mean regressions (GMRs) to treat variations in both AVHRR and

MODIS datasets equally.

Two AVHRR-MODIS data translation equations (Eq. A and Eq. B ) have been developed to

support United States drought monitoring (T. Miura, pers. communication). Phenological regions

(Gu et al., 2010) were used as a geographic framework for extracting data to derive the

equations. Eq. A was developed using a combination of canonical correlation analysis and GMR

methods and is a phenological region-independent translation equation of the form:

NDVIp AVHRR = 9.89 + 0.76 * (0.99 * NDVI MODIS + 0.002 * NDVI MODIS 2)

Where NDVIp AVHRR is the NDVI from AVHRR data, and NDVI MODIS is the NDVI from

the MODIS instrument.

Eq. B was developed based on a maximum likelihood regression or estimation method and the

phenological regions framework. Eq. B formulation was accomplished with 1 km2 resolution

AVHRR and eMODIS Terra NDVI data for 2005-2007. There are 19 phenological class-

dependent translation equations and one general (or default) translation equation in Eq. B. Eq. B

was evaluated with 2008 satellite data using mean difference and root mean square error (RMSE)

between NDVIp AVHRR and NDVI MODIS. Overall the RMSE between translated AVHRR

NDVI and MODIS NDVI was <0.03 and the mean differences in NDVIp AVHRR for all but one

of the phenological classes was < 0.02 NDVI units.


81

Operational Status and Future Plans for the VegDRI System

Operational Performance

As of 8/10/2010, the VegDRI system was fully functional and producing weekly drought

products delivered to meet the schedule requirements of the USDM authors. USGS EROS has

transitioned the system to operational status under the direction of the EROS Long Term Archive

project as of 12/31/2010.

Challenges, Risks, and Recommendations

VegDRI is based on an empirical modeling strategy and, as such, is not a true turn-key system.

Model recalibration is advised on a yearly schedule to update the historical archive that is used to

calculate the averages and anomalies of the time-series satellite VI measures. This is especially

needed because of the fairly short historical period for the satellite data (now 21 years for the

AVHRR and 10 years for Terra MODIS). Model recalibration requires participation and support

from science staff at the NDMC and EROS. Risks include inadequate funding or losing staff

institutional knowledge. However, the current system can continue to operate without model

recalibration.

There are additional risks to input data streams incorporated into VegDRI. The weekly gridded

Standardized Precipitation Index data is periodically (but infrequently) delivered either as an

incomplete raster file or delivered later than the 00:01 a.m. Monday deadline due to source data

or software issues at the High Plains Regional Climate Center (which is co-located with the

National Drought Mitigation Center). The NDMC staff are dedicated to solving these problems

as they occur.
82

The eMODIS system at EROS is going through a transition to operations at the USGS EROS at

the time of this report. Efforts are ongoing to improve and fortify the eMODIS hardware and

software systems, improving redundancies and reducing risk. An important risk reduction

strategy is to have access to multiple satellite VI records for input into VegDRI. Both Terra

(morning) and Aqua (afternoon) composite data are being produced from the eMODIS system.

This project has not directly addressed a potential future transition of the VegDRI system after

the conclusion of the lifespans of the Terra and Aqua MODIS instruments. Given the success of

the index, we recommend a transition of VegDRI to ingest VIIRS observations, given the

similarities in spatial resolution and multi-spectral band widths of this instrument compared to

MODIS.

Results, Communication, and Dissemination

Accessibility/Distribution

The VegDRI is currently distributed as raster image data (geotiff format), graphical maps, and

interactive map layers in a web map viewer (http://vegdri.cr.usgs.gov/viewer/viewer.php).

Improved access to historical VegDRI archives is desirable.

Multiple websites display some version of the VegDRI products:

1. USGS VegDRI website: http://vegdri.cr.usgs.gov/

2. NIDIS website: http://www.drought.gov/

3. NDMC Website: http://drought.unl.edu/


83

Starting in mid-2010, EROS has also been staging weekly graphic maps (see Figure 1) and

sending notices out to the USDM authors each week via the Drought Exploder listserve.

Currently, around 290 local experts across 45 states are members of this listserve.

Outreach and Communication

During the course of the project (2006-2010), various methods have been used to carry out

project outreach and communication. Project team members have led workshops, given multiple

oral presentations, made posters, and published in the scientific literature.

NIDIS Workshop

Over thirty researchers, scientists, and natural resource managers representing a variety of

federal and non-federal agencies convened in at the National Integrated Drought Information

System (NIDIS) Contributions of Satellite Remote Sensing to Drought Monitoring Workshop

at NOAA in Boulder, Colorado on February 6-7, 2008.

Project Presentations

1. Brown, J.F., VegDRI evaluation: Focus on Owyhee and Upper Colorado Basins,

VegDRI/VegOUT Workshop, Boise, ID, Jul 27, 2010 [ORAL PRESENTATION].

2. Brown, J.F., Advantages of near-real time satellite data for operational drought

monitoring: the utility of MODIS and AVHRR, Annual Meeting of the Association of

American Geographers, Washington, D.C., Apr 14-18, 2010 [ORAL PRESENTATION].


84

3. Brown, J.F., EROS NASA drought project status and plans, Drought Forum, Austin TX,

Oct 6, 2009 [ORAL PRESENTATION].

4. Brown, J.F., Anderson, M., Wardlow, B.D., and Svoboda, M., Remote sensing

techniques for monitoring drought hazards: an intercomparison, 2009 AGU Fall Meeting,

San Francisco, CA, Dec 14-18, 2009 [ORAL PRESENTATION].

5. Brown, J.F., Miura, T., Gu, Y., Jenkerson, C., and Wardlow, B.D., Utilizing a multi-

sensor satellite time series in real-time drought monitoring across the United States, 2009

Joint Assembly of the American Geophysical Union, Toronto, Ontario, Canada May 24-

27, 2009 [ORAL PRESENTATION].

6. Gu, Y., Brown, J.F. Phenological classification of the United States: A framework for

vegetation drought monitoring, 2009 Annual Meeting of the Association of American

Geographers, Las Vegas, NV, Mar 22-27, 2009 [ORAL PRESENTATION].

7. Brown, J.F., Wardlow, B.D., Tadesse, T., Callahan, K., and Pervez, M.S., Monitoring

recent drought effects on corn yields across the Corn Belt with the Vegetation Drought

Response Index, 2009 Annual Meeting of the Association of American Geographers, Las

Vegas, NV, Mar 22-27, 2009 [ORAL PRESENTATION].


85

8. Brown, J.F., Wardlow, B.D., Tadesse, T., and Gu, Y., Improving decision support for

drought using new geospatial models and online tools, ACES: A Conference for

Ecosystem Services, Naples, FL, Dec 8-11, 2008 [POSTER].

9. Brown, J.F., Pervez, S., Wardlow, B., Tadesse, T. and Callahan, K., Assessment of 2006

and 2007 drought patterns in the vegetation drought response index across Nebraska,

Pecora 17 Symposium, Denver, CO Nov 18, 2008 [ORAL PRESENTATION].

10. Brown, J.F., and Hayes, M.J., National drought monitoring progress and future plans,

Geography staff briefing, USGS Headquarters, Reston, VA, Aug 17, 2008 [ORAL

PRESENTATION].

11. Brown, J.F., Wardlow, B., Gu, Y., Tadesse, T., Pervez, S., Hayes, M., and Verdin, J.,

2008, The value of satellite observations in drought and phenology research and

applications, Climate Change Workshop 2008, North Platte, NE, May 19-22, 2008

[POSTER].

12. Brown, J.F., Wylie, B., Homer, C., and Zhu, Z., Integrating remote sensing tools for

National inventory and monitoring, Briefing for M. Myers, USGS Director, Sioux Falls,

SD. Mar 1, 2007 [ORAL PRESENTATION].

13. Brown, J.F., Verdin, J., Wardlow, B., and Tadesse, T., Remote sensing tools for

improving drought decision support, Managing Drought and Water Scarcity in


86

Vulnerable Environments--Creating a Roadmap for Change in the U.S., Geological

Society of America Meeting, Longmont, CO, Sep 18-20, 2006 [POSTER].

14. Gu, Y., Brown, J.F., Verdin, J.P., and Wardlow, B., A five-year analysis of MODIS

NDVI and NDWI for rangeland drought assessment, Global Vegetation Workshop,

Missoula, MT, Aug 7-10, 2006 [POSTER].

Publications

1. Brown, J.F. 2010, Drought monitoring with VegDRI, U.S. Geological Survey Fact Sheet

20103114, 2p. [http://pubs.usgs.gov/fs/2010/3114/]

2. Tadesse, T., Wardlow, B.D., Hayes, M.J., Svoboda, M.D., and Brown, J.F., 2010, The

Vegetation Outlook (VegOut): a new method for predicting vegetation seasonal

greenness, GIScience & Remote Sensing, 47 (1), p. 25-52.

3. Gu, Y., Brown, J.F., Miura, T., van Leeuwen, W., and Reed, B.C., 2010, Phenological

classification of the United States: A geographic framework for extending multi-sensor

time-series data, Remote Sensing, 2(2), 526-544; doi: 10.3390/rs2020526.

4. Gu, Y., Hunt E., Wardlow, B., Basara, J., Brown, J.F., and Verdin, J.P., 2008, Evaluation

of MODIS NDVI and NDWI for vegetation drought monitoring using Oklahoma

Mesonet soil moisture data, Geophysical Research Letters, 35 (22), L22401.


87

5. Brown, J.F., Wardlow, B.D., Tadesse, T., Hayes, M.J., and Reed, B.C., 2008, The

Vegetation Drought Response Index (VegDRI): a new integrated approach for

monitoring drought stress in vegetation: GIScience & Remote Sensing, 45 (1), p. 16-46.

6. Gu, Y., Brown, J.F., Verdin, J.P. and Wardlow, B., 2007, A five-year analysis of MODIS

NDVI and NDWI for grassland drought assessment over the central Great Plains of the

United States, Geophysical Research Letters, 34, L06407, doi:10.1029/2006GL029127.

References

Asner, G.P., Townsend, A.R., & Braswell, B.H. (2000). Satellite observation of El Nino effects

on Amazon forest phenology and productivity. Geophysical Research Letters, 27, 981-984

Breshears, D.D. (2005). Regional vegetation die-off in response to global-change-type drought.

Proceedings of the National Academy of Sciences, 102, 15144-15148

Brown, J.F., Wardlow, B.D., Tadesse, T., Hayes, M.J., & Reed, B.C. (2008). The Vegetation

Drought Response Index (VegDRI): A new integrated approach for monitoring drought stress in

vegetation. GIScience and Remote Sensing, 45, 16-46


88

Eidenshink, J. (2006). A 16-year time series of 1 km AVHRR satellite data of the Conterminous

United States and Alaska. Photogrammetric Engineering & Remote Sensing, 72, 1027-1035

Goetz, S., Fiske, G., & Bunn, A. (2006). Using satellite time-series data sets to analyze fire

disturbance and forest recovery across Canada. Remote Sensing of Environment, 101, 352-365

Jenkerson, C.B., Maiersperger, T., Schmidt, G., 2010. eMODIS: A user-friendly data source:

U.S. Geological Survey Open-File Report 20101055, 10 p.

Kogan, F.N. (1990). Remote sensing of weather impacts on vegetation in non-homogeneous

areas. International Journal of Remote Sensing, 11, 1405-1419

Parker, M.D., Ratcliffe, I.C., & Henebry, G.M. (2005). The July 2003 Dakota hailswaths:

Creation, characteristics, and possible impacts. Monthly Weather Review, 133, 1241-1260

Peters, A.J., Griffin, S.C., Vina, A., & Ji, L. (2000). Use of remotely sensed data for assessing

crop hail damage. Photogrammetric Engineering and Remote Sensing, 66, 1349-1355

Peters, A.J., Walter-Shea, E.A., Ji, L., Via, A., Hayes, M., & Svoboda, M.D. (2002). Drought

monitoring with NDVI-based Standardized Vegetation Index. Photogrammetric Engineering and

Remote Sensing, 68, 71-75


89

Svoboda, M., LeComte, D., Hayes, M., Heim, R., Gleason, K., Angel, J., Rippey, B., Tinker, R.,

Palecki, M., Stooksbury, D., Miskus, D., & Stephens, S. (2002). The Drought Monitor. Bulletin

of the American Meteorological Society, 83, 1181-1190

Wang, J., Rich, P.M., & Price, K.P. (2003). Temporal responses of NDVI to precipitation and

temperature in the central Great Plains, USA. International Journal of Remote Sensing, 24,

2345-2364

Wells, N., Goddard, S., & Hayes, M.J. (2004). A self-calibrating Palmer Drought Severity Index.

Journal of Climate, 17, 2335-2551


90

Appendix 1 (Section 2)
91
92

Section 3 (Led by National Drought Mitigation Center)

Integration of Soil Moisture and Vegetation Index for Drought

Monitoring

B. Wardlow and M. D. Svoboda

National Drought Mitigation Center, School of Natural Resources

University of Nebraska-Lincoln

815 Hardin Hall, 3310 Holdrege Street

Lincoln, NE 68583-0988 U.S.A.

The National Drought Mitigation Center (NDMC) served as the lead integrator of both soil

moisture and Vegetation Drought Response Index (VegDRI) products generated from this

project within the current operational U.S. Drought Monitor (USDM) framework

(http://drought.unl.edu/dm). The NDMC was specifically tasked with ingesting these results

from our project partners into the USDM (and National Integrated Drought Information System

(NIDIS) Portal (http://drought.gov) where relevant) as a means of evaluating and improving

decision support. The NDMC also helped direct development and benchmarked the results and

value of the products developed as part of this project. Within the USDM framework, case

studies were conducted by soliciting and obtaining feedback from the USDM monitoring

community (~290 members), including the weekly USDM authors themselves. The USDM

author team consists of a total of 10 members from the National Drought Mitigation Center,
93

USDA, NOAA (National Climatic Data Center and Climate Prediction Center) and the Western

Regional Climate Center (also sponsored by NOAA). In addition, the NDMC developed and

currently hosts and maintains the projects website at http://drought.unl.edu/nasa. The following

sections summarize in more detail the specific tasks and deliverables from the NDMC that were

highlighted above.

Product Development Guidance

In the initial stages of the project, the NDMC provided guidance to establish data and processing

requirements for the remote sensing products generated by the project participants for effective

integration into the operational USDM system. This guidance included defining specific product

requirements for the USDM system including product types, data format, update (frequency)

schedule, and a classification color scheme for the maps utilized within a GIS system. As

products developed using this initial set of requirements became available, the NDMC worked

with the USDM authors and USGS EROS, NASA JPL, and NOAA PSD project participants to

ensure that operational needs were met. This was an iterative process where adjustments and

modifications were made to various products and deliverables based on an interactive feedback

process between project partners and the USDM author community.

For the NASA JPL and NOAA PSD derivative products involving QuikSCAT soil moisture

change (SMC), information on the USDM ranking percentile criteria, cartographic color scheme

and operational USDM GIS methods, and the delta map and temporal production needs of the

USDM were provided to apply to the SMC product. In addition, feedback was provided on the
94

SMC maps on map legend development and the incorporation of USDM vector overlays on the

SMC maps as a valuable-added enhancement.

For the USGS EROS eMODIS VegDRI product, similar feedback and evaluation of VegDRI

maps were provided related to map color scheme, data format, updating schedule, and data

delivery mechanism. Product definition and formatting was prioritized early in the project in

order to ensure more effective integration of project data deliverables into the USDM system and

NIDIS portal. This product is currently ingested automatically by the NIDIS Portal Map Viewer

for the USDM author community on a weekly basis within the core suite of overlay products

used by the authors in the making of the Drought Monitor.

NDMC-NASA Partnership Website

An NDMC-NASA Partnership website (http://drought.unl.edu/nasa) was developed in Year 1

and is currently hosted and maintained by the NDMC for project-related information. To date, it

has not been made open to the public at the request of the Principal Investigators while it

contains information related product under development and testing, but it can be made fully

accessible at any time it is deemed necessary. An operational email list server

(nasadrought@unl.edu) was also created by the NDMC to facilitate communication among all

project participants throughout the life of the project. The intent of the project web site was to

serve as central access point and repository for the project partners remote sensing-derived

products, general information, project reports and presentations that were developed for the

advanced USDM prototype system that was being supported. Once the data feeds were
95

established and automated from USGS EROS and NOAA PSD to NIDIS, demonstrations of the

advanced USDM system that included these new NASA products was conducted for the USDM

community and all project partners. In addition to the project website, the NDMC worked with

the NIDIS Portal team to post eMODIS-based VegDRI and SMC delta maps within the USDM

Authors Community map viewer to allow USDM authors to visualize the information and

assess the validity and potential use of these new products into the making of the USDM.

The eMODIS VegDRI continues to be fed in near real-time to the NIDIS Portal and the USDM

authors on a weekly basis while the SMC was stopped because of an interruption of the

QuikSCAT data stream due to QuikSCAT instrument problems. At the completion of the

project, the NDMC will continue to update, enhance and maintain the project website (and list

server if needed) for project communication and facilitate the continued integration of project

products, as well as other NASA Earth Science data, into the USDM after this project expires, as

long as the NDMC has the in-house resources to maintain this ongoing activity.

eMODIS VegDRI Model and Map Development

In support of eMODIS VegDRI development, the NDMC prepared and maintained a database of

data inputs for eMODIS VegDRI and generated the models that were operationally implemented

in this project. The NDMC also coordinated with USGS and the High Plains Regional Climate

Center (HPRCC) the automated delivery of current climate data (i.e., Standardized Precipitation

Index, SPI) on a customized time schedule (early Monday morning) to allow the data to be
96

acquired and ingested into the eMODIS system for near real-time, weekly VegDRI map

production. Beta eMODIS VegDRI maps/data were evaluated by USDM authors at the NDMC

and feedback was provided regarding their formatting, cartographic presentation, and ease of

use. This information was used to define the final formatting of the weekly operational eMODIS

VegDRI maps that began to be routinely ingested into the USDM system in 2010.

Product Evaluation

The NDMC engaged USDM authors and other drought experts within the USDM drought

community to assess the accuracy and utility of the remote sensing-derived products as they

became available. Initial assessments were completed for both the SMC and VegDRI products,

but only for an abbreviated period because operational products were only available towards the

end of the project period.

Case studies were undertaken with JPL to evaluate how the SMC responded to drought

conditions and helped determine its potential for being utilized in the USDM system with several

conclusions reported earlier in Section 1. However, interruption of the SMC creation because of

sensor issues led us to rely only on a number of case studies instead of a near-real-time

operational growing season approach. A full investigation for the full capability of the SMC in

enhancing the USDM was thus impacted by the discontinuity of the QuikSCAT dataset. Future

data from the Oceansat-2 scatterometer (similar to QuikSCAT) will be useful to continue the

QuikSCAT data stream for drought applications.


97

For eMODIS VegDRI, full assessment of the accuracy and utility of these products began in the

summer 2010 and continued through the Fall. The eMODIS maps were presented to USDM

authors for their review within the NIDIS Portal map viewer and USDM authors at the NDMC

provided more directed feedback on the performance and complimentary information that

VegDRI could provide to the USDM process. Given the short evaluation period and the

relatively drought-free conditions across most of the United States in 2010, a complete and

thorough review of VegDRI was not possible. However, the initial investigation showed that the

drought conditions/patterns depicted by VegDRI were relatively consistent with the USDM and

other indicators, which suggests that VegDRI represents a higher spatial resolution product that

can be used to improve the characterization of agricultural drought patterns within the USDM.

This is critical in those vast areas which lack observed climate data, particularly over rangeland.

VegDRI has begun to establish traction in being utilized by the USDM author in their weekly

suite of USDM overlay maps. An extended, post-project evaluation of near real-time VegDRI

maps for the 2011 growing season by NDMC, USGS EROS, and the USDM community is

planned to more fully evaluate the accuracy and contribution of the eMODIS VegDRI tool.

Coordination with USDM Authors to Implement Remote Sensing Data Products into an

Advanced USDM Prototype System

Throughout the project, the NDMC continually engaged the USDM authors to communicate and

describe the information content of each project deliverable as they became available and

assisted them with the evaluation/integration of this information into the USDM development

process and within the USDM and NIDIS Portals. Feedback from USDM authors was collected,
98

organized, and communicated to other project partners by the NDMC to facilitate product

development and to promote the increased and continual application of the information. The

initial response, based on a limited amount of operational integration, was sufficiently favorable

to begin to include and test the eMODIS VegDRI product within their weekly suite of drought

monitoring products. Authors use the products with overlays of the USDM to aid them in the

placement of the D0-D4 line work on the map. This process has been automated and will

continue to operate after this project has ended, as long as resources can be leveraged by the

partners.

Communication of Project Activities

The NDMC has presented and highlighted project activities and the new tools and products

through numerous presentations and publications. The NDMC (Wardlow and Svoboda) worked

with other project participants to prepare book chapters for the VegDRI (lead author: Brian

Wardlow) and SMC (lead author: Son Nghiem) approaches for submission to a new book

entitled Remote Sensing and Drought: Innovative Monitoring Approaches by CRC Press (co-

editors: Brian Wardlow - NDMC, Jim Verdin USGS & NIDIS, and Martha Anderson USDA

ARS). The NDMC (Svoboda and Wardlow) is currently assisting NASA JPL lead, Son Nghiem,

in the writing of a peer-reviewed journal article on some of the project deliverables. Brian

Wardlow, of the NDMC, has also summarized the VegDRI work in a World Meteorological

Organization (WMO) publication.


99

Mark Svoboda and Brian Wardlow have presented on various aspects of the NDMCs role in this

project at the past three Drought Monitor Forums (both U.S. and North American venues), which

are held annually and bring together ~100 drought experts and media from Canada, Mexico, and

the United States. Potential continental/global applications of the VegDRI and SMC information

have also been discussed at these same U.S. and North American Drought Monitor (USDM and

NADM) Forums. Project deliverables and products have also been shared with the National

Integrated Drought Information System (NIDIS) Program Office (tasked with, among other

things, developing a national drought early warning system) and presented at each of the NIDIS

Pilot Drought Early Warning System (DEWS) basin meetings for the Upper Colorado River and

Apalachicola-Chattahoochee-Flint. We plan to do the same for the California NIDIS Pilot

Project, scheduled to start in 2011. Internationally, the NDMC has presented project activities at

a WMO-sponsored workshop in West Africa and at various meetings involving the Global Earth

Observations (GEO) office and its potential usage within the Global Earth Observation System

of Systems (GEOSS).

Publications

Wardlow, B.D., T. Tadesse, J. Brown, K. Callahan, S. Swain, and E. Hunt, 2012. Integration of

Satellite, Climate, and Biophysical Data for Drought Monitoring of Vegetation. In B. Wardlow

et al. (Eds), Remote Sensing of Drought: Innovative Monitoring Approaches, CRC Press, Boca

Raton, FL, In press.


100

Nghiem, S., B.D. Wardlow, D. Allured, M.D. Svoboda, D. LeComte, M. Rosencrans, S.K. Chan,

and G. Neumann, 2012. Microwave remote sensing of soil moisture Science and Applications.

In B. Wardlow et al. (Eds), Remote Sensing of Drought: Innovative Monitoring Approaches,

CRC Press, Boca Raton, FL, In press.

Presentations

Wardlow, B.D., GIScience Activities at the National Drought Mitigation Center, USDA

National Agricultural Statistics Service (NASS), Fairfax, VA., August 12, 2010.

Wardlow, B.D., An Overview of Remote Sensing Activities at the National Drought Mitigation

Center, Hydrology and Remote Sensing Laboratory, USDA Agricultural Research Service (ARS),

Beltsville, MD, August 11, 2010.

Wardlow, B.D., The Future of Remote Sensing Applications for Drought Monitoring, World

Meteorological Organization (WMO) Drought Workshop, Lincoln, NE, December 10, 2009.

Wardlow, B.D., Vegetation Drought Response Index (VegDRI): 2009 Update and Ongoing

Activities, U.S. Drought Monitor Forum, Austin, TX, October 8, 2009.

Wardlow, B.D., Remote Sensing and Drought An Overview and Opportunities for Mali, WMO

Drought Workshop, Bamako, Mali, September 17, 2009.


101

Wardlow, B.D., Remote Sensing and Drought Monitoring - New Tools and Future Directions,

National Hydrologic Warning Council Conference, Vail, CO, May 5, 2009.

Wardlow, B.D., VegDRI A New Hybrid Drought Index for Monitoring Vegetation in the U.S.,

U.S.-Canada GEO Bilateral Workshop on Ice and Water, National Science Foundation (NSF),

Arlington, VA, October 28, 2008.

Wardlow, B.D., Vegetation Drought Response Index (VegDRI): A Hybrid-Based Approach for

Vegetation Drought Monitoring, North American Drought Monitor Workshop, Ottawa, Canada,

October 16, 2008.


102

Section 4 (Led by Dartmouth College)

MODIS Surface Water Status

G. Robert Brakenridge

Director, Dartmouth Flood Observatory

http://floodobservatory.colorado.edu/

Senior Research Scientist

CSDMS, INSTAAR, University of Colorado

Campus Box 450, Boulder, CO 80309-0450 USA

Office: 303-735-5485, Cell: 603-252-0659

Email: Robert.Brakenridge@Colorado.edu

The Dartmouth portion of this project was to develop MODIS surface water status products, and

use them as a metric for evaluating the system performance of the Drought Monitor. This

project is a transfer from its former location at Dartmouth College to University of Colorado and

the work plan is extended for completion of the remaining tasks in the current year. This section

contains the research progress to date. As noted in the original proposal, U.S. stream gaging

stations, which also measure surface water, provide a critical input to the present drought

assessment algorithms. Streamflow integrates the effects of temperature, precipitation, watershed

geomorphology and geology, and antecedent conditions. The U.S. has been divided into 5348
103

hydrologic accounting units (watersheds) and the present array of gaging stations provide

information on the status of moisture deficits or surpluses for many of these units, and including

contributing units further upstream. However, many of these stations measure major rivers and

their linked tributary systems, and data from these are not independent measurements of the local

contributing areas until the imported water from upstream watersheds is accounted for. This is

not accomplished in the 7 day streamflow averages prepared by the USGS (used in the USDM)

and presented in the form of percentiles for period of record: instead, as one moves downstream

along all streams and rivers, the additional gaging station discharge percentile information

applies to the entire upstream contributing areas instead of the local incremental contributing

areas. Remote sensing offers, instead, the opportunity for discrete land parcel or catchment-unit

measurements of existing surface water status. As well, there exist hundreds of U.S. ponds and

reservoirs, including many small water bodies < 100 km2, which are ungauged (Figure 1)

Figure 1. MODIS-based information on ungauged reservoir variability in northern Oklahoma

(left) and southern Texas (right). Different shades of blue show MODIS hydrographic mapping
104

during normal years (light blue) and during intervals of moisture surpluses (darker blues).

Red crosses show all operating gaging stations: note that many reservoirs are not measured by

the in situ network, but can be by MODIS. Also shown are the hydrologic units, whose

hydrologic status can be estimated by precise measurements of the ungaged reservoirs.

Across much of the U.S, these water bodies are critical to local agriculture, livestock, recreation,

and other uses, and their areal extents are also observable indicators of drought status. Increases

in contributing watershed runoff to many reservoirs (depending on its mode of operation) are

accompanied by changes in reservoir water surface areas. If these areas can be consistently,

frequently, and economically monitored, through time, then such data can provide significant

enhancement and testing of map evaluations of regional drought conditions (such as the U.S.

Drought Monitor, or USDM). We proposed to obtain such measurements, and to facilitate their

incorporation into the USDM.

During the first project year, we evaluated the capability of MODIS data to numerically estimate

the surface water status of selected U.S. hydrologic accounting units. We began as planned, by

testing change detection/measurement approaches (Figure 2). Band 2 on both operating MODIS

sensors very sensitively discriminates surface water changes at a spatial resolution of 250 m: the

challenge is to design a processing methodology that, working around the abundant noise

(changes in cloud cover and cloud shadows, terrain shadowing, seasonal changes in vegetation

and illumination, atmospheric changes), can produce a weekly MODIS-based product covering

all of the U.S. (excluding Alaska, Hawaii, and territories) and which can be directly compared to

the USDM).
105

Figure 2. MODIS band 2 near-IR geocoded 250 m image of southern Indiana (Ohio River is in

lower portion of scene) during dry conditions (left, September 18, 2004), compared to an

interval of moderate moisture surplus (right, change detection product using the dry MODIS

scene and one from July 16, 2003). Red shows increasing water as inferred by declining band 2

radiances between the two dry and wet scenes. Surface water variability at several small

reservoirs as well as along the Wabash River is detected.

During further work, we developed several other methods for surface water classification and

mapping. Change detection for water can proceed either by direct comparison of raster-based

data (as in Figure 2) or as Geographical Information System (GIS) water polygon-based

comparison of mapped water. Mapping the water, in this case, makes use of remote sensing

image classification procedures to discriminate water, then polygons are fit to the classified

water areas, and then, within the GIS, map displays can show areas of expanded or decreased

water area (Figures 3 and 4). Partly with the support of this projects first year, we have
106

developed an efficient water classification algorithm using a widely available MODIS data

product (the rapidfire Rapid Response Subsets).

Figure 3. Sample of the planned MODIS-based Reduced Surface Water display, for one of the

twelve 10 deg. x 10 deg. map sheets covering the coterminous U.S. This display is here shown at

much reduced scale. Small areas of yellow are dry land, as imaged in March of 2009, compared

to that imaged by the same sensor using the same classification techniques in March of 2005,
107

when no drought conditions were occurring according to the USDM. Display thus shows at

maximum sensor resolution all areas of shrinkage of ponds, reservoirs, lakes, and rivers. Careful

choice of non-drought comparison dates is essential. These GIS data can be easily ingested into

an all-U.S. mosaic at a scale comparable to the USDM, or overlain onto the USDM map. Where

actual dimensions of water area shrinkage are too small to show at USDM scale, GIS

cartography techniques still allow display of the locations of such change.

Figure 4. A small portion of Figure 3, illustrating the preservation of detailed spatial

information. Planned work for year 2 will provide detailed, regional maps depicting which water

bodies are affected, while also providing a unified weekly all-US map showing the locations of

surface water reductions as observed via MODIS.

During the first year, the advantage of the GIS-based approach became apparent. It offers the

possibility to retain the relatively fine spatial resolution of the two MODIS 250 m bands, by
108

performing new data retrieval and water classification and mapping over discrete, 10 deg x 10

deg portions of the U.S. (Figure 3). It also allows for easy combination of such detailed, regional

displays into all-U.S. displays that can be directly compared to the USDM (to be accomplished

in the second and final project year).

Note that in the original proposal, we also considered a gauging reach approach, whereby

changes in reach surface water area along rivers could be used as a predictor of river discharge

(when river discharge rises, mean river width and total reach water area increase), and a paired

measurement parcel/calibration parcel technique could be used to produce very well-calibrated

time series of local water area changes, including small reservoir water surface areas. We

proposed to incorporate this approach within a semi-automated system that could ingest these

same NASA MODIS data over selected U.S. hydrologic accounting units, with the first year of

work planned to develop this system, the second to provide demonstrations of its use for

enhanced-USDM validation, and the third to transition to operational incorporation of MODIS

surface water data. However, NASA funding limitations de-scoped the effort to a two year

run, and with the input being limited to USDM validation rather than work toward actual

MODIS water area data input into the USDM operational product. We have thus focused on the

MODIS mapping approach, as per Figures 3 and 4, and because the remaining work then

requires only expansion of our trial approach to cover 12 US map sheets, and production of

weekly surface water change maps at the much-reduced spatial scale of the USDM (Figure 5).

These remain feasible within the time and budgetary constraints of the second and final project

year.
109

Figure 5. Left: location of the planned map sheets. Right: example of USDM output. Surface

water data from each sheet will be depicted at much-reduced resolution on a U.S. map at same

scale and in same projection as the USDM.

During the first project year, we also prepared a web site location to host these products (for

example, http://www.dartmouth.edu/~floods/hydrography/W100N40Drought.html) for the

above. Joint publications have been listed in the first section of this report.
110

Participants in the 6th U.S. Drought Monitor Forum, Lower Colorado River Authority (Redbud

Center), Austin, TX October 7-8, 2009. Members of the NASA drought project team, and Brad

Doorn from NASA Headquarters participated in this forum.

Hosted by: Lower Colorado River Authority

Sponsored by: National Drought Mitigation Center

You might also like