Vous êtes sur la page 1sur 20

Remote Sensing of Environment 85 (2003) 1 – 20

www.elsevier.com/locate/rse

Review article
Remote sensing of soil salinity: potentials and constraints
G.I. Metternicht a,*, J.A. Zinck b
a
Department of Spatial Sciences, Curtin University of Technology, GPO Box U 1987, Perth, WA 6845, Australia
b
International Institute for Geo-Information Science and Earth Observation (ITC), Department of Earth Systems Analysis (ESA),
PO Box 6, 7500 AA Enschede, The Netherlands
Received 24 June 2002; received in revised form 29 October 2002; accepted 2 November 2002

Abstract

Soil salinity caused by natural or human-induced processes is a major environmental hazard. The global extent of primary salt-affected
soils is about 955 M ha, while secondary salinization affects some 77 M ha, with 58% of these in irrigated areas. Nearly 20% of all irrigated
land is salt-affected, and this proportion tends to increase in spite of considerable efforts dedicated to land reclamation. This requires careful
monitoring of the soil salinity status and variation to curb degradation trends, and secure sustainable land use and management.
Multitemporal optical and microwave remote sensing can significantly contribute to detecting temporal changes of salt-related surface
features. Airborne geophysics and ground-based electromagnetic induction meters, combined with ground data, have shown potential for
mapping depth of salinity occurrence. This paper reviews various sensors (e.g. aerial photographs, satellite- and airborne multispectral
sensors, microwave sensors, video imagery, airborne geophysics, hyperspectral sensors, and electromagnetic induction meters) and
approaches used for remote identification and mapping of salt-affected areas. Constraints on the use of remote sensing data for mapping salt-
affected areas are shown related to the spectral behaviour of salt types, spatial distribution of salts on the terrain surface, temporal changes on
salinity, interference of vegetation, and spectral confusions with other terrain surfaces.
As raw remote sensing data need substantial transformation for proper feature recognition and mapping, techniques such as spectral
unmixing, maximum likelihood classification, fuzzy classification, band ratioing, principal components analysis, and correlation equations
are discussed. Lastly, the paper presents modelling of temporal and spatial changes of salinity using combined approaches that incorporate
different data fusion and data integration techniques.
D 2003 Elsevier Science Inc. All rights reserved.

Keywords: Review; Salinity; Alkalinity; Videography; Remote sensing; Hyperspectral; Microwave; Image classification; Modelling; Monitoring

1. Introduction approximately 10 times the size of a country like Venezuela


or 20 times the size of France. In addition to these naturally
Soil salinity might not be as dramatic and damaging as salt-affected areas, about 77 M ha have been salinized as a
earthquakes or large-scale landslides, but it is certainly a consequence of human activities, with 58% of these con-
severe environmental hazard. This results in increasing centrated in irrigated areas. On average, 20% of the world’s
impact on crop yields and agricultural production in both irrigated lands are affected by salts, but this figure increases
dry and irrigated areas due to poor land and water manage- to more than 30% in countries such as Egypt, Iran and
ment and expansion of the agricultural frontier into marginal Argentina (Ghassemi et al., 1995). In the future, more
drylands. drylands will be put into agricultural production because
Statistics relating to the extent of salt-affected areas vary of increasing population pressure. This will mainly be
according to authors, but estimates are in general close to 1 achieved with irrigation and will thus expand the saliniza-
billion hectares, which represent about 7% of the earth’s tion hazard.
continental extent (Ghassemi, Jakeman, & Nix, 1995) or Furthermore, salinity also affects other major soil degra-
dation phenomena such as soil dispersion, increased soil
erosion, and engineering problems. When soil salinization is
* Corresponding author. Tel.: +61-892663935; fax: +61-892662703.
assessed in economic terms, reasons to be concerned about
E-mail addresses: Metternicht@vesta.curtin.edu.au (G.I. Metternicht), it become more apparent. For instance, the economic
Zincka@itc.nl (J.A. Zinck). damage caused by secondary salinization was estimated at

0034-4257/03/$ - see front matter D 2003 Elsevier Science Inc. All rights reserved.
doi:10.1016/S0034-4257(02)00188-8
2 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

US$750 million per year for the Colorado River Basin in the sources, and data transformation and classification techni-
USA, US$300 million per year for the Punjab and North- ques usually implemented for detection of salt-affected
west Frontier Provinces in Pakistan, and US$208 million areas. The second part deals with modelling of temporal
per year for the Murray– Darling Basin in Australia (Ghas- and spatial changes of salinity, and alternative approaches
semi et al., 1995). Frequently, these costs do not consider that incorporate different data fusion and data integration
the losses on property values of farms with degraded land, techniques to overcome some constraints of current remote
and other indirect costs such as eutrophication of rivers and sensing technologies.
estuaries, damage to infrastructure (including roads and
buildings), and the social cost of farm businesses. A study
focussed on dryland salinity risk assessment in Australia 2. Constraints on the use of remote sensing data
(National Land and Water Resources Audit, 2000) estimates
that some 20,000 km of major roads and 1600 km of The presence of salts at the terrain surface can be
railways are already at risk, with a potential increase to detected from remotely sensed data either directly on bare
52,000 and 3600 km, respectively, by 2050, unless proper soils, with salt efflorescence and crust, or indirectly through
measures to halt degradation are taken. However, remedial vegetation type and growth as these are controlled or
actions require reliable information to help set priorities and affected by salinity (Mougenot, Pouget, & Epema, 1993).
to choose the type of action that is most appropriate in each Surface salinity is a highly dynamic process, causing
case. Decision-makers and growers need confidence that all identification constraints derived from the proper behavior
technical estimates and data provided to them are reliable of the salt features, spectrally, spatially, and temporally. Salt
and robust, as the economic and social effects of over- or detection can also be blurred by the presence of vegetation
underestimating the extent, magnitude, and spatial distribu- and other surface features, which might contribute to creat-
tion of salinity can be disastrous. ing spectral confusions with the salt reflectance properties.
To keep track of changes in salinity and anticipate further
degradation, monitoring is needed so that proper and timely 2.1. Spectral behaviour of salt types
decisions can be made to modify the management practices
or undertake reclamation and rehabilitation. Monitoring Ground observations and radiometric measurements
salinity means first identifying the places where salts con- show that quantity and mineralogy of salts, together with
centrate and, secondly, detecting the temporal and spatial soil moisture, colour, and roughness, are the main factors
changes in this occurrence. Both largely depend on the affecting salt reflectance. Mougenot et al. (1993) mention
peculiar way salts distribute at the soil surface and within reflectance variations according to variable terrain surface
the soil mantle, and on the capability of the remote sensing conditions, including crusts with or without low salt evi-
tools to identify salts (Zinck, 2001). dence, salt crusts less than 1 mm to 1 m thick, puffy
This paper attempts, from an end-user’s point of view, to structures containing soil aggregates and salt crystals
review some of the key issues concerning the use of remote (0.5 – 5 mm) derived from salty clays and sometimes from
sensing data for soil salinity mapping and monitoring from a salt crusts, and wind erosion of puffy structure layers. Fig. 1
set of recent papers selected, among others, because they shows some examples of salt- or sodium-related surface
provide appropriate supporting data to the points addressed. features that influence salt identification from remote sens-
Salts tend to concentrate on the soil surface in dry and ing data.
irrigated areas. As salinity increases, more salts will appear Salts also cause variations in surface roughness. Fig. 2
at the soil surface, favouring the use of conventional remote illustrates the influence of different salt types and land
sensing tools. At the same time, the proper nature of the management practices on the reflectance of crusted surfaces.
salts, the way they distribute at the terrain surface, and the Salt mineralogy produces distinctive macromorphological
fast rate of change over time place some constraints on features at the soil surface. For instance, puffy crusts form
using remote sensing technologies because not all remote because of the abundance of sodium sulphates, while
sensing data offer the same possibilities for identifying salt- smooth salty crusts are due to the presence of chlorides
affected areas and monitoring their evolution. (Driessen & Schoorl, 1973; Eghbalm, Southard, & Whittig,
The first part of this paper highlights constraints on the 1989). In general, salty crusts are smoother than nonsaline,
use of remote sensing data, advantages of the various data usually cultivated surfaces and have an overall higher

Fig. 1. Examples of surface features, or combinations thereof, that influence salt identification from remotely sensed imagery because of distribution pattern,
morphology, spectral similarity, spectral interference, and temporal change. (a) Puffy sodium sulphate-rich crust, characteristic of alkaline areas; (b) puffy crust
containing soil aggregates, spots of thin salt crust, and salt-tolerant Chenopodiaceae; (c) sodium sulphate-rich crust broken by cattle or sheep trampling, leaving
the underlying dispersed soil material exposed to wind erosion; (d) alkaline areas with pH values up to 9, used for cultivation of salt-tolerant crops (e.g. barley,
alfalfa); (e) salt crust maximized at the end of the dry season; (f) chloride sulphate-dominated surface, with patches of salt-tolerant Cynodon dactilon; (g)
surface roughness caused by ploughing; (h) irrigated agricultural field with local salt concentration impeding alfalfa growth; (i) dispersed patches of halophytic
vegetation on soils severely affected by salinity – alkalinity; (j) poor barley development in a salt-affected paddock; (k) dark reddish saline crust; (l) structural
crust.
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 3

reflectance in the visible and near-infrared wave bands. tively higher roughness produced by large clods in ploughed
Cracks, common in silt- and clay-rich nonsaline crusts, areas (Fig. 1).
lower the overall crust reflectance. Similarly, reflectance Salt mineralogy (e.g. carbonates, sulphates, chlorides)
of cultivated nonsaline surfaces decreases because of rela- determines the presence (or absence) of absorption bands in
4 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

Fig. 2. Reflectance variations of surface features due to differences in crusting and land management practices. Data recorded with a ground-based Crop Scan
radiometer.

the electromagnetic spectrum. For instance, pure halite vibrations of constitutional water molecules, as reported by
(NaCl) is transparent and its chemical composition and Mulders (1987) and Siegal and Gillespie (1980), and shown
structure preclude absorption in the visible and near to in Fig. 3. Carbonate absorption features are also reported at
thermal infrared bands (Hunt, Salisbury, & Lenhoff, 2.34-Am wavelength (Siegal & Gillespie, 1980). Middle-
1972), as shown in Fig. 3. On the contrary, carbonates infrared bands, reflecting water and OH absorption, allow
present absorption features in the thermal range (between 11 differentiating between chlorides (as halite) and sulphates
and 12 Am) due to internal vibrations of the CO32 group, when both are dry. Mulders (1987) reports the 1.50– 1.73-
whereas sulphate anions have an absorption band near 10.2 Am range as one of the absorption bands for soil surface
Am caused by overtones or combination tones of internal features containing gypsum (CaSO4H2O).

Fig. 3. (a) Spectra of gypsum (CaSO42H2O), halite (NaCl), calcium carbonate (CaCO3), sodium bicarbonate (NaHCO3), and sodium sulphate (Na2SO4) in the
visible, near, and midinfrared (0.4 – 2.5 Am), as recorded by the GER 3700 spectroradiometer. Redrawn from Hovari (2002). (b) Spectra of carbonates,
chlorides, and sulphates (gypsum, anhydrite, apatite, and halite) in the thermal infrared (6 – 20 Am). Modified from Lane and Christensen (1997, 1998).
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 5

Under laboratory conditions, Csillag, Pásztor, and Biehl has higher reflectance than the nonsaline soil surface.
(1993) identified six spectral ranges characterising the However, alteration of the saline crust by trampling
salinity status of soils exposed to different salinization and increases the surface roughness, which causes reflectance
alkalinization processes, including the visible (0.55 –0.77 to decrease at similar salt contents. The shape of the spectral
Am), near-infrared (0.9 –1.3 Am) and middle-infrared (1.94 – curve does not change, but the overall reflectance does.
2.15, 2.15 –2.3, 2.33 –2.4 Am) bands. With low moisture
content, salts have high reflectance values in the visible 2.3. Temporal changes in salinity
(especially the blue) region of the spectrum. Low reflectance
occurring in the middle-infrared bands is attributed to the Good timing for passive (e.g. optical and infrared) remote
presence of hygroscopic water in salt minerals or high sensing data acquisition must take into account that salt
moisture content of fresh salts. Likewise, the type of salt identification is easiest at the end of the dry season (Fig. 1),
crystallization pattern and salt mineralogy interfere, as more as salts dissolve during the rainy season. In contrast to white
light is trapped in coarse lattices (e.g. gypsum) than in fine saline surfaces, pure alkaline soils are usually dark at surface,
lattices (e.g. halite). because excess sodium causes organic matter to disperse
when the soil is moist. On the other hand, the character-
2.2. Spatial patterns of salinity isation of saline soils using active (radar) imagery and
complex dielectric constants determined by radar backscat-
The tendency of salts to concentrate locally, in patchy tering inversion techniques requires some soil moisture, as
spots, creates variable mixtures of features within individual the measurements depend principally on dielectric constant
pixels. In addition, saline crusts vary considerably in struc- and permittivity determined by moisture conditions, chem-
ture, from smooth to rough, and in colour, from white to ical and biological compositions (Mougenot et al., 1993).
dark. These cause large variations in reflectance (Epema, For optical remote sensing, the possibility to identify
1992; Escadafal, 1989; Metternicht & Zinck, 1996). Fig. 4 salt-affected soils varies largely with moisture content, salt
shows the influence of surface roughness and crust colour pureness, coatings, and spectral contrast with other surface
on reflectance in the visible and near-infrared regions of the features. In general, moderately to highly saline areas are
spectrum (450 –800 nm), as recorded from ground-based easily detected, while low salinity levels and initial stages of
measurements, with an eight-band Crop Scan radiometer, of salinization frequently cause identification failures. For
soil surfaces belonging to very strongly saline and slightly instance, McGowen and Mallyon (1996) report significant
alkaline class. A decrease of two chroma units in the overestimation of saline areas when using Landsat TM data
Munsell soil colour chart, from 10 YR 6/5 to 10 YR 6/3, and a maximum likelihood classification. The reasons are
causes reflectance reduction in all bands. The saline crust adduced to land management practices, native vegetation

Fig. 4. Influence of surface morphology and colour on the reflectance of very strongly saline crusts. Data recorded with a ground-based Crop Scan radiometer.
6 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

and soil types present across the study area. The problem C. dactilon increases monotonically throughout the visible
was resolved by combining multitemporal imagery of differ- and near-infrared wave bands, indicating lower chlorophyll
ent cropping periods. On the other hand, underestimation of content. Metternicht (1996) found that halophytic vegetation
salt-affected surfaces using remote sensing has been attrib- was a good indicator to separate saline– alkaline areas from
uted to confusion between slightly saline and nonsaline soils nonaffected ones but caused spectral confusion among
(Jain, Sharma, Manchanda, & Singh, 1988; Manchanda & salinity – alkalinity classes. Very strongly saline – alkaline
Iyer, 1983). Hence, detection accuracy can be improved by areas, with soil reaction (pH)>9.5 and electrical conductivity
combining field data and multitemporal or multispectral (EC)>32 dS/m, were mistakenly labelled as strongly saline,
(e.g. optical, infrared, and microwave) image data, sup- slightly alkaline areas with pH 7 –8 and EC 8.1 – 32 dS/m
ported by geographic information system technology. because they had the same herbaceous cover of C. dactilon.
Although the presence of vegetation interferes with the
2.4. Interference of vegetation reflectance of salinity features, it can also be used as an
indirect indicator to correlate remote sensing data with soil
Salt- and sodium-affected areas are usually free of salinity (Szabo, Pásztor, Suba, & Varallyay, 1998; Zhang,
vegetation, but at lower concentrations specialized plant Ustin, Rejmankova, & Sanderson, 1997). For instance, high
communities can be present in natural conditions (e.g. correlation was found between NDVI values and soil
Salicornia spp.), and salt-tolerant crops such as barley, electrical conductivity, allowing to separate saline– alkaline
cotton, and alfalfa, can be cultivated (Fig. 1). Salt-tolerant soils from nonaffected ones, but not to discriminate between
vegetation modifies the overall spectral response pattern of saline and alkaline soils (Wiegand, Anderson, Lingle, &
salt-affected soils, especially in the green and red bands Escobar, 1994; Wiegand, Rhoades, Escobar, & Everitt,
(Metternicht, 1998a; Rao et al., 1995). 1994).
Halophytic salt-tolerant vegetation, such as Cynodon Some researchers propose to detect salinity distribution
dactilon and Chenopodiaceae, is common in saline and by using contrasted associations of vegetation and bare soils
alkaline areas. Data collected with a ground-based radio- (Metternicht & Zinck, 1997; Szabo et al., 1998; Zhang et al.,
meter (Skye Instruments, UK) in Bolivia shows the spectral 1997), while others recommend procedures of classification
curve of salt-tolerant Chenopodiaceae to be comparable to that take into account vegetation types at macroscale level
that of chlorophyll-rich vegetation, with strong absorption in and vegetation status, as influenced by salinity stress, at a
the visible range of the spectrum and higher reflectance in microscale level (Dale, Hulsman, & Chandica, 1986; Wie-
the near-infrared (Fig. 5). In contrast, the spectral curve of gand, Anderson, et al., 1994; Wiegand, Rhoades, et al.,

Fig. 5. Reflectance curves of salt-tolerant vegetation and saline crust. Data recorded with a ground-based Crop Scan radiometer.
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 7

1994). A range of optical, infrared, or microwave sensors, or gates are destroyed during the process of crust formation
combinations thereof, can be used to this end, as shown in and the finer soil particles form a seal or crust at the surface,
Section 3. leading to an increase of the overall reflectance of a soil (De
Among cultivated crops, cotton appears to be the ideal Jong & Epema, 2001). Fig. 1l shows the increased albedo of
indicator because it is largely cultivated in irrigated dry- a crusted surface. De Jong (1992) and Goldshleger et al.
lands. The health status of cotton plants and cotton yield (2001) conducted controlled spectral investigations of the
have been used as salinity indicators because of their strong behaviour of these structural crusts across the NIR and
correlation with electrical conductivity. Both criteria can be SWIR spectral regions, and reported that these crusts cause
estimated from colour infrared aerial photographs, video an overall increase in reflectance (e.g. up to 12%). Gold-
images, and multispectral images using spectral brightness shleger et al. found spectral information to be related to
coefficients, photo densities, and NDVI (Golovina, Minskiy, changes in particle size distribution and texture at the soil
Pankova, & Solovyev, 1992). surface, suggesting the crust might bias thematic remote
sensing of soils.
2.5. Spectral confusions with other terrain surface features Fig. 6 shows spectral curves of different crust types
collected in the Cochabamba area of Bolivia using a Crop
Features quite different in nature and dynamics can Scan multiband radiometer, with eight bands of 7– 12-nm
generate high levels of reflectance, similar to those of areas bandwidth located in the range between 450 and 800 nm
with high salt concentration. This is the case of surface (Metternicht, 1996). Spectral confusions occur between
features common in drylands, such as braided stream beds, salty crusts and bright silt loam structural crusts in the blue
eroded terrain surfaces with truncated soils, and nonsaline and green regions of the spectrum (450 – 550 nm). Surface
silt-rich structural crusts which cause spectral confusions brightness due to high silt content determines higher reflec-
because of their spectral similarity with salt-related features tance than that of a puffy or smooth salty crust. In general,
(Fig. 6). nonsaline crusts have lower reflectance and slightly different
Structural crusts are thin layers formed on the soil surface curve shape than salty crusts. The puffy, sodium sulphate-
after a rainstorm (or irrigation) event takes place (Gold- rich crusts seem to respond more to the influence of silt than
shleger, Ben-Dor, Benyamini, Agassi, & Blumber, 2001), to salt content, as the curve shape and level of reflectance
and it significantly changes the soil color. The soil aggre- are similar to those of the bright silt loam crust. A clear

Fig. 6. Influence of soil properties on the reflectance of saline and nonsaline crusts. Data recorded with a ground-based Crop Scan radiometer.
8 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

Table 1
Ground-based air- and satellite-borne sensors reported in the literature for detection and mapping of salt-affected areas
Sensor No. of bands Spectral range (Am) Spatial resolution
Satellite-borne sensors
Landsat TM4 and TM5, 7 (1 – 7), 8 (1 – 8) Visible, NIR, mid- and Bands 1 – 5 and 7: 30 m,
Landsat TM7-ETM + thermal infrared Band 6: 120 m, Band 8: 15 m
B1: 0.45 – 0.52
B2: 0.52 – 0.60
B3: 0.63 – 0.69
B4: 0.76 – 0.90
B5: 1.55 – 1.75
B6: 10.40 – 12.50
B7: 2.08 – 2.35
B8: 0.52 – 0.90 (pan)
SPOT 1 – 3 4 (Xs1 – 3 and Pan) Visible, NIR Xs and Xi: 20 m,
Xs1: 0.50 – 0.59 Pan and Mono: 10 m
Xs2: 0.61 – 0.68
Xs3: 0.79 – 0.89
SPOT 4 5 (Xi1 – 4 and Mono) Xi4: 1.58 – 1.75
Pan: 0.51 – 0.73
Mono: 0.61 – 0.68
LISS-III 4 Visible, NIR, midinfrared Bands 1 – 3: 23 m,
B1: 0.52 – 0.59 Band 4: 70 m
B2: 0.62 – 0.68
B3: 0.77 – 0.86
B4: 1.55 – 1.70
LISS-II 4 Visible, NIR 36.25 m
B1: 0.45 – 0.52
B2: 0.52 – 0.59
B3: 0.62 – 0.68
B4: 0.77 – 0.86
IRS-1C 1 0.5 – 0.75 (pan) 5.8 m
JERS-1 1 Microwave 18 – 12.5 m
L-band (23 cm)-HH pol

Airborne sensors
Aerial photographs B/W; colour infrared variable, depending on
flight height
Narrow-band videography 3 Visible, NIR 3.4 m
0.54 – 0.55
0.64 – 0.65
0.84 – 0.85
DMSV (digital multispectral 4 Visible, NIR variable: 0.25 – 2 m
video systems) 0.44 – 0.46
0.54 – 0.56
0.64 – 0.66
0.74 – 0.76
AIRSAR – TOPSAR 3 Microwave 10 m
P-, L-bands (full polarimetric),
C-band (VV polarization)
Hyperspectral Hymap 128 Visible, NIR, midinfrared 2 – 10 m
0.45 – 2.5
Hyperspectral DAIS-7915 79 0.4 – 12 3 – 20 m
0.4 – 1 (32 bands)
1.5 – 1.8 (8 bands)
2 – 2.5 (32 bands)
3 – 5 (1 band)
8.5 – 12.3 (6 bands)
Airborne geophysics Gravity
Magnetic
Electromagnetic
Gamma ray
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 9

Table 1 (continued)
Sensor No. of bands Spectral range (Am) Spatial resolution
Ground-based sensors
Electromagnetic induction meter Electromagnetic conductivity meter
(EM31, EM34-3, EM38, EM39) measures the bulk electrical
conductivity of soils
Crop Scan multiband radiometer 8 Visible, NIR
(Skye Instruments, UK) 0.36 – 1.1

decrease in reflectance is observed as the crusts become and assessment of saline problem areas. Multispectral video
nonsaline, together with lower contents of silt and clay. systems, recording spectral responses of vegetation in the
visible and infrared regions of the spectrum, have been
applied for mapping crop variations due to salinity. Everitt
3. Sources of remote sensing data et al. (1988) used narrow-band videography to detect and
estimate the extent of salt-affected soils in Texas. Wiegand,
A variety of remote sensing data has been used for Everitt, and Richardson (1992), Wiegand, Rhoades, et al.
identifying and monitoring salt-affected areas, including (1994) and Wiegand, Richardson, Escobar, and Gerbermann
aerial photographs, video images, infrared thermography, (1991) analyzed and mapped the response of cotton to soil
visible and infrared multispectral and microwave images, salinity using colour-infrared photographs and videography
and even remotely sensed data collected by airborne geo- in three bands covering the 0.84– 0.85-, 0.64 –0.65-, and
physics and electromagnetic induction (Table 1). 0.54 – 0.55-Am ranges of the spectrum, with a spatial reso-
lution of 3.4 m. By relating video data and field data, such
3.1. Aerial photographs as soil electrical conductivity, plant height, and percent bare
area, they determined the interrelations between plant, soil
Delineation of salt-affected soil units on aerial photo- salinity, and spectral observations. More precisely, for
graphs usually proceeds from a combination of geomorphic detecting salt-affected areas, colour-infrared composite and
features and grey tones or colours. Field verification allows red narrow-band images have proven to be better than green
relating these units to variations in salt content. Despite all and near-infrared narrow-band images (Escobar, Gerber-
the advances in spectral and spatial resolutions of satellite mann, & Alaniz, 1988; Wiegand et al., 1992; Wiegand,
and airborne remote sensing imagery, aerial photographs Rhoades, et al., 1994).
remain useful for detailed or historical studies. They have Other airborne sensors, such as the digital multispectral
been intensively used in soil salinity mapping, especially video (DMSV) or the digital multispectral imaging (DMSI)
colour-infrared photographs in which barren saline soils (in systems developed by SpecTerra Systems (2000), operate in
white) and salt-stressed crops (in reddish brown) can be similar regions of the spectrum but provide a wider cover-
easily discriminated from other soil surface and vegetation age, with one extra band in the 0.44– 0.46-Am range, and
features (Rao & Venkataratnam, 1991; Wiegand, Rhoades, variable pixel resolution as a function of the flight height
et al., 1994). Colour infrared aerial photography has also (0.25 –2 m). These systems have been tested for agricultural
been used in small areas requiring high resolution, which applications in Western Australia (Delfos, 2001; Drysdale,
cannot be provided by current satellite sensors, or to 2001; Islam & Metternicht, 2002; Metternicht, Gonzalez,
validate satellite data (Everitt, Escobar, Gerbermann, & Beeston, & Honey, 2002; Metternicht, Hageman, Beeston,
Alaniz, 1988; Manchanda & Iyer, 1983). Colour infrared & Honey, 2001; Stovold, Steber, & Smith, 1994). Due to
has proven to be most useful to identify variations in their flexibility in frequency and time of data acquisition,
salinity-induced plant stress. Because the response of crops near real-time display and high spatial resolution (see Table
to salinity severity usually varies from season to season and 1), these systems have shown potential for qualitative
from crop to crop, infrared photographs are a good means monitoring and rapid detection of vegetation degradation
for recording multitemporal and crop-to-crop responses to in agricultural landscapes, for keeping track of the effect of
salinity and changes in management practices (Ghassemi et farm management practices (e.g. fertilizer application, lim-
al., 1995). ing), and for assessing the performance of revegetation
programs in saline areas where the rate of tree survival is
3.2. Airborne videography and digital multispectral cam- an important factor to consider.
eras
3.3. Satellite sensors covering the visible to middle infrared
Videography appears as a promising means because of its
advantages in terms of recording ease, immediate playback, Extensive research on the application of satellite imagery
and real-time display, enabling preliminary identification covering the visible to infrared regions of the spectrum for
10 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

identification and mapping of saline areas has been con- canopy measured by infrared thermometers. It has also been
ducted in countries like Australia, Bolivia, China, Egypt, used to monitor saline seeps from irrigation channels by
India, Iran, and USA. Works by Csillag et al. (1993), Epema estimating the temperature difference between wet and dry
(1990), Evans and Caccetta (2000), Everitt et al. (1988), soils (Prathapar et al., 1990).
Kalra and Joshi (1996), Metternicht (1996), Metternicht and
Zinck (1996), Mougenot et al. (1993), Mulders (1987), Rao 3.5. Microwave sensing
et al. (1995), Srivastava, Tripathi, and Gokhale (1997), and
Verma, Saxena, Barthwal, and Deskmukh (1994) provide With the availability of space-borne synthetic aperture
illustrative application examples. In general, good results radar (SAR) systems (e.g. ERS 1/2, JERS-1, Radarsat),
are reported when discriminating only two surface types, much research has been conducted on detecting, assessing,
namely, saline and nonsaline (Evans & Caccetta, 2000). and mapping agriculture, forest, wetland, rangeland, and
Attempts have also been made to verify the efficiency of this urban features (Metternicht, 1999). However, relatively few
kind of remote sensing data in mapping salinity types and studies have investigated the possibility of using air- and
degrees (Kalra & Joshi, 1996; Metternicht, 1998; Metter- space-borne microwave data for mapping areas degraded by
nicht & Zinck, 1996). salinization. In general, microwave C-, P- and especially L-
In general, reflectance increases with increasing quantity bands are considered adequate for detecting salinity in
of salts at the terrain surface. This is particularly true for the different settings (Carver & Bush, 1979; Chaturvedi et al.,
blue band, in which the interference caused by ferric oxides 1983; Metternicht, 1998; Schmullius & Evans, 1997; Singh,
is masked. Salt-affected soils show relatively higher spectral Kumar, & Srivastav, 1990; Taylor et al., 1996).
response in the visible and near-infrared regions of the Compared to optical sensing, microwave sensing
spectrum than nonsaline soils do, and strongly saline-sodic presents advantages in special soil conditions, which turn
soils present higher spectral response than moderately saline- out to be in general also salt-affected (Taylor et al., 1996), as
sodic soils (Rao et al., 1995). Conversely, reflectance de- is the case of sandy coastal and desert zones, waterlogged
creases and salts are difficult to identify with increasing areas, and places with irregular microtopography such as
moisture, the presence of ferric oxides, and the inclusion of puffy crusts and cloddy surfaces (Metternicht, 1998; Singh
clay. Moreover, the presence of water and OH groups in the & Srivastav, 1990). Furthermore, the independence from
crystal lattice contributes to decreasing the reflectance in the atmospheric conditions offers an advantage over the optical
middle- and near-infrared bands (Epema, 1990; Mougenot, and infrared satellite-borne sensors that can be exploited for
1993). developing operational tools to support the mapping and
The use of multispectral images for detecting salts is management of salt-affected areas.
hampered by a set of factors. Among them, the quantity of Because of the differential behaviour of the real and
salts concentrated on the soil surface plays a primary role. It imaginary parts of the dielectric constant, microwaves are
seems to be that below 10 – 15% salt content, salts are efficient in detecting soil salinity. While the real part is
difficult to discriminate from other soil surface components independent of soil salinity and alkalinity, the imaginary
(Mougenot, 1993). Thus, ground validation is indispensable part is highly sensitive to variations in soil electrical con-
to correlate surface features, salt content, and reflectance. ductivity, but with no bearing on variations in alkalinity.
This allows the separation of saline soils from others
3.4. Satellite and airborne sensors covering the thermal (Sreenivas, Venkataratnam, & Rao, 1995). Previous studies
infrared have focused on the following features: (a) saline water
detection by analysing the dependence of microwave
The thermal infrared is commonly used to estimate responses on salinity and temperature (Singh et al., 1990;
moisture and salinity. This range of the electromagnetic Singh & Srivastav, 1990), (b) soil salinity identification by
spectrum registers features caused by energy absorption of relating salinity levels to the imaginary part of the complex
different salt types such as sulphates, phosphates, and dielectric constant (Bell, Menges, Ahmad, & Van Zyl, 2001;
chlorides (Mulders, 1987). Moderate to poor results were Carver & Bush, 1979; Sreenivas et al., 1995), and (c) soil
obtained when using the thermal range of the spectrum salinity mapping, including discrimination of salinity levels
alone (Metternicht, 1996). In case studies from Iran and by mapping surface roughness and vegetation types related
Egypt, the inclusion of thermal band 6 into selected (best) to salinity, and then using this information as ancillary data
Landsat TM band combinations improved the separation of to estimate the extent of salinity at regional level (Metter-
saline soils from gypsiferous and coarse-textured desert nicht, 1998; Taylor et al., 1996).
soils, and substantially increased both the overall and the
class accuracies (Goossens, Alavi Panah, De Dapper, & 3.6. Hyperspectral sensing
Kissyar, 1999; Goossens & Van Ranst, 1998).
In irrigation schemes, infrared thermometry is commonly The capabilities of hyperspectral imagery for salinity
used to detect salt-affected areas from the relationship mapping have been recently investigated by Ben-Dor,
between crop water stress and temperatures of the crop Patkin, Banin, and Karnieli (2002) and Taylor and Dehaan
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 11

(2000). Using the Hymap airborne sensor that acquires 3.8. Electromagnetic induction
images over the spectral range of 450 –2500 nm (i.e. visible,
near- and mid-infrared) in 128 bands, Taylor and Dehaan The electromagnetic induction meter (McNeill, 1986)
mapped saline areas characterised by salt scalds, halophytic estimates soil salinity by measuring the bulk electrical
vegetation, and soils with varying salinity degrees and conductivity of the soil, which depends on the salinity of
types. Using spectral unmixing, they found that saline the soil solution, porosity, and the type and amount of clay
endmembers presented a pronounced high reflectance at in the soil. By inducing an electromagnetic field, the instru-
800 nm, a shallow and wide hydroxyl feature at 2200 nm, ment is able to measure the electrical conductivity of the
and broad absorption features at 1450 and 1900 nm. Wet bulk soil (ECa), also referred to as the apparent soil salinity,
saline soils presented barely distinguishable hydroxyl fea- in a volume of soil below the transmitter and receiver coils
tures at 2200 nm. Spectral features in the visible and near- (Johnston, Savage, Moolman, & du Plessis, 1997).
infrared parts of the spectrum, related to combined water in The instrumentation has evolved from geophysical
hydrated evaporite minerals, allowed saline soils to be equipment, with a large number of electromagnetic con-
mapped by the Hymap scanner. Further tests showed that ductivity meters being manufactured by Geonics, Canada
the subtle hydrate absorption features at 980 and 1170 nm (e.g. EM31, EM34-3, EM38, EM39). Its value for rapid
are crucial for correct mapping of salt-affected soils. diagnosis and mapping of soil salinity has been recognised
Likewise, Ben-Dor et al. (2002) report successful results in works by George, Bennett, Arkell, and Vukelic (1994)
in using the hyperspectral DAIS-7915 sensor and the Visible and Mackenzie, Bellamy, Fraser, and Ellis (1989) in Aus-
and Near-Infrared Analysis (VNIRA) approach to produce tralia, Rhoades (1992) in USA, Cannon, McKenzie, and
quantitative soil surface maps of organic matter, soil field Lachapelle (1994) in Canada, and Johnston et al. (1997) in
moisture, and soil salinity. Africa. In the agricultural belt of Western Australia, conven-
tional EM salinity measurements have been carried out with
3.7. Airborne geophysics data the unit mounted on a nonmagnetic vehicle such as an all-
terrain vehicle, allowing to survey around 120– 150 ha per
Airborne geophysics offers new possibilities for detect- day, but this varies largely according to terrain conditions,
ing salt-affected areas. Measurements are made with instru- such as topography and land use (Metternicht, 1998b).
ments mounted on or towed by an aircraft. A variety of Correct interpretation of measurements presents a major
geophysical methods is available, including seismology, challenge. Some authors directly correlate the apparent
gravity, magnetics, and electrical, electromagnetic, and electrical conductivity (ECa) with salinity levels to deter-
gamma-ray spectrometry (George, Campbell, Woodgate, mine salinity hazards, while others convert the measure-
Farrell, & Taylor, 2000). For instance, the National Airborne ments into saturated paste extract equivalents (ECe) through
Geophysics project in Australia (George et al., 1998) regression equations. Different calibration models and cor-
analyzed the performance of airborne electromagnetic, mag- relation factors have been developed, which show that the
netic, and gamma-ray spectrometry techniques for rapidly accuracy of the measurements (e.g. from F 1 to F 7.59 dS/
obtaining data over large areas that can assist salinity m) and the correlation between the estimated and real depth
investigations. The study shows that surface salinity hazard of salinity occurrence are highly variable (Cannon et al.,
maps can be generated provided the remotely sensed data 1994; Johnston et al., 1997). Accuracy is affected by
are integrated with proper ground data (e.g. electromagnetic variations in soil texture and water content and, therefore,
ground survey), bore hole information (e.g. water table most investigations suggest establishing proper calibration
depth and salinity data), and Landsat TM images. between EM values and electrical conductivity (ECe) for the
In general, good correlation exists between the apparent specific soil and moisture conditions of a particular survey
electrical conductivity as measured by an electromagnetic area. Despite this accuracy issue, EM series have the
induction meter, mounted on an airplane, and electrical potential to collect nondestructive soil samples for rapid
conductivity from field samples (Hallett, 1994). According assessment, mapping, and monitoring soil salinity.
to airborne EM surveys in Western Australia (Street &
Duncan, 1992), distribution patterns and values of conduc-
tivity similar to those gathered by ground-based systems 4. Image processing of remotely sensed data
(e.g. chloride concentrations from drill hole data) were
obtained, confirming that apparent electromagnetic conduc- Raw data acquired from ground-, air-, or satellite-borne
tivity responded to variations in salt stored in the regolith. sensors are usually transformed to allow for better discrim-
However, a report of the Waters and Rivers Commission of ination between saline and nonsaline soils or among salinity
Western Australia found that the system was not able to classes. A variety of transforms of remote sensing data has
identify all salt hazard sites and that some areas identified as been used in soil salinity studies, including best band
having low conductivity were indeed highly saline, espe- selection, principal components analysis, the Kauth –Tho-
cially where a thin soil layer covered the rock substratum mas transform, intensity – hue – saturation transformation,
(Metternicht, 1998b). image ratioing, image differencing, and pattern recognition
12 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

techniques using maximum likelihood classifier, neural Csillag et al. (1993) used a modified stepwise principal
networks, decision trees, unmixing of surface features, components analysis to assess the effectiveness of individ-
fuzzy classification, and radar backscattering inversion ual bands for discriminating salinity states from high-reso-
techniques. lution spectra provided by narrow absorption bands (10 nm)
within the range of 495– 2395 nm. Working with a large
4.1. Selection of best band combination data set from California and Hungary, they achieved 80–
90% overall recognition accuracy.
Selection of the best band combination is usually the first
step to compress remote sensing data, while securing class 4.2.2. Intensity– hue –saturation transformation
separability. The optimum index factor (OIF) was applied to The intensity –hue – saturation (IHS) transformation, with
remote sensing data covering salt-affected areas in the Indo- subsequent supervised classification using per-pixel max-
Gangetic plain, and revealed that the best TM three-band imum likelihood classifier, was the approach adopted by
combination was 1 –3 – 5, but band combinations 3 –5 – 7 Dwivedi et al. (2001) for mapping salt-affected soils in the
and 3– 4– 5 could serve as well because their OIF values Indo-Gangetic alluvial plains of northern India. The IHS
were close to that of the first ranked. Recurrent bands 3 and transform was applied for merging high-spatial resolution
5 seemed to contain the most important part of the infor- panchromatic data of the IRS-1C (5.8 m) sensor with low-
mation. However, there was a lack of consistency between resolution multispectral data of the LISS-III (23.5 m) and
OIF values and field salinity estimates, as shown by band LISS-II (36.5 m) sensors. The IRS-1C and LISS-III data
combination 1 –2 – 7, which ranked last-but-one while giv- without any transformation ranked last in terms of overall
ing 96% interpretation accuracy (Dwivedi & Rao, 1992). accuracy. After transformation using the IHS approach,
Transformed divergence (TD) analysis was applied to accuracy figures for LISS-II, LISS-III and IRS-1C, and
evaluate the separability among salinity classes and select LISS-III hybrid data were 89.5%, 85.9% and 81.5%,
the best band combination to be used for the classification respectively.
of a Landsat TM image of a salt-affected area in Bolivia
(Metternicht & Zinck, 1996, 1997). Two band combina- 4.3. Unmixing of surface features
tions, comprising six (1, 2, 4, 5, 6, 7) and five (1, 2, 4, 6, 7)
bands, resulted with very high TD values of 1957 and One major drawback to the use of remote sensing data
1959, respectively, thus, close to the maximum possible of for salinity and alkalinity mapping is the intricate distribu-
2000. tion of salts on the terrain surface, causing frequent occur-
rence of heterogeneous pixels. The mixed surface
4.2. Image transformations components can be separated using linear mixture model-
ling from the reflectance of ‘pure’ pixels corresponding to
4.2.1. Band ratios and principal components analysis selected endmembers. A variety of techniques, including
Band ratios of visible to near-infrared and between minimum noise fraction, principal components analysis, and
infrared bands have proven to be better for identifying salts identification of average pure pixels representing a partic-
in soils and salt-stressed crops than individual bands (Craig, ular surface component from false colour composites, can
Shih, Boman, & Carter, 1998; Hick & Russell, 1990; Hick, be applied for identifying endmembers directly from the
Davies, & Steckis, 1984). imagery (Metternicht & Fermont, 1998; Taylor & Dehaan,
Theoretically, principal components analysis offers an 2000).
interesting approach to deal at the same time with salt The output of the unmixing analysis is represented in a
identification and change detection. The stable brightness proportion map for each of the considered surface compo-
of PC1 and the stable greenness of PC2 allows the separa- nents, one image representing the sum of all the abundances
tion of saline from nonsaline soils, while the differential and a root –mean –square (RMS) image. The RMS image
brightness in PC3 and the differential greenness in PC4 represents the error between the original mixed spectrum
accounts for the changes occurring in surface salinity. A and the best-fit spectrum computed from the resulting
case study in the Indo-Gangetic plain showed that PC1 was endmember abundances (Smith, Ustin, Adams, & Gillespie,
very effective in identifying salt-affected areas, but changes 1990). The error ranges generated by the spectral unmixing
taking place between two reference dates (1975 and 1992) allow to reclassify the image and improve its interpretability
could not be detected from PC3 and PC4 (Dwivedi, 1996; (Metternicht & Zinck, 1996, 1997).
Dwivedi & Sreenivas, 1998). This shortcoming is attributed According to Taylor and Dehaan (2000), who derived
to the fact that PCA uses the spectral response pattern of the salinity maps from hyperspectral data, image endmembers
entire scene. It can be removed using the Kauth – Thomas can efficiently illustrate the distribution of salinity, though
transform, which deals separately with soil and vegetation they are likely to have different compositions within multi-
reflectances to generate individual brightness and greenness temporal data. They should be combined with maps of field-
images. The higher the brightness, the lower the humidity derived endmembers to be useful for monitoring changes in
and the higher the salinity (Peng, 1998). salinity and implementing remedial measures.
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 13

Fig. 7. Framework for fuzzy modelling of saline and alkaline classes, and approach to fuzzy classification of remotely sensed data (source: Metternicht, 1998).
14 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

4.4. Fuzzy classification 4.6. Radar backscatter inversion techniques

Successful discrimination of saline and alkaline areas In addition to fuzzy supervised classification of SAR
from remotely sensed data requires correct determination of satellite images (Metternicht, 1998), radar backscattering
information classes. The traditional approach to soil salinity inversion techniques have been applied for characterisation
mapping is based on rigid information classes with crisp of saline areas using airborne radar imagery (Taylor et al.,
boundaries, using soil reaction (pH) and electrical conduc- 1996). They found a good agreement between anomalous
tivity ranges as established, for instance, by the US Salinity complex dielectric constants determined by inversion of the
Staff Laboratory (Richards, 1954). However, in nature, salt polarised returns of AIRSAR images acquired in wet con-
contents vary in gradual manner, horizontally as well as ditions and saline areas as defined by electrical conductivity
vertically, even when conspicuous white patches of salt and dielectric constants determined in the field.
crust appear at the soil surface. As a consequence, broad
zones of gradual transition may be misrepresented because
of the arbitrary assignment of sharp class boundaries. Fuzzy 5. Assessing temporal and spatial changes of salinity
classification, based on the fuzzy set theory proposed by
Zadeh (1965), where a given class has a continuum of Monitoring salinity changes from past to present faces
membership grades, can cope with the diffuse spatial dis- the difficulty that, in general, there is no ground-truth
tribution of salts and sodium, better than crisp classification. information available for past situations. Thus, validating
Expanding the concepts highlighted by Burrough (1989) historical remote sensing data involves uncertainties. To
for soil classification to the issue of soil salinity, Metternicht overcome this difficulty, fusion of multisource remote
(1998) proposes the determination of transitional ‘fuzzy’ sensing data and their integration with field and laboratory
class boundaries, derived from continuous salinity classes data have been advocated.
that intergrade gradually, to better represent real-world
situations. Fig. 7 illustrates the approach devised for deter- 5.1. Issues in soil salinity monitoring
mining fuzzy classes and the fuzzy classification applied to
a JERS-1 SAR data image from Cochabamba, Bolivia. This Moving from simple spatial identification of salt-affected
method provided a reliable detection of salt-affected areas, areas to the monitoring of temporal changes in salinity
with an overall accuracy of 81% (Metternicht, 1998). In all requires at least two identification dates along a given time
cases, fuzzy classification improved, or at least equalled, the span. As salt-related surface features drastically change with
crisp classification. One main drawback was that surface seasons, time series of remote sensing data must be captured
roughness produced by the cultivation of slightly alkaline in similar periods of the year to be comparable, preferably at
land affected the radar backscattering, causing erroneous the end of the dry season, if passive remote sensors are used.
allocations of alkaline and saline – alkaline soils to non- Georeferencing and co-registration of multitemporal data
affected areas. are essential to enable ground sites to be traced over time
and satellite data to be compared with ancillary data.
4.5. Decision trees and neural networks Approaches, such as those developed by the Land Monitor
project in Western Australia, require calibration between
The performance of maximum likelihood classifiers, images to ‘like-values’ so that digital numbers from differ-
nonparametric decision trees, neural networks, and condi- ent dates can be compared (Campbell, Furby, & Fergusson,
tional probabilitistic networks for discriminating saline areas 1994). In addition to that, ground-truth data must be
from nonaffected ones has been tested using Landsat TM collected contemporaneously to validate the spectral signa-
data, landform data derived from a digital elevation model, tures from the satellite images. These include radiometric
and conditional statements about salinity – landscape rela- field measurements, surface features description and map-
tionships (Evans, 1998). Conditional probabilitistic net- ping, as well as soil sampling for laboratory analyses.
works, enabling the inclusion of prior knowledge about
the relationships between input attributes and salinity in the 5.2. Uncertainties in soil salinity monitoring
classification of Landsat TM data, showed the best techni-
que for salinity mapping in the southwest of Western Errors can arise from inaccurate image registration, poor
Australia (Evans, 1998). This research highlights one sig- pixel identification and matching, and feature misclassifica-
nificant disadvantage of decision trees and neural networks, tion. There are also error sources when the two images do
which is their inability to incorporate prior knowledge about not represent similar field conditions at the selected dates.
the relationships between input attributes and the relation- This creates uncertainties concerning the likelihood, nature,
ships between these input attributes and the output classes. and magnitude of the changes.
This might significantly affect the accuracy of mapping, To cope with these uncertainties, expert knowledge can
leading to over- or underestimate the extent of salt-affected be mobilized and formalized as ‘if – then’ rules in a mon-
areas. itoring model, as discussed by Metternicht (2001). In
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 15

particular, expert judgement allows us to assess whether tion can be the basis for change prediction. Such an attempt
given changes, as detected from the difference image, might of change prediction has been done for the Cochabamba
really occur or are rather artifacts resulting from inaccurate region in Bolivia. Areas which have undergone significant
map overlay. For example, increases of two degrees in changes in the recent past are considered as particularly
salinity or changes from full saline to full alkaline are prone to potential increase in salinity and present, therefore,
unlikely to take place in a short time interval. This infor- a severe hazard of salinization (Metternicht, 1996, 2001).
mation is incorporated in the model as degrees of likelihood The degree of probability for such an evolution to occur can
(Fig. 8). The expert grades the likelihood of occurrence in be attached to the mapped areas, as was done for a region in
terms of certainty factors within a range of 0 – 1, from Russia (Krapilskaya & Sadov, 1987).
absolutely unlikely to very likely. The values so obtained Other methods for predicting areas at risk of dryland
are further used to generate a map of likelihood of changes salinity use a combination of multitemporal satellite
between two selected reference dates. Changes unlikely to imagery, information derived from large-scale DEMs, and
occur are corrected to validate two difference maps. One subjective probability (Evans & Caccetta, 2000; Florinsky,
shows the nature of the changes: for example, from non- Eiler, & Lelyk, 2000). Evans and Caccetta (2000) used
affected to saline or from saline to saline– alkaline. The decision trees based on the relationships between salinity
other shows the magnitude of the changes: for example, one risk and variables describing landscape features to repro-
class in alkalinity or two degrees in salinity. duce expert opinions about the future extent of salinity in a
certain area, whereas Florinsky et al. (2000) assumed that
5.3. Salinity –alkalinity hazard prediction the build-up of salts at macrotopographic scale occurs in
depressions.
The monitoring procedure based on fuzzy logic (Metter-
nicht, 2001) highlights areas where major changes in salinity
and/or alkalinity have taken place between the current 6. Data fusion and data integration
situation and some date in the past. This retrospective
monitoring shows trends, which can be extended into the Current mapping and prediction approaches can benefit,
future, within given possibility ranges. Thus, change detec- in terms of improved salinity mapping and prediction, from

Fig. 8. Graphic representation of the likelihood of changes. Numbers in the matrix boxes indicate information classes. Arrows show the likelihood or certainty
of change (source: Metternicht, 2001).
16 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

the synergistic use of airborne and satellite-borne sensors A first step of improvement consists in incorporating
covering different regions of the spectrum, and or the contextual landscape information. For example, in the
integration of soilscape information into the modelling Cochabamba region, Bolivia, highly saline– alkaline areas
process, as described hereafter. are strongly correlated with specific geomorphic positions,
in particular with playas and lagunary flats, while non-
6.1. Spectral confusions and spectral efficiency (data affected areas are mainly on piedmont glacis. The inclusion
fusion) of this relationship in spatial analysis and modelling led to
better spatial discrimination of salinity (Metternicht &
In salt-affected areas, spectral confusions between differ- Zinck, 1996, 1997). However, by doing so, we are still
ent surface features are common. For example, a white looking at salt-affected areas as mere surfaces, as 2D spaces.
saline crust and a shiny silty crust have both similar high This is one dimension short, as soils are 3D bodies and
reflectance values. This hampers the spectral separability of salinity can substantially change with depth, even more that
salt-affected areas. In an attempt to evaluate the efficiency of surface salinity is frequently controlled by subsoil or sub-
the different regions of the electromagnetic spectrum in stratum salinity and by the seasonal fluctuations of the water
allowing for good spectral separability, Metternicht (1996) table.
obtained the following results for the Cochabamba area in According to McGowen and Mallyon (1996), highest
Bolivia. accuracy in detecting dryland salinity in New South Wales,
Australia, was obtained when using multitemporal Landsat
(a) The visible range performed well in separating non- TM imagery and ground salinity data measured with an
affected areas from saline and saline – alkaline ones. EM31 instrument providing information on salinity at sub-
(b) The infrared range provided the lowest spectral soil level, and geo-referenced with a GPS. The different data
confusion when separating saline– alkaline areas from sets were integrated within a GIS. A similar approach was
pure alkaline areas. implemented by Mackenzie et al. (1989) to map dryland
(c) The thermal range improved the separability of the salinity recharge areas from the combination of EM34
alkaline areas because alkalinity is controlled by the measurements, vegetation information, infiltration rates,
amount of free carbonates, which have a strong and airborne scanner imagery.
absorption feature in TM band 6. Another effective approach for mapping salt-affected
(d) Overall, the microwave region (L-band) had low soils that combines digital Landsat TM image and ancillary
amounts of spectral confusion and the best separability data such as topographic maps, geologic maps, soil maps,
between saline and alkaline areas, nonaffected, and land use maps, ground water quality, and historical data on
saline –alkaline areas, as well as among saline and salinity was developed by Sah, Apisit, Murai, and Parkpian
saline –alkaline areas. (1995). GIS-assisted data integration was found effective in
classifying low-salinity and potential saline areas, as well as
Thus, the different regions of the electromagnetic spec- correcting areas misclassified as extremely or moderately
trum have complementary capabilities for spectral separa- saline. Likewise, Eklund, Kirkby, and Salim (1998) used
bility of the salinity –alkalinity classes. This calls for data GIS modelling for monitoring salinization, by isolating
fusion to increase the efficiency of the remote sensing data. proper indicators and determining their ‘weight’. They
The research of Metternicht showed that the merging of found certain attributes to be more significant than others
visible and infrared with microwave data achieved a better for discriminating salinity classes. Ground-water depth, in
discrimination of the alkaline and nonaffected areas than particular, was a prime discriminator of saline discharge.
Landsat TM or JERS-1 SAR alone, with 35 –40% increase Yet, one very effective monitoring approach consists in
in accuracy (Metternicht, 1996). comparing historical and current soil salinity maps, when
historical field and laboratory data are available. This is the
6.2. Landscape and soilscape information (data integra- case for a region of the Golestan province in northeast Iran
tion) (Naseri, 1998). Soil salinity/alkalinity maps of several
irrigation schemes were available for 1972, providing elec-
It is not enough to separate classes of salinity and trical conductivity and pH data from composite samples
alkalinity. The reclamation, rehabilitation, and management taken at three depths (0– 50, 50 –100 and >100 cm) on a
of salt-affected areas require that not only classes but also regular grid of 500-m intervals. The same points were
severity levels of salinity –alkalinity be determined. This visited again in 1995 and sampled for electrical conductivity
cannot be achieved from remote sensing data alone and and pH determinations. The values of the two dates were
needs full synergy between remote sensing, field, and compared and the differences expressed as discrete map
laboratory data (De Dapper, Goossens, Gad, & El Badawi, units, in terms of percent salinity/alkalinity increase or
1996; Khalil, Fahim, & Hawela, 1995; Metternicht & Zinck, decrease and in terms of class changes. This method
1997; Mongkolsawat, Thirangoon, & Eiumnoh, 1991; Peng, requires accurate interpolation techniques to generate con-
1998; Younes, Gad, & Rahman, 1993). tinuous salinity maps from field point observations and the
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 17

design of proper sampling strategies to cater for the spatial reclamation, and rehabilitation. Monitoring soil salinity is
variations proper of salt-affected soils, as shown by Utset, greatly enhanced when using multispectral remote sensing
Ruiz, Herrera, and Deleon (1998). This kind of information, data, which however need transformation before being used
resulting from monitoring along the third dimension of the for identification and change detection. Detecting and
soil cover as a volume, lies largely beyond the current assessing changes between a current and a past situation
technical capabilities of satellite-based remote sensing. Air- involve uncertainties about the likelihood, nature, and
borne geophysics (e.g. EM surveys) show some promise in magnitude of the changes, which can be solved using expert
measuring soil salinity in the third dimension (e.g. depth), as knowledge.
shown by studies conducted in Australia by George et al. The best monitoring results are obtained when integrat-
(1998) and Street and Duncan (1992). ing remote sensing data with field and laboratory data. In
It is, thus, obvious that monitoring soil salinity and early this regard, geographic information systems offer the
warning of salinization cannot be achieved from remote advantage of integrating data of diverse nature in terms of
sensing data alone and require a solid synergy between scale, time, source, and structure. It is the researcher’s
remote sensing data, field observations, and laboratory challenge to identify the most adequate salinity indicators
determinations, as sources of data, and GIS facilities for for a particular area, so that appropriate ground and remote
processing, transforming, and displaying the data (Tóth, sensing techniques can be applied to extract information in
Kertesz, & Pásztor, 1998). an accurate and cost-effective manner.

7. Conclusion References

By providing fast, timely, relatively cheap, and repetitive Bell, D., Menges, C., Ahmad, W., & van Zyl, J. (2001). The application of
data, remote sensing plays an important role for detecting, dielectric retrieval algorithms for mapping soil salinity in a tropical
coastal environment using airborne polarimetric SAR. Remote Sensing
mapping, and monitoring salt-affected surface features. of Environment, 75, 375 – 384.
Ground validation remains necessary to determine relation- Ben-Dor, E., Irons, J., & Epema, G. (1999). Remote sensing for the earth
ships between reflectance and salt type and degree in soil science, manual of remote sensing (3rd ed.) (pp. 111 – 188). New York:
and groundwater, deduced from surface characteristics Wiley.
directly on bare soils or indirectly on vegetation. Likewise, Ben-Dor, E., Patkin, K., Banin, A., & Karnieli, A. (2002). Mapping of
several soil properties using DAIS-7915 hyperspectral scanner data.
sensors with capabilities to penetrate the soil surface (e.g. A case study over soils in Israel. International Journal of Remote Sens-
airborne electromagnetic, magnetic, and gamma-ray spec- ing, 23, 1043 – 1062.
trometry, or ground-based electromagnetic induction Burrough, P. (1989). Fuzzy mathematical methods for soil survey and land
meters) require integration of field data (e.g. water table evaluation. Journal Soil Science, 40, 477 – 492.
depth and salinity data) and data from other regions of the Campbell, N., Furby, S., & Fergusson, B. (1994). Calibrating different
images from different dates. Report to LWRRDC, Project CDM1. Perth,
spectrum to provide reliable information on the distribution WA: CSIRO Division of Mathematics and Statistics.
of salt-affected areas. Cannon, M., McKenzie, R., & Lachapelle, G. (1994). Soil salinity mapping
As earlier summarized by Mougenot et al. (1993), direct with electromagnetic induction and satellite-based navigation methods.
and precise estimation of salt quantities from satellite remote Canadian Journal of Soil Science, 74, 335 – 343.
sensing is rather difficult because of: Carver, K., & Bush, T. (1979). Airborne multispectral remote sensing of
saline seeps: the 1978 Harding Co., South Dakota experiment. NASA
Contract NAS9-15421, Final Report. Las Cruces, NM: New Mexico
(a) Lack of specific absorption bands of some salt types State University.
(e.g. halite) and occurrence of spectral confusions. Chaturvedi, L., Carver, K., Clifford Harlan, J., Hancock, G., Small, F., &
(b) Low spectral resolution of satellite bands. Dalstead, K. (1983). Multispectral remote sensing of saline seeps. IEEE
Transactions on Geoscience and Remote Sensing, 21, 231 – 239.
(c) Vertical, spatial, and temporal variability of salinity in
Craig, J. C., Shih, S. F., Boman, B. J., & Carter, G. A. (1998). Detection of
the soil mantle and substratum. salinity stress in citrus trees using narrow-band multispectral imaging.
(d) The presence of other soil chromophores1 (e.g. structural ASAE paper no. 983076. ASAE Annual International Meeting, Orlan-
crusts, soil surface texture, soil colour, soil surface do, FL, USA, 12 – 16 July 1998. 10 pp.
roughness as determined by different salt types) as Csillag, F., Pásztor, L., & Biehl, L. (1993). Spectral band selection for the
discussed in Section 2 of this paper. characterization of salinity status of soils. Remote Sensing of Environ-
ment, 43, 231 – 242.
Dale, P., Hulsman, K., & Chandica, A. (1986). Classification of reflectance
Soil salinization represents an increasing environmental on colour infrared aerial photographs and sub-tropical salt marsh veg-
hazard, especially in irrigated areas. This calls for monitor- etation types. International Journal of Remote Sensing, 7, 1783 – 1788.
ing to support timely decision making on land management, De Dapper, M., Goossens, R., Gad, A., & El Badawi, M. (1996). Model-
ling and monitoring of soil salinity and waterlogging hazards in the
desert-delta fringes of Egypt, based on geomorphology, remote sensing
1
The term chromophore has been used to describe parameters and and GIS. In O. Slaymaker (Ed.), Geomorphic hazards ( pp. 169 – 182).
substances (chemical or physical) that significantly affect the shape and Chichester, UK: Wiley.
nature of a soil spectrum (Ben-Dor, Irons, & Epema, 1999). De Jong, S. (1992). The analysis of spectroscopical data to map soil types
18 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

and soil crusts of Mediterranean eroded soils. Soil Technology, 5, George, R., Bennett, D., Arkell, P., & Vukelic, B. (1994). Soil salinity in
199 – 211. the southwest irrigation area; extent and management options. Proceed-
De Jong, S., & Epema, G. (2001). Imaging spectrometry for surveying and ings of the Soils ’94 Conference ( pp. 101 – 107). Broadwater Resort,
modelling land degradation. In F. D. van der Meer, & S. de Jong (Eds.), Busselton: Australian Society of Soil Science.
Imaging spectrometry ( pp. 65 – 86). Dordrecht, The Netherlands: George, R., Campbell, C., Woodgate, P., Farrell, S., & Taylor, P. (2000).
Kluwer Academic Publishers. Complementary data and cost benefit analysis of utilising airborne
Delfos, J. (2001). Land degradation in the agricultural belt of Western geophysics for salinity management purposes. Draft report, Project
Australia. Project report. Perth, WA: Department of Spatial Sciences, no. RM17. Land and Water Resources Research and Development.
Curtin University of Technology. 43 pp.
Driessen, P. M., & Schoorl, R. (1973). Mineralogy and morphology of salt Ghassemi, F., Jakeman, A. J., & Nix, H. A. (1995). Salinisation of land and
efflorescences on saline soils in the Great Konya Basin, Turkey. Journal water resources: human causes, extent, management and case studies.
of Soil Science, 24, 436 – 442. Canberra, Australia: The Australian National University, Wallingford,
Drysdale, G. (2001). Exploring the relationship of Digital Multispectral Oxon, UK: CAB International.
Imagery with agricultural crops and soil properties. Honors Disserta- Goldshleger, N., Ben-Dor, E., Benyamini, Y., Agassi, M., & Blumber, D.
tion, Department of Spatial Sciences, Curtin University of Technology, (2001). Characterization of soil’s structural crust by spectral reflectance
Perth, WA. 150 pp. in the SWIR region(1.2 – 2.5 Am). Terra Nova, 13, 12 – 17.
Dwivedi, R., Ramana, K., Thammappa, S., & Singh, A. (2001). The Golovina, N. N., Minskiy, D., Pankova, Y., & Solovyev, D. A. (1992).
utility of IRS-1C, LISS-III and PAN-merged data for mapping salt- Automated air photo interpretation in the mapping of soil salinization
affected soils. Photogrammetric Engineering and Remote Sensing, 67, in cotton-growing zones. Mapping Sciences and Remote Sensing, 29,
1167 – 1175. 262 – 268.
Dwivedi, R. S. (1996). Monitoring of salt-affected soils of the Indo-Gang- Goossens, R., Alavi Panah, S. K., De Dapper, M., & Kissyar, O. (1999).
etic alluvial plains using principal component analysis. International The use of thermal band of Landsat TM for the study of soil salinity in
Journal of Remote Sensing, 17, 1907 – 1914. Iran (Ardakan area) and Egypt (Ismailia Province). Proceedings Inter-
Dwivedi, R. S., & Rao, B. R. M. (1992). The selection of the best possible national Conference on Geoinformatics for Natural Resource Assess-
Landsat TM band combination for delineating salt-affected soils. Inter- ment, Monitoring and Management ( pp. 454 – 459). Dehradun, India:
national Journal of Remote Sensing, 13, 2051 – 2058. Indian Institute of Remote Sensing, NRSA.
Dwivedi, R. S., & Sreenivas, K. (1998). Image transforms as a tool for the Goossens, R., & Van Ranst, E. (1998). The use of remote sensing to
study of soil salinity and alkalinity dynamics. International Journal of map gypsiferous soils in the Ismailia Province (Egypt). Geoderma,
Remote Sensing, 19, 605 – 619. 87, 47 – 56.
Eghbalm, M. K., Southard, J., & Whittig, L. D. (1989). Dynamics of Hallett, M. (1994). Groundwater and salinity studies using airborne geo-
evaporite distribution in soils on a fan – playa transect in the Car- physics. Australian Journal of Soil and Water Conservation, 7, 24 – 27.
rizo Plain California. Soil Science Society of America Journal, 53, Hick, P. T., Davies, J. R., & Steckis, R. A. (1984). Mapping dryland salinity
898 – 903. in Western Australia using remotely sensed data. Satellite remote sens-
Eklund, P., Kirkby, S., & Salim, A. (1998). Data mining and soil salinity ing: review and preview ( pp. 343 – 350). Reading, UK: Remote Sensing
analysis. International Journal of Geographical Information Science, Society.
12, 247 – 268. Hick, P. T., & Russell, W. G. R. (1990). Some spectral considerations for
Epema, G. F. (1990). Effect of moisture content on spectral reflectance in remote sensing of soil salinity. Australian Journal of Soil Research, 28,
a playa area in southern Tunisia. Proceedings International Symposi- 417 – 431.
um, Remote Sensing and Water Resources, Enschede, The Netherlands Hovari, F., 2002. Spectral library of surface soil salinity. Available at http://
( pp. 301 – 308). International Institute for Aerospace Survey and Earth www2.utep.edu/~fhowari/spectral/library.htm.
Sciences. Hunt, G., Salisbury, J., & Lenhoff, C. (1972). Visible and near infrared
Epema, G. F. (1992). Mapping surface characteristics and their dynam- spectra of minerals and rocks: V. Halides, phosphates, arsenates, ven-
ics in a desert area in southern Tunisia with Landsat Thematic Map- adates and borates. Modern Geology, 3, 121 – 132.
per. In G. F. Epema, 1992. Spectral reflectance in the Tunisian Islam, Z., & Metternicht, G. (2002). Fuzzy classification approach to detect
desert. PhD thesis, Wageningen Agricultural University, the Nether- tree crowns and species from a forested area using DMSV data. Pro-
lands, pp. 123 – 142. ceedings of the 29th International Symposium on Remote Sensing of
Escadafal, R. (1989). Caractérisation de la surface des sols arides par Environment, Paper 9.17, Buenos Aires, Argentina, April. Environmen-
observation de terrain et par télédétection. Paris, France: Editions de tal Research Institute of Michigan.
l’ORSTOM. Jain, S., Sharma, K., Manchanda, M., & Singh, C. (1988). Physiographic
Escobar, D. E., Gerbermann, A. H., & Alaniz, M. A. (1988). Detecting soil classification from multi remote sensing data. Proceedings of the
saline soils with video imagery. Photogrammetric Engineering and 9th Asian Conference on Remote Sensing, No. 23 – 29, Bangkok,
Remote Sensing, 54, 1283 – 1287. Thailand, vol. 17 (pp. 1 – 8). Asian Centre for Research on Remote
Evans, F. (1998). An investigation into the use of maximum likelihood Sensing.
classifiers, decision trees, neural networks and conditional probabili- Johnston, M., Savage, M., Moolman, J., & du Plessis, H. (1997). Evalua-
tistic networks for mapping and predicting salinity. MSc thesis, School tion of calibration methods for interpreting soil salinity from electro-
of Computer Science, Curtin University of Technology, Perth, WA. magnetic induction measurements. Soil Science Society of America
Evans, F., & Caccetta, P. (2000). Broad-scale spatial prediction of areas at Journal, 61, 1627 – 1633.
risk from dryland salinity. Cartography, 29, 33 – 40. Kalra, N. K., & Joshi, D. C. (1996). Potentiality of Landsat, SPOT and IRS
Everitt, J., Escobar, D., Gerbermann, A., & Alaniz, M. (1988). Detecting satellite imagery for recognition of salt affected soils in Indian arid
saline soils with video imagery. Photogrammetric Engineering and zone. International Journal of Remote Sensing, 17, 3001 – 3014.
Remote Sensing, 54, 1283 – 1287. Khalil, K. A., Fahim, M., & Hawela, F. (1995). Soil reflectances as
Florinsky, I., Eiler, R., & Lelyk, G. (2000). Prediction of soil salinity risk affected by some soil parameters. Egyptian Journal of Soil Science,
by digital terrain modelling in the Canadian prairies. Canadian Journal 35, 477 – 491.
of Soil Science, 80, 455 – 463. Krapilskaya, N. M., & Sadov, A. V. (1987). Aerospace monitoring of
George, R., Beasley, R., Gordon, I., Heislers, D., Speed, R., Brodie, R., hydrogeological conditions and reclamation status of lands in southern
McConnell, C., & Woodgate, P. (1998). National Airborne Geophysics Aral coastal strip. Problems of Desert Development, 1, 22 – 26.
Project. Final Report, AFFA, NDSP. Lane, M. D., & Christensen, P. R. (1997). Thermal infrared emission spec-
G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (20032002) 1–20 19

troscopy of anhydrous carbonates. Journal of Geophysics Research, Mulders, M. (1987). Remote sensing in soil science. development in soil
102, 25581 – 25592. science. Amsterdam, The Netherlands: Elsevier, 379 pp.
Lane, M. D., & Christensen, P. R. (1998). Thermal infrared emission spec- Naseri, M. Y. (1998). Characterization of salt-affected soils for modelling
troscopy of salt minerals predicted for Mars. Icarus, 135, 528 – 536. sustainable land management in semi-arid environment: a case study
Mackenzie, M., Bellamy, G., Fraser, D., & Ellis, G. (1989). Mapping dry- in the Gorgan region, Northeast Iran. PhD thesis, Ghent University,
land salinity recharge using EM conductivity measurements and air- Belgium.
borne scanner imagery. Proceedings of the 5th Australian Soil National Land and Water Resources Audit (2000). Australian dryland sal-
Conservation Conference, vol 1 (pp. 41 – 48). Perth, WA: Hamilton, inity assessment 2000: extent, impacts, processes, monitoring and man-
Howes and Attwater. agement options (p. 129). Canberra, Australia: National Land and Water
Manchanda, M., & Iyer, H. (1983). Use of Landsat imagery and aerial Resources Audit.
photographs for delineation and categorization of salt-affected soils of Peng, W. L. (1998). Synthetic analysis for extracting information on soil
part of north-west India. Journal Indian Society Soil Science, 31, salinity using remote sensing and GIS: a case study of Yanggao Basin in
263 – 271. China. Environmental Management, 22, 153 – 159.
McGowen, I., & Mallyon, S. (1996). Detection of Dryland salinity using Prathapar, S. A., Smith, R. C. G., Barrs, H. D., Slavich, P. G., Humphreys, W.
single and multi-temporal Landsat imagery. Proceedings of the 8th A., Muirhead, & Lelij, A. van der (1990). Use of canopy temperatures
Australasian Remote Sensing Conference, Canberra ( pp. 26 – 34). to study soil spatial variability. Proceedings of Symposium on Manage-
McNeill, J. (1986). Rapid, accurate mapping of soil salinity using electro- ment of Soil Salinity in South East Australia, Albury, New South Wales,
magnetic ground conductivity meters. Technical Note TN-18. Missis- Australia, 18 – 20 September 1989 ( pp. 367 – 368). Auscript.
sauga, Ontario: Geonics, 15 pp. Rao, B., Sankar, T., Dwivedi, R., Thammappa, S., Venkataratnam, L.,
Metternicht, G. (1998a). Analysing the relationship between ground based Sharma, R., & Das, S. (1995). Spectral behaviour of salt-affected soils.
reflectance and environmental indicators of salinity processes in the International Journal of Remote Sensing, 16, 2125 – 2136.
Cochabamba Valleys (Bolivia). International Journal of Ecology and Rao, B. R. M., & Venkataratnam, L. (1991). Monitoring of salt-affected
Environmental Sciences, 24, 359 – 370. soils—a case study using aerial photographs, Salyut-7 space photo-
Metternicht, G. (1998b). Environmental indicators to map soil salinity graphs, and Landsat TM data. Geocarto International, 6, 5 – 11.
distribution in the Intermediate Rainfall Zone. a review. Report for Rhoades, J. (1992). Instrumental field methods of salinity appraisal. In G.
Alcoa of Australia, Perth, WA, 27 pp. Topp, et al. (Eds.), Advances in measurement of soil physical proper-
Metternicht, G. (1999). Current status and future prospectives of radar ties, bringing theory into practice. SSSA Special Publication, vol. 30
remote sensing for cartographic applications. Cartography, 28, 1 – 16. (pp. 231 – 248). Madison, WI: SSSA.
Metternicht, G. (2001). Assessing temporal and spatial changes of salinity Richards, L. (1954). United States salinity laboratory staff, diagnosis and
using fuzzy logic, remote sensing and GIS. Foundations of an expert improvement of saline and alkali soils. In L. Richards (Ed.), Agriculture
system. Ecological Modelling, 144, 163 – 177. handbook, 60. USA: US Department of Agriculture. 160 pp.
Metternicht, G., & Fermont, A. (1998). Estimating erosion surface fea- Sah, A., Apisit, E., Murai, S., & Parkpian, P. (1995). Mapping of salt-affected
tures by linear mixture modelling. Remote Sensing of Environment, soils using remote sensing and geographic information systems: a case
64, 254 – 265. study of Nakhon Ratchasima, Thailand. Proceedings of the 16th Asian
Metternicht, G., Hageman, E., Beeston, G., & Honey, F. (2001). Mapping Conference Remote Sensing, 20 – 24 November ( pp. G-3-1 – G-3-6).
variations in crop conditions using airborne videography. Proceedings Thailand: Nokhon Ratchasima.
of the 20th International Cartographic Conference, Beijing, China, Schmullius, C., & Evans, L. (1997). Table Summary of SIR-C/X-SAR
August ( pp. 732 – 743). International Cartographic Association. results: SAR frequency and polarization requirements for applications
Metternicht, G., Gonzalez, S., Beeston, G., & Honey, F. (2002). Applica- in ecology and hydrology. Proceedings of the IGARSS’97, Singapore
tions of high-resolution digital multispectral imagery for rapid assess- ( pp. 1734 – 1736). Institute of Electrical and Electronics Engineers.
ment of vegetation conditions in agricultural landscapes: the Australian Siegal, B., & Gillespie (1980). Remote sensing in geology. New York: Wiley.
experience. Proceedings of the 29th International Symposium on Re- Singh, R., Kumar, V., & Srivastav, S. (1990). Use of microwave remote
mote Sensing of Environment, Paper 3.18, Buenos Aires, Argentina, sensing in salinity estimation. International Journal of Remote Sensing,
April. Environmental Research Institute of Michigan. 11, 321 – 330.
Metternicht, G. I. (1996). Detecting and monitoring land degradation fea- Singh, R. P., & Srivastav, S. K. (1990). Mapping of waterlogged and salt-
tures and processes in the Cochabamba valleys, Bolivia: a synergistic affected soils using microwave radiometers. International Journal of
approach. ITC Publication, vol. 36. Enschede, The Netherlands: Inter- Remote Sensing, 11, 1879 – 1887.
national Institute for Aerospace Survey and Earth Sciences. (390 pp). Smith, D., Ustin, S., Adams, J., & Gillespie, A. (1990). Vegetation in
Metternicht, G. I. (1998). Fuzzy classification of JERS-1 SAR data: an deserts: I. A regional measure of abundance from multispectral images.
evaluation of its performance for soil salinity mapping. Ecological Remote Sensing of Environment, 31, 1 – 26.
Modelling, 111, 61 – 74. Sreenivas, K., Venkataratnam, L., & Rao, P. V. N. (1995). Dielectric prop-
Metternicht, G. I., & Zinck, J. A. (1996). Modelling salinity – alkalinity erties of salt-affected soils. International Journal of Remote Sensing,
classes for mapping salt-affected topsoils in the semi-arid valleys of 16, 641 – 649.
Cochabamba (Bolivia). ITC Journal, 1996-2, 125 – 135. Srivastava, A., Tripathi, N., & Gokhale, K. (1997). Mapping groundwater
Metternicht, G. I., & Zinck, J. A. (1997). Spatial discrimination of salt- and salinity using IRS-1B LISS II data and GIS techniques. International
sodium-affected soil surfaces. International Journal of Remote Sensing, Journal of Remote Sensing, 18, 2853 – 2862.
18, 2571 – 2586. Stovold, R., Steber, M., & Smith, P. (1994). Evaluation of low-cost digital
Mongkolsawat, C., Thirangoon, P., & Eiumnoh, A. (1991). A practical partially corrected multi-spectral video system data for operational use
application of remote sensing and GIS for soil salinity potential map- in Western Australia. Proc. of the 7th Australasian Remote Sensing
ping in Korat basin, northeast Thailand. Proceedings International Conference, Melbourne ( pp. 119 – 127). International Soil Conservation
Workshop on Conservation and Sustainable Development, 22 – 26 Organisation.
April 1991 ( pp. 290 – 297). Bangkok, Thailand: Asian Institute of Street, G., & Duncan, A. (1992). The application of airborne geophys-
Technology. ical surveys for land management. Proc. of the 7th International Soil
Mougenot, B. (1993). Effets des sels sur la réflectance et télédétection des Conservation Organisation, Sydney, 27 – 30 September 1992, vol. 2
sols salés. Cahiers ORSTOM, Série Pédologie, 28, 45 – 54. (pp. 762 – 770).
Mougenot, B., Pouget, M., & Epema, G. (1993). Remote sensing of salt- Szabo, J., Pásztor, L., Suba, Z., & Varallyay, G. (1998). Integration of
affected soils. Remote Sensing Reviews, 7, 241 – 259. remote sensing and GIS techniques in land degradation mapping. Pro-
20 G.I. Metternicht, J.A. Zinck / Remote Sensing of Environment 85 (2003) 1–20

ceedings 16th International Congress of Soil Science, Montpellier, Vegetation indices in crop assessments. Remote Sensing of Environ-
France, 20 – 26 August. International Society of Soil Science. 6 pp. ment, 35, 105 – 119.
Taylor, G., & Dehaan, R. (2000). Salinity mapping with hyperspectral Wiegand, C. L., Everitt, J. H., & Richardson, A. J. (1992). Comparison of
imagery. Available at www.geology.unsw.edu.au/research/R~Sensing. multispectral video and SPOT-1 HRVobservations for cotton affected by
Taylor, G. R., Mah, A. H., Kruse, F. A., Kierein-Young, K. S., Hewson, R. D., soil salinity. International Journal of Remote Sensing, 13, 1511 – 1525.
& Bennett, B. A. (1996). Characterization of saline soils using airborne Wiegand, C. L., Rhoades, J. D., Escobar, D. E., & Everitt, J. H. (1994).
radar imagery. Remote Sensing of Environment, 57, 127 – 142. Photographic and videographic observations for determining and map-
Tóth, T., Kertesz, M., & Pásztor, L. (1998). New approaches in salinity/ ping the response of cotton to soil salinity. Remote Sensing of Environ-
sodicity mapping in Hungary. Proceedings 16th International Congress ment, 49, 212 – 223.
of Soil Science, Montpellier, France, 20 – 26 August 1998. International Younes, H. A., Gad, A., & Rahman, M. A. (1993). Utilization of different
Society of Soil Science. 6 pp. remote sensing techniques for the assessment of soil salinity and water
Utset, A., Ruiz, M., Herrera, J., & Deleon, D. (1998). A geostatistical table levels in the Serry Command Area, Egypt. Egyptian Journal of
method for soil salinity sample site spacing. Geoderma, 86, 143 – 151. Soil Science, 33, 343 – 354.
Verma, K., Saxena, A., Barthwal, A., & Deshmukh, S. (1994). Remote Zadeh, L. (1965). Fuzzy sets. Information and Control, 8, 338 – 353.
sensing technique for mapping salt affected soils. International Journal Zhang, M., Ustin, S., Rejmankova, E., & Sanderson, E. (1997). Monitoring
of Remote Sensing, 15, 1901 – 1914. Pacific coast salt marshes using remote sensing. Ecological Applica-
Wiegand, C., Anderson, G., Lingle, S., & Escobar, D. (1994). Soil salinity tions, 7, 1039 – 1053.
effects on crop growth and yield. Illustration of an analysis and map- Zinck, J. A. (2001). Monitoring salinity from remote sensing data. In R.
ping methodology for sugarcane. Journal of Plant Physiology, 148, Goossens, & B. M. De Vliegher (Eds.), Proceedings of the 1st Work-
418 – 424. shop of the EARSeL Special Interest Group on Remote Sensing for
Wiegand, C., Richardson, A., Escobar, D., & Gerbermann, A. (1991). Developing Countries ( pp. 359 – 368). Belgium: Ghent University.

Vous aimerez peut-être aussi