Next Article in Journal
Structural Health Monitoring of Above-Ground Storage Tank Floors by Ultrasonic Guided Wave Excitation on the Tank Wall
Next Article in Special Issue
Synergy Effect of Combining Fluorescence and Mid Infrared Fiber Spectroscopy for Kidney Tumor Diagnostics
Previous Article in Journal
Improved Durability and Sensitivity of Bitterness-Sensing Membrane for Medicines
Previous Article in Special Issue
Development and Testing of an LED-Based Near-Infrared Sensor for Human Kidney Tumor Diagnostics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Application of EM38: Determination of Soil Parameters, Selection of Soil Sampling Points and Use in Agriculture and Archaeology

Chair of Plant Nutrition, Technical University of Munich, D-85350 Freising, Emil-Ramann-Str. 2, Germany
*
Author to whom correspondence should be addressed.
Sensors 2017, 17(11), 2540; https://doi.org/10.3390/s17112540
Submission received: 14 June 2017 / Revised: 7 October 2017 / Accepted: 27 October 2017 / Published: 4 November 2017
(This article belongs to the Special Issue State-of-the-Art Sensors Technology in Germany)

Abstract

:
Fast and accurate assessment of within-field variation is essential for detecting field-wide heterogeneity and contributing to improvements in the management of agricultural lands. The goal of this paper is to provide an overview of field scale characterization by electromagnetic induction, firstly with a focus on the applications of EM38 to salinity, soil texture, water content and soil water turnover, soil types and boundaries, nutrients and N-turnover and soil sampling designs. Furthermore, results concerning special applications in agriculture, horticulture and archaeology are included. In addition to these investigations, this survey also presents a wide range of practical methods for use. Secondly, the effectiveness of conductivity readings for a specific target in a specific locality is determined by the intensity at which soil factors influence these values in relationship to the desired information. The interpretation and utility of apparent electrical conductivity (ECa) readings are highly location- and soil-specific, so soil properties influencing the measurement of ECa must be clearly understood. From the various calibration results, it appears that regression constants for the relationships between ECa, electrical conductivity of aqueous soil extracts (ECe), texture, yield, etc., are not necessarily transferable from one region to another. The modelling of ECa, soil properties, climate and yield are important for identifying the location to which specific utilizations of ECa technology (e.g., ECa−texture relationships) can be appropriately applied. In general, the determination of absolute levels of ECa is frequently not possible, but it appears to be quite a robust method to detect relative differences, both spatially and temporally. Often, the use of ECa is restricted to its application as a covariate or the use of the readings in a relative sense rather than as absolute terms.

1. Introduction

Fast and accurate detection of within-field variation is essential for the detection and management of the environment. The EM38 device (Geonics. Ltd., Mississauga, ON, Canada), a sensor that delivers dense datasets, can be used to accomplish this goal. The EM38 meter is the most widely used EMI sensor in agriculture [1,2].
Researchers have related EM38-ECa (apparent electrical conductivity—ECa) to a number of different soil properties either within an individual field or across the entire landscape [3]. The application of EM38 began with the detection of salinity and continued with the determination of clay and water content [2]. Currently, areas of application include the estimation of nutrient levels and other soil chemical and physical properties, soil sampling points, the determination of soil types and their boundaries, the prediction of yield and the delineation of crop management zones. The increasing application especially during the last decade is also caused by various technical developments: Global Positioning Systems (GPS), surface mapping programs and systems for data analysis and interpretation. Technical data, construction and tool specification are described in Heil and Schmidhalter [4].
This device consists of a receiver and a transmitter coil installed 1.0 m apart at the opposite ends of a nonconductive bar. The investigated depth range depends on the coil configuration and the distance between the coils. While the distance is fixed, the orientation of the coils can be changed. In the vertical mode, the device is in a position perpendicular to the soil, whereas in the second case, the device lies parallel to the soil surface [5,6]. The sensitivity in the horizontal mode is the highest directly below the instrument, while the sensitivity in the vertical position reaches a maximum at approximately 30–40 cm below the instrument. The depth-weighted nonlinearity of the response is shown in Figure 1. The cumulative relative contributions of all soil EC are R(z).
An exact determination of the measurement depth is difficult. Theoretically, the readings acquire an unlimited depth, but in reality, it depends on the electrical contrast. The most common definition is a depth range up to 1.5 m when using the vertical dipole mode and 0.75 m in the case of the horizontal mode [4,5,6].
For wide area measurements e.g., in precision agriculture as well as in field-scale soil property measurements the sensor is mounted on metal-free sledge and pulled behind an all-terrain vehicle equipped with a GPS receiver and data collection computer (Figure 2).
Beside the EM38, EM31 and EM34 electromagnetic devices are also available on the market. In contrast to the EM38, the other devices are designed for the detection of deeper areas of soils, e.g., geological layers, ground water and other subsurface feature associated with changes in ground conductivity. The EM31 has an effective exploration depth of about six metres with an intercoil spacing of 3.66 m. The EM34-3 uses three intercoil spacings—10, 20 and 40 m—to provide variable depths of exploration down to 60 m.

2. Goal of this Study

The objective of this study is to summarize the results of recent measurements and the development of algorithms from ECa measurements obtained with the geophysical sensor EM38. Given the numerous possible subject matters for research in using EM38, this review paper has focused on the following specific fields:
  • Salinity
  • Soil-related properties in non-saline soils
    • Soil texture
    • Soil water content, water balance
    • Soil horizons and vertical discontinuities
    • N-turnover, cation exchange capacity, organic matter and additional soil parameters
    • Soil sampling designs
    • Soil type boundaries
  • Agriculture
    • Agricultural yield variability and management zones
    • Efficiency of agricultural field experimentation
    • Additional application of EM38 in agriculture and horticulture
  • Archaeology
The rationale of this compilation should allow the users of this sensor to understand which variables are today detectable, which objectives are realistic and in which regions applications are widely used. The users have to note that these sensor readings are a composite of soil properties and therefore not a replacement for in-depth knowledge’s about soil and site.

3. Surveying Soil Salinity

Ample information can be found in the literature that describes the potential of EM-38 measurements for the non-invasive detection of in situ soil salinity (Table 1).
Corwin and Lesch [97] summarized five methods that have been used for determining soil salinity in the field: (1) visual crop observations; (2) the electrical conductivity of the soil solution (soil paste or extracts); (3) in situ measurement of electrical conductivity with electrical resistivity (with the Wenner array method); (4) non-invasive measurement of electrical conductivity with ECa and, most recently; (5) in situ measurement of electrical conductivity with time domain reflectometry. Frequently, ECe (e.g., conductivity of aqueous extracts of soil saturated soil paste, EC1:5, EC1:2 or EC1:1, conductivity of soil: water suspensions) was indicated as the most useful and reliable measurement of point-wise salinity detection [43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,81,82,95,96,97,98]. In older publications, ECe alone was often used to identify salt-affected areas [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,81,82,86,93,95,96,97,98,99] Norman [30] developed a salinity classification system based on the range of total dissolved salt concentration (EC1:5) with corresponding groupings of crops with different tolerances to root zone salinity. Soil salinity can be derived from the conductivity of the bulk soil (ECa). For example, salinity is quantified and monitored in irrigated agricultural areas of arid zones by means of ECa measurements using EM38 [28,29,30,31,32,33,34,35,36,37,38,39,40,86]. In areas where saline soils exist, 65% to 70% of the variance in ECa can be explained by the changes in salinity alone [51]. ECa readings can be used to predict the exchangeable sodium percentage and ECe as well [27]. The different terms of salinity can be inferred from the equation ECa = f(ECe(0−Z cm)).
During the last three decades, several calibration methods have been published describing EM38-ECe relationships [27,28,41,60]. Following the classification of Triantafilis et al. [43] and Vlotman et al. [49], further calibration approaches have been proposed, using linear regression, multiple regression coefficients [15,31,37], simple depth weighted coefficients [5,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,67,69,70,71,72,73,74,75,76,77,78,79,80,81,82,86,87,88,89,90,91,92,93,94], established-coefficients [10,11], modelled coefficients [38], mathematical coefficients [9], a logistic profile model [43] and inverted salinity profiles [56].
Johnston et al. [19] reported that EM38 readings are not highly accurate but that categories of soil salinity for large areas can be readily established. Coefficients of determination between 0.88 and 0.9 at depth levels of 30–60, 60–90 and 0–90 cm in soils in which salinity was the dominant factor influencing the EM38 readings were described by Amezketa [59]. A more complex example of these regressions is the dual pathway parallel conductivity (DPPC) model developed by Rhoades et al. [32]. This model indicates the major contribution to ECa readings from conductivity in the water-filled pores that contain the majority of the solved salts with a relatively small contribution from the exchangeable cations. When comparing different ECa-ECe prediction models, the relationships often show low accuracy [5,19,78]. These results suggest that it is essential to establish calibration relationships between ECa and ECe that depend on the soil type and water status for the specific site conditions for a particular survey [19,20]. The variability of ECa to ECe conversion is greater in coarse-textured soils than in medium- or fine-textured soils [24].
The effect of soil salinity and soil water content on the ECa has been described e.g., by Hanson and Kaita [91], Bennett et al. [65], Gill and Yee [16], Turnham [81] and Wittler et al. [54]. The results indicated substantial changes in the ECa readings as soil-water content changed. A linear relationship existed between soil-water content and ECa for each level of soil salinity across the range of measured soil water contents [91]. Norman [30] stated that, for clay soils (i.e., >40% in the top 30 cm), the water content of the soil profile should be greater than 20% to allow soil salinity values to be accurately derived from the observed ECa data. In Iranian investigations, Rahimian and Hasheminejhad [35] found that more reliable regression equations between ECah (horizontal mode) and ECav (vertical mode) and soil salinity could be derived at 35% water content in comparison to 25% water content. Arndt et al. [60] cited similar values from the USDA-Soil Conservation Service. For field surveys where ECa was closely related to salinity, Corwin and Lesch [97] used relationships between the v- and h-mode to derive new variables. The geometric mean (sqrt(ECah*ECav)) provides a measure of the cumulative ECa through the root zone and the ratio mean (ECah/ECav ) characterizes the degree of leaching. A ratio greater than 1 indicates that the net flow of water and salts is upward, and a ratio less than 1 indicates a downward net flow.
Broadfoot et al. [66] and Mankin and Karthikeyan [80] described similar classifications:
  • Leached soils, where salinity increases with depth, defined by ECah/ECav ≤ 1.0
  • Uniform, where salinity does not change significantly with profile depth and where 1.0 < ECah/ECav ≤ 1.05, and
  • Inverted salinity profiles, where salinity decreases with depth and where ECah/ECav > 1.05.
A similar representation was chosen by Spies and Woodgate [93]. Subsoil (EM31) salinity maps and root zone (EM38) maps were combined to provide an assessment of salinity hazard. The EM38 instrument had a depth range of less than 1.5 m, while the EM31 probes had a depth range of 4 to 6 m. Triantafilis et al. [42] developed a leaching fraction model in combination with ECa based on the amount of deep drainage and the average root zone ECe. However, the present investigations are not limited to the creation of real-time inventories but are also of value in forecasting temporal changes in the salinity status. Lesch et al. [100] used pre- and post-ECa surveys to quantify the degree of salt removal from a field. However, the spatial variability impeded the derivations, particularly for subareas with high salinity levels. Salama et al. [101] related apparent conductivity to recharge/discharge mechanisms within watersheds. They associated low values of ECa with low concentrations of total soluble salts and recharge areas. Discharge areas were associated with high values of ECa, indicating greater concentrations of soluble salts near the surface and inverted salt profiles. The latter were associated with rising groundwater tables, increased groundwater flow with mobilization of soluble salts, and greater discharge at or near the surface. All of these factors are related to saline seep development [102].
In an advanced application, EM38-ECa was used to help to assess the salt tolerance of trees, forages and turf grasses [14,16,23,24,25,26,65]. The authors also studied the usefulness of ECa to predict the survival and growth of eucalyptus and pastures in saline soils. According to McKenzie et al. [24,25] and McKenzie [26], close correlations between salinity measured as ECa to the yield of wheat and salinity measured by the saturated paste extract by McKenzie [26] were equal. In contrast, relationships of ECa with observations on the establishment and growth of perennial pasture species were weak [16]. Kaffka et al. [20] reported that, in locations where crop growth were influenced by salinity, ECa was useful for estimating optimum N-fertilizer application and for identifying areas of the field with unprofitable yields. Horney et al. [92] developed a four-step method for site-specific salinity management in commercial fields. The steps included (1) generation of an ECa map; (2) directed soil sampling for ECe; (3) determination of the estimated amendment requirement as a function of location in the field; and (4) integration of the individual amendment requirements into a practical spatial pattern for amendment application. As early as 1997, McKenzie et al. noticed that EM38 is a cost-effective tool for assessing field salinity and for use in experiments on the salt tolerance of crops.
Vaughan et al. [48] combined ECe and water content of soil samples with field wide ECah measurements. The prediction of soil salinity at unsampled points was carried out by co-kriging of logECe with ECah. In a comparison to the work of Triantafilis et al. [44] co-kriging and regression kriging of the ECa readings also showed minimum errors compared to ordinary and three-dimensional kriging.
All of the cited procedures are practical only if salinity is the main factor influencing ECa and if ECe shows a close relationship to ECa [65]. Otherwise, a multiple regression model with further independent influencing factors is required. Consequently, calibration equations and modelled results cannot be used on other sites very often.

4. Detecting Soil-Related Properties in Non-Saline Soils by EM-38

4.1. Influence of Soil Water Content Conditions

In soils where salinity is not a significant factor, ECa values primarily represent as a function of soil water content and the amount of electrical charge. Many researchers recommend measurements with the EM38 at a soil water content close to or at field capacity [49,103,104].
This praxis has its basis in the theory of Rhoades et al. [32] and Corwin and Lesch [97]. In sufficiently wet soils, soil water is the major conductive pathway. Here ECa is determined by the volumetric content of soil water. However, to an increasing extent of researchers noticed that the spatial patterns of ECa, measured under different soil water conditions, are relatively stable with time; only the level indicates a change [105]. However, the relationship between ECa and soil physical and chemical properties varied considerably depending on the actual water conditions. This weak temporal stability of relationships between ECa and other soil properties indicated that soil water conditions have a significant influence on ECa. When there is not enough water in the continuous pores, the surfaces of soil particles and the small discontinuous pores of the soil are the main pathways (e.g., when soil water content is <60 to 70% of field capacity). Under these conditions, the influence of the soil particle volume, the volume and conductivity of water in the small pores, as well as the surface- conductivity of soil particles, increases [32].
Bang [106] showed that several variables (i.e., bulk density, percentage of sand, silt, and clay, plant-available water content, cone index, and saturated hydraulic conductivity) and chemical parameters (i.e., extractable P and K, pH, cation exchange capacity, organic matter, and micronutrients) presented different strengths of the correlations with ECa. Few direct strong correlations were found between ECa and the soil physical properties studied (R2 < 0.50), yet overall, the correlation improved when ECa was measured under relatively dry conditions. Furthermore, according to Bang, the utility of ECa as a variable in cluster analysis to indicate management or soil sampling zones was influenced by variations in ECa measured under different soil water conditions. Bang suggested “that the spatial and temporal ECa variability measured under different soil water conditions could be a critical factor when evaluating the ability of ECa to predict soil chemical and physical characteristics important to soil and crop productivity and management”. Therefore, Bang [106] recommended that an ECa survey be conducted under relatively dry conditions in similar coastal plain soils.
Lück et al. [107] carried out measurements on loamy fields, partly with coarse textured sediments. The authors found the most pronounced ECa distributions during summer (relatively dry conditions). This may has been caused by the larger water content fluctuations in the sandy soils due to their lower water-holding capacity. In contrast to these soils, the loamy parts of the fields had a higher water content as a consequence of higher water-holding capacity as well as better water delivery via capillary rise. Conversely, at sites with dominant Pleistocene loess soils, readings taken during periods when soil water content was at field capacity produced more pronounced maps [108]. Under drier conditions, the ECa readings indicated lower, more similar values. Some researchers recommend a different procedure. Mertens et al. [109] suggested the creation of an averaged map from repeated recordings made at different dates. This procedure is scientifically more appropriate than a water correction. Zhu et al. [110] indicated that the best mapping of major soil distribution across a landscape studied in Pennsylvania required optimal timing, meaning the occurrence of a wet period. No single survey or relative differences in ECa obtained by repeated measurements was sufficient to obtain the best possible soil map for the study area. A combination of repeated surveys, depth to bedrock, and terrain attributes provided the best mapping of soils in this agricultural landscape and doubled the accuracy of the map. ECa measurements collected during the wetter periods (i.e., >10-mm antecedent precipitation during the previous 7 days) showed greater spatial variability (i.e., greater sills and shorter spatial correlation lengths), indicating the influence of soil water distribution on soil ECa [111].

4.2. Soil Texture

Frequently, in non-saline soils, ECa is used to indicate soil texture, particularly clay content. Simulations of silt and sand are rare and seem more likely a by-product. However, the quality of the single relationships are often rather confounding (Table 2). As noted by Corwin and Lesch [112] the target variables correlate inconsistently with ECa mainly as a consequence of: (1) the complex interaction of soil properties; (2) a temporal component of variability that is only weakly indicated by an expected constant variable such as ECa and (3) variable climatic factors.
McBratney et al. [113] and McBratney and Minasny [114] demonstrated that differences in the mineral composition influence the magnitude of the ECa values and therefore the strength of the relationship to the clay content. Kaolin-dominant soil minerals will have smaller conductivities, and soil that mainly contains illite or has a mixed mineralogy will have larger ECa values, but these values are smaller than those for smectitic materials. Furthermore, the authors noticed that at low conductivities (<50 mS m−1), it is quite difficult to separate clay. Wayne et al. [115] derived texture fineness classes from ECa readings. A conductivity greater than 30 mS m−1 indicated clay, and a conductivity less than 5 mS m−1 indicated sand. Furthermore, ECa values between 0 and 10 were classified as sandy loam, and values between 10–20 mS m−1 indicated clay loam. These fineness classes represented a basis for the derivation of the plant-available water content. Domsch and Giebel [116] described another approach to delineate clay content. Working with predominantly sandy soils, the authors indicated that, at field capacity, ECa reflected this property well. However, for soils with water-influenced horizons (gleyic soils), such relationships are very weak and should not be introduced in calculations for mineral soils. A factor scoring that used clay and silt content showed a closer relationship with ECa. Furthermore, the authors related ECa to soil textural classes: an ECa of 0–10 mS m−1 indicated sand or loamy sand, an ECa of 10–20 mS m−1 indicated sand or loamy sand over loam, and an ECa of 20–30 mS m−1 indicated sandy loam or loam. Vitharana et al. [104] used the geometric mean ((ECav·ECah)0.5) to delineate the clay content of the top- and subsoils. Doolittle et al. [117] used ECa to locate small inclusions of sandy soils within a predominately fine-textured alluvial landscape. Bobert et al. [103] improved the relationships between ECa and clay, silt and clay + silt by extracting the drift caused by soil water content calculated from a wetness index map. A multi-site/multi-season approach to calibrate ECa models for predicting clay content across large landscapes was developed by Harvey and Morgan [118]. The fact that the relationships between clay and ECa were similar in all 12 fields, indicated that a single linear regression model could be used to describe the spatial variability of the clay content across all of the fields. This “single calibration approach” used data from a designated calibration area to estimate ECa model parameters that were then combined with data from subsequent fields to predict the soil variability in the observed fields. The single calibration approach is likely applicable to other areas, providing requirements for its use are met. Those requirements include the following: (1) the distribution of the soil property or properties of interest in calibration area should be representative of the study area; (2) the soil property or properties that influence ECa should be the same across the study area; and (3) management practices (e.g., crop rotation and irrigation) should be similar across the study area.
To an increasing extent, methods other than linear regression have been used. Response surface sampling design, fuzzy k-means classification, hierarchical spatial regression modelling and ECa (EM38 and EM34) surveys were applied by Triantafilis and Lesch [119] to produce a map of spatial clay content. Triantafilis et al. [44] combined ECa values (EM38 and EM31) and clay content with different geostatistical methods (co-kriging, regression-kriging and ordinary-kriging). The results suggested that the linear relationship of clay content against ECa (EM38) data used in combination with kriging of regression residuals was the most accurate. Vitharana et al. [104] showed that standardized ordinary kriging of subsoil clay content as the primary variable and the geometric mean ((ECav*ECah)0.5) as the secondary variable gave better results when compared to ordinary kriging and traditional ordinary kriging.

4.3. Soil Water Content, Water Balance

The derivation of the water storage capacity, particularly the field capacity, and the plant-available water content based on electrical conductivity measurements has gained increasing importance. Table 3 provides an overview of current investigation areas and target variables.
Water content, like salinity, is a horizontally and vertically effective dynamic property. In areas where water content is the dominant factor that influences ECa and where water content decreases with depth, ECah > ECav and vice versa [167]. Wayne et al. [115] developed a hierarchical procedure for calculating available water content. ECa was used to target the location for neutron probe samples. The construction of a water content–texture relationship allowed the determination of the available water content and the soil water deficit. Kachanoski et al. [141] found that in soils with a low electrolyte content and a wide range of texture, ECa explained more than 90% of the water content. Additionally, Kachanoski et al. [142] correlated ECa readings with water storage and found that 50–60% of the variations in ECa were explained by water content. Similar levels for coefficients of determination were described by Sheets and Hendrickx [150] and Khakural et al. [143]. Morgan et al. [147] noted that ECa is only applicable in areas with a greater range of water content. The same observation was made by Hedley et al. [135],who calculated an R2 of 42%. Substantial changes in the relationships between ECa readings and soil water content were shown by Hanson and Kaita [91]. The higher the soil salinity was, the more sensitive the ECa readings were to changes in soil water content. A linear relationship existed between soil water content and ECa for each level of soil salinity over the range of measured soil water contents. In a Mollic catena, Brevik et al. [139] found significant relationships between ECa and soil water content that explained 50% to over 70% of the variability. The greatest difference between ECa values in any soils was observed when the soils were moist. Regression line slopes tended to be lower in higher landscape positions indicating greater ECa changes with a given change in soil water content. A relationship between increasing water content and ECa readings from a summit-to-foot slope area of calcareous till parent material with a coefficient of determination of 0.86 was described by Clay et al. Wilson et al. [161,162] derived areas with different water movements from EM31 and EM38 readings. Drying/draining patterns were characterised by a downward shift in ECa with time. Follow-up ECa surveys across high-to-low patterns showed a positive correlation between ECa and water content. Regions with increased horizontal flow showed high conductivities after rainfall. Areas that had preferential vertical flow showed lower EM38 readings after periods of rainfall. For a prototype engineered barrier soil profile designed for waste containment, Reedy and Scanlon [148] and Reedy [149] predicted the average volumetric water content at any location at any time with a linear regression model (R2 = 0.80) and spatially averaged volumetric water contents over the entire area (R2 = 0.99).
Bang [106] described weak and negative relationships between soil water content and ECa values in North Carolina’s Coastal Plains. Little variation in subsoil water content across the study site for each survey date combined with a relatively narrow range of variability in soil texture was the main reason for this result. Furthermore, the variability in other factors (e.g., soil compaction and texture) might have masked the contribution of the water content to ECa variation., The author concluded that the spatial variability of soil water content at a 0- to 75-cm depth could not be directly determined by a field-scale ECa survey at this site, due to the weak relationships between soil water content and ECa. Relationships between plant-available water content and ECa (R2 = 0.78) were derived by Wong and Asseng [152] to transform a water storage capacity map of the field into yield maps for three major season types (dry, medium and wet) and nitrogen fertilizer management scenarios. Hall et al. [159] reported that ECa methods (i.e., EM38 and the use of a borehole conductivity meter) could accurately characterize water and solute distributions in the vadose zone. Saey et al. [168] developed an index to register the area-wide soil heterogeneity. After calculating the relationship between clay content and ECa, this equation was converted so that ECa was the target variable. In the next step, the authors calculated a quotient of the measured ECa and the ECa reading derived from the clay content. This result was called ECref and was used as measure for soil heterogeneity.
Variables other than water content are targets of ECa measurements to an increasing extent; for example, hydraulic conductivity, water table depth, drainage classes and groundwater recharge. In developing a relationship between ECa and estimated deep drainage (mm/year) Triantafilis et al. [42,45] developed four-parameter broken-stick models fitted to ECav beyond 120 cm. Vervoort and Annen [163] showed that the overall patterns of the hydraulic conductivity of palaeochannel in alluvial plains could be inferred from the combination of EM inversion using EM38 and EM34 measurements. However, the absolute magnitude of hydraulic conductivity could not be easily predicted.
Sherlock and McDonnell [169] used simple linear regression analyses to compare terrain electrical conductivity measurements from EM31 and EM38 to a distributed grid of water table depth and soil- water content measurements in a highly instrumented 50 by 50 m hill slope in Putnam County, New York. Regression analysis indicated that EC measurements from the EM31 meter (v-mode) explained over 80% of the variation in the water table depth across the test hill slope. Despite problems with sensitivity and zeroing the EM38 could explain over 70% of the gravimetrically determined soil water variance.
The depth of the water table was also detected by Schuman and Zaman [160]. Knowledge of the water table depth was necessary to select a suitable field for new citrus plantings and for drainage systems. With ECa in the vertical mode, the authors could estimate these values with a RMSE of approximately 4–15 cm. ECa, the topographical wetness index and the rainfall time series gave good predictions of water content and water table depth using the models derived according to Hedley et al. [140]. Further investigations determined soil drainage classes [144], groundwater recharge [170], water drainage [46] and irrigation [164].

4.4. Detection of Soil Horizons and Vertical Discontinuities

To an increasing extent, investigations were carried out to calculate ECa depth profiles in combination with the detection of vertical discontinuities (Table 4). Refining and improving of soil maps is necessary for soil protection and the description of soil functions.
ECa profiling by depth requires more intensive measurements. Usually, this investigation is carried out with measurements made at different heights above the soil surface or repeated measurements at different coil spacing using regressions between ECa and depth for the further calculation [5,9,185,212]. As the instrument is raised above the ground, the relative influence of deeper layers on the measurements decreases. Visual comparison of ECa values and instrument height and inverse modelling (inversion, optimization) are often used. However in numerous cases, the alternating influencing factors impede the retrieval of adequate results; for example, both texture and salinity can cause strong vertical fluctuations. Sudduth et al. [196], Sudduth and Kitchen [155,175,176,177,178,179,181,184,185,186,187,188,195,196,197,198,201,202,203,204,205,206,207,208,209], Kitchen et al. [213] and Noellsch [214] used ECa to determine the depth to the claypan (the sublayer with 50 to 60% clay, varying in depth from 0.1 to 1 m) in nonsaline soils (Missouri). A high correlation between increasing ECa and decreasing depth to the claypan was observed by Doolittle et al. [184]. The depth of boulder clay was estimated by Brus et al. [193], and Bork et al. [191] estimated the loess thickness above basalt. Mapping of sand deposition after floods was carried out by Kitchen et al. [187]. In the investigations of Boettinger et al. [190] soil depth to the petrocalcic horizon was positively and significantly correlated with ECa. Doolittle and Collins [183] reported that bedrock depths on a Pennsylvania site, based on depth classes, could be estimated with ECa data.
Knotters et al. [188] introduced ECa as an auxiliary variable in co-kriging and kriging with regression to predict the depth of Holocene deposits. Vitharana et al. [189] improved the content of a soil map with the calculation of the depth of a Tertiary stratum.

4.5. Relationships to N-turnover, Cation Exchange Capacity, Organic Matter and Additional Soil Parameters

In addition to the previously listed soil properties, further parameters have been combined with ECa readings, including cation exchange capacity, organic matter, bulk density, nutrients (e.g., NO3, Olsen-P) and elements such as Ca, Mg, K, Na (exchangeable or in saturation extract), B, Mo, H and other anions. For close relationships, field-wide ECa measurements allow mapping of soil properties (Table 4). The dominant target variable was the cation exchange capacity [3,132,135].
The leaching rates calculated from a field study were related to changes in ECa readings [209]. This enabled the derivation of a spatially averaged leaching rate. The spatial distribution of N seems to be an increasingly attractive parameter to be estimated via soil conductivity. Eigenberg and Nienaber [215,216] and Eigenberg et al. [217,218] related ECa maps made at different times to temporal values of available N and other specific mobile ions that were associated with animal waste and cover crops, and concluded that ECa can be used as an indicator of the content and loss of water-soluble N. Eigenberg and Nienaber [215,219] isolated and detected areas of nutrient build-up in a cornfield receiving waste. Different manure and compost rates had been applied for replacement of commercial fertilizer. ECa measurements differentiated commercial N-fertilized plots from those that had manure applied at the recommended P rate, compost applied at the P rate, and compost applied at the N rate. In another publication, the same authors [220] discriminated areas with synthetic fertilizer from areas with feedlot manure and compost application. Differences between ECa maps before and after the applications were partly explained by N decompositions. Furthermore, Eigenberg et al. [221] reported that ECa (EM38 and Dualem-2) soil conductivity appeared to be a reliable indicator of soluble N gains and losses in a soil under study in Nebraska, a measure of available N sufficiency for corn mainly in the early growing season, and an indicator of NO3N surplus after harvest when soluble N was vulnerable to loss as a consequence of leaching and/or runoff.
Johnson et al. [204] stated that in soils where ECa is dominated by NO3-N, ECa was applicable for tracking spatial and temporal variations in crop-available N (manure, compost, commercial fertilizer, and cover crop treatments). Furthermore, the calculation of fertilizer rates for site-specific management was possible. Stevens et al. [222] studied ECa as an indirect measure for NH4+ and K+ in animal slurries. The predictive capability of soil conductivity to estimate soil nitrate was demonstrated by Doran and Parkin [223]. Korsaeth [125] found an explanation of a variance of 27–69% (average 47%) of topsoil inorganic N concentration by means of ECa. In general, the author stated that determination of absolute levels of this parameter was difficult with ECa, but it appeared to be quite a robust method for detection of both spatial and temporal relative differences. Some authors described relationships between ECa and soil conditions that influenced soil mineral N [224,225]. Fritz et al. [224] suggested the application of ECa to predict NO3 concentrations in the soil. A comparison of the EM38 and the Veris 3100 sensor cart showed a correlation with soil NO3, but the authors indicated that further studies were necessary to confirm their results.
The studies of Jaynes et al. [226] and Kitchen et al. [213] assumed a possible relationship between soil ECa and N mineralization and denitrification rates. Soil conditions, especially the texture, influenced the rate of denitrification and N mineralization [227]. The relationships between soil texture and N mineralization and denitrification should aid in developing an in-season variable-rate N fertilizer recommendation [224]. Soil organic matter, ECa, and soil texture are properties that might aid in predicting mineralization and denitrification in soil. Dunn and Beecher [228] detected large differences in surface soil acidity and a strong relationship (R2 = 0.49 to 0.91) compared to ECa readings in individual rice fields in NSW, Australia. The proposed ECa levels for the delineation of zones were <80, 80–140 and >140 mS m−1 for the EM31 vertical mode, and <80, 80–110 and >110 mS m−1 for the EM38 vertical mode. Many rice growers in southern NSW currently have EM maps of their fields. Using these maps soil sampling for soil acidity would be a more cost-effective method than grid sampling.
Triantafilis and Momteiro Santos [200] indicated the cation exchange capacity (CEC) as one of the most important soil properties because it is an index of the shrink–swell potential and is thus a measure of soil structural resilience to tillage. The authors used the readings from EM38 and EM31, and additionally remotely sensed spectral reflections (red, green and blue spectral brightness), and two trend surface (Easting and Northing) variables as ancillary data or independent variables, and a stepwise MLR model was used to predict the CEC. The x and y variables accounted for any distinct drift in the residual error pattern. The correlation coefficient (R2 = 0.76) for the regression model was much larger than that achieved with any of the individual ancillary data variables. The adjusted R2 was 0.69, and the estimated RMSE was 1.86 cmol kg−1.
In other studies, the results were more confusing. Heininger et al. [229] and Nadler [230] indicated that salinity, soil texture, or soil water content were masking the response of ECa to other physical, chemical and nutrient levels in soil. Cations, such as Ca, Mg, or K, commonly associated with binding sites on soil particles, could influence ECa with variations in ECS (i.e., conductivity of the solid soil). However, the common assumption is that in most field solutions, changing levels of soil cations have a minor influence on ECS [229,231]. Heininger and Crosier [232] demonstrated that under saturated conditions changes in nutrient levels (e.g., soluble N and S), changes in ECWC could influence ECa. In a study by Heiniger et al. [229], ECa was evaluated as a means to estimate plant nutrient concentrations (i.e., P, K, Ca, Mg, Mn, pH, CEC, and humic content). This study indicated that it was unlikely that ECa could be used to directly estimate the soil nutrient content in a field. However, the authors suggested that additional research on the relationships of ECa with soil water content and soil texture was necessary to determine whether ECa could be used to establish nutrient management zones. The authors concluded that “ECa can be valuable tool when used in conjunction with multivariate statistical procedures in identifying soil properties and their relationship to nutrient availability”.
According to Martinez et al. [203], ECa can provide inexpensive and useful information to capture soil spatial variability and characterization of organic carbon. ECa data were used to elucidate differences in soil properties as a consequence of topography and management, explaining >25% of the spatial variation. With normalized ECa (ΔECa) the authors successfully applied fuzzy k-means to delimit homogeneous soil units related to soil management and the spatial distribution of organic carbon. Grigera et al. [131] related soil microbial biomass to organic matter fractions in a field using ECa. Soil properties (0–90 cm) that showed higher correlations with ECav (Ct (R = 0.87), clay (R = 0.83), total dissolved solids (R = 0.68), and depth of topsoil (R = 0.70)) influenced soil water availability in this field. Soil microbial groups were correlated with different soil C fractions in the uper 15 cm and were similar across ECa zones. Motavalli et al. [233] assessed variation in soil Bray 1 P levels in litter amended landscapes at 0–5 and 5–15 cm depths. ECa was also applied as subsidiary variable in a co-kriging method for improving the map accuracy interpolation of P, K, pH, organic matter and water content [210]. Jung et al. [207] described a similar effect for the application of ECa. Cross-semivariance analysis with ECa as a secondary variable were better than by a simple semivariance analysis.
Bekele et al. [234] reported that ECa was strongly related to ammonium extractable K, organic matter (OM), pH and Bray-2 phosphorus with factor analysis but not to ammonium extractable Ca and the sum of bases in fields in LA, USA. Furthermore Lukas et al. [127] examined soil chemical characteristics (i.e., P, K, Mg content and pH value) and humus content and showed relatively balanced, moderately strong correlations with ECa.
Additionally, the use of ECa for the detection of soil compaction has become increasingly important [192,208]. Krajco [208] discovered that the ECa readings measured in the horizontal mode distinguished the areas with no compaction above 0.3 m and areas with soil compacted in the entire soil profile with less precision. The EM38 operated in the vertical mode was not sensitive enough to measure any differences in soil bulk density.

4.6. Derivation of Soil Sampling Designs

ECa measurements are frequently applied to devise soil sampling schemes to reduce soil sampling points (Table 5) [88,114,115,235,236].
In addition to finding representative locations, the goal is to significantly reduce the number of samples required to effectively calculate the target variable. Frequent selection of sampling points by means of ECa surveys is performed empirically. In principle, design-based (probability-based) and model-based (prediction-based) sampling schemes are applicable.
Triantafilis et al. [42,45] used the ratio (ECav(EM38)/ECav(EM31)) to determine soil sampling points on salt affected areas. Lower ratios appeared when EM38 was sensing the relatively sandy and less conductive topsoil. The results of Shaner et al. [243] support the utilization of ECa-directed zone sampling as an alternative to grid sampling if the transition zones of soil texture and soil organic matter are avoided. Approximately 80% of the samples in grid sites 10 m from the zone boundaries were classified correctly compared to the samples <10 m from the boundary, in which only 50–54% were classified correctly. Corwin et al. [237] described a procedure that was the basis for the development of the ESAP software package [240,241]. In this model-based sampling approach, a minimum set of calibration samples was selected based on the measured ranges and spatial locations of the ECa readings. This sampling approach originated from the response surface sampling design (RSSD) methodology of Box and Draper [244]. The ESAP software was specifically designed for use with ground-based EM signal readings. The ESAP software package tried to identify the optimal locations for soil sampling (6–20 sites depending on the level of variability of ECa) by minimizing the mean square deviation. Zimmermann et al. [235] developed a hierarchical system with (1) ECa measurements; (2) kriging; (3) cluster analysis; (4) principal component analysis and (5) formation of a pseudo-response surface design to select subsets of appropriate sites for soil sampling. The number of samples could be minimized while still retaining the prediction accuracy inherent in statistical sampling techniques. Horney et al. [92] suggested a methodology for salt affected soils with the following steps: (1) building an ECa map; (2) directed sampling for salinity; (3) as a function in the field determination of the estimated improvement requirement and (4) integration into a practical spatial pattern. Tarr et al. [245] used stratification of ECa and terrain attributes to derive a heterogeneous pasture in relatively homogenous sampling zones with fuzzy k-means clustering. The five zones had significant differences in the target variables (i.e., P, K, pH, organic matter and water content). However, the reduction of sampling points from 116 to 30 to 15 points resulted in a loss of accuracy, but this loss may not have an economic or management consequence to the producer. Yao et al. [242] described a completely new method based on Minasny and McBratney [246]. The authors developed the application of the VQT (variance quad-tree) method on sampling design with the digital elevation model and its derivatives and Landsat TM images. ECa was selected as an additional variable, and the spatial distribution map of ECa was used as design detecting salinity. The results show that the spatial distribution of soil salinity detected with the VQT scheme was similar to that produced with grid sampling, while the sample quantity was reduced to approximately one-half. The spatial precision of the VQT scheme was considerably higher than that of the traditional grid method with respect to the same sample number. Fewer samples were required for the VQT scheme to obtain the same precision level. The authors suggested that VQT and ECa provide an efficient tool for lowering sampling costs and improving sampling efficiency in the coastal saline region.

4.7. Derivation of Soil Type Boundaries

Delineating soil classifications has quite different levels of complexity and accuracy. ECa is applied to support the derivation of soil types (Table 4). Very often, the first question concerns the interpolation of the ECa procedure. Niedźwiecki et al. [247] gave an overview of ECa field-wide variability with variograms. The authors recommended an individual interpolation because of differing variability between fields. Selection of parameters for semivariograms has a strong influence on the ability to identify significant spatial autocorrelation of data. Lag parameter size and directional analysis of variance are particular concerns.
The next question concerns the interpolation of ECa across field boundaries. As a consequence of land use, time of measurement, wetness, and fertilization differences between single fields, considerable differences in the ECa levels frequently exist. Weller et al. [121] presented a method for unifying ECa across boundaries with a “nearest-neighbours ECa correction”. ECa measurements near field boundaries were correlated with ECa values of the neighbouring field, resulting in the same level of ECa in both fields. This procedure also enhanced the coefficients of determination.
Another procedure was described by Heil and Schmidhalter [108] (Figure 3). To reduce the levels and to obtain reliable ECa values across field boundaries, the following steps were used: (1) The field-by-field means (mfield) were subtracted from individual observations (Figure 3b); (2) The resulting new ECa (zresidual) values were then used as input to estimate the residual variogram. The ECa data were interpolated, and continuous maps of ECa residuals were obtained (Figure 3c); (3) Finally, the field-by-field means (mfield) were added back to the estimated point-kriged surfaces (zkrig) for each particular field (Figure 3d). With this procedure it is possible to interpolate point wise or row wise measurements with a single interpolation calculation.
Nehmdahl and Greve [128] compared soil profile descriptions and interpolated ECa measurements to derive areas with more or less similar soil types. Stroh et al. [181] distinguished boundaries of soil map units in a relative manner. In different instances, gradients or contrasting inclusions within map units were also identified. In this investigation, correlations between ECa readings and soil properties such as CEC, pH, particle size distribution and extractable bases were low (i.e., explained <6% of the variance) or non-significant. James et al. [178] used confusion matrix analysis to determine whether ECa and a clustered k-means algorithm accurately delineated soil textural boundaries in a field containing clay loam and sandy loam soils. The agreement between the ECa data and the two soil classes was 62%. Hedley et al. [135] derived two soil units (clayey soils and silty loamy soils) with a discriminant analysis of an ECa survey. A more detailed prediction was not possible.
Often, the use of ECa is restricted to its application as covariate or the readings are used in a relative sense, not as absolute terms. In some studies, combination with further predictors such as terrain attributes or yield deliver an acceptable result [179]. Rampant and Abuzar [179] predicted soil types from the various combinations of geophysical (EM38, EM31, airborne gamma radiometrics) and terrain attributes with a decision tree classifier. Individually, the geophysical data were relatively weak predictors of soil information. Using all of the geophysical and terrain data, the soil types were predicted very well, with less than 2% of the area misclassified. Clay et al. [248] empirically derived soil patterns from ECa readings and elevation data. Generally, well-drained soils in the summit area and poorly-drained soils in the valley bottoms had low and high ECa values, respectively.
An interesting comparison between ECa and the soil values of the German national soil inventory (Bodenzahlen) was presented by Neudecker et al. [249]. In 11 fields in four different German regions, R2 varied between 0.1 and 0.71. Highly heterogeneous fields showed a range of R2 values from 0.03–0.71. The authors concluded that ECa measurements were much better in delineating zones of different soil substrates than other, rather subjective methods such as the German national soil inventory.

5. Applications in Agriculture

5.1. Derivation of Agricultural Yield Variability and Management Zones

ECa is used to reflect crop yields and to derive management zones. Different studies show that crop yields vary due to site-specific differences and temporal climatic changes (Table 6).
Management (productivity) zones with similar yields and used by farmers to make application decisions based upon calculations of the expected yield. The applied methods and additional predictors are different in this context. In fact, ECa has no direct relationship to the growth and yield of plants, but the spatial variation of ECa is partly correlated with soil properties that do affect crop productivity. Several studies have shown this connection [88,127,213,226,271]. The advantage of ECa in comparison to yield measurements is its relative temporal stability, which offers a better basis for the delineation of management zones than variable yield mapping information does. With cluster analysis, Fleming et al. [258] confirmed that management zones represented different suites of soil. In one field, soil organic matter, clay, nitrate, potassium, zinc, ECa and corn yield data corresponded to the levels indicated by the management zones. In a different field, only the medium productivity zone had the highest values for these parameters. Cockx et al. [253,254] used the spatial distribution of NO3 in addition to ECa to create nitrogen management zones. The interpolated ECa measurements were the input for a fuzzy k means classification. This procedure placed each single point in a membership in each class [46]. The method minimized the multivariate within-class variance, and consequently, individuals in the same class had similar attributes [283]. Using a principle compound analysis, (PCA) Vitharana et al. [189,281] detected the importance of pH, ECa-v and organic matter as independent key variables to characterize overall soil variation. The authors identified and delineated four classes (with a fuzzy k-means algorithm) with these variables. Clear differences in soil properties and landscape positions were found between these classes, and the three-year average standardized yields (grain and straw) were also different across the classes. Schepers et al. [277] aggregated brightness images, elevation, ECa and yield into management zones using principal component analysis in combination with unsupervised classification. Domsch et al. [257] correlated ECa and yield within the boundary lines method. In this context, Corwin et al. [284] combined ECa with leaching of pollutants and Johnson et al. [204] combined ECa with soil quality parameters (measured as bulk density, water content, clay content, organic matter, N, extract-able P, pH, microbial biomass C and N, potentially mineralizable N). In an investigation on claypan soil, Sudduth et al. [196] described a negative relationship between ECa and grain yield in a dry year. The correlations with corn and soybean in a wet year in topographically highly variable landscape were also negative, as observed by Jaynes et al. [226,266]. However, in both studies no significant relationships were observed in years with a more normal water supply. In a newer study of claypan areas, Jung et al. [132] described negative relationships for corn and soybean in years with more than 150 mm precipitation, while in contrast, ECa was positively correlated in years with less than 150 mm precipitation. In both cases, the correlation coefficients were not higher than 0.74. However, the authors concluded, “while correlation analysis itself is far from a definitive analysis, we suspect this similar pattern (between ECa and yield) in correlation is not coincidental”. Kitchen et al. [213] related ECa to yield applying boundary line analysis on claypan soils. A significant relationship (boundary lines with R2 > 0.25 on most areas) was apparent, but climate, crop type, and specific field information was also necessary to explain the structure of the potential yield by ECa interaction. The authors divided the relationships between productivity and ECa into four categories: (1) positive; (2) negative; (3) positive in some portions of the field and negative in others; and (4) no relationship. The strongest relationships were negative, reflecting the tendency of claypan soils to be water-limited for crop production in the majority of growing seasons [133]. Figure 4 and Table 7 show the relationships between ECa (EM38 in both configurations) and yield of the long-term field experiment Dürnast 020 (South Germany, (4477221.13E, 5362908.78N), Heil, unpublished).
Here the application of different N-fertilizers with two fertilization levels has been tested since 1979. In the Figure 4 the multi annual means of the yields of wheat (1980, 1983, 1986, 1989, 1992, 1995, 1998, 2001, 2004, 2007, 2010, 2012) were divided in the two fertilization levels and the unfertilized control plots. Within this site, soils were mapped as deposits of Pleistocene loess, and the dominating soil types were fine-silty Dystric Eutrochrept and fine-loamy Typic Udifluvent (German Soil Survey, Bodenkundliche Kartieranleitung 2005). On this productive field (plant available water capacity 250 mm until 100 cm depth , C-content: 1.4% (0–30 cm) and 0.4% (50–75 cm)) all relationships are negative with always significant R2 and also linear or weak quadratic curves. Remarkable is that the curves have similar slopes, at least in the higher ECa range. The always lower coefficients of determination in the case of the vertical configuration could reflect, that the deeper soil is less important to the plant growth.
After a first visual inspectation the lowest values of yield correspond with higher contents of clay. The curve progressions allow further interpretations:
  • The spatial distribution of the yield was at first influenced by the ECa across the field. Treatment effects (fertilizing level, fertilizer form) were overlain by soil conditions with different ECa values.
  • The height of the yield was secondly assumedly determined by the level of fertilization.
In claypan soils, Fraisse et al. [260] also used a combination of ECa and topographic features (with unsupervised classification) to develop zones and evaluated their ability to describe yield variability. By dividing a field into four or five zones based on ECa, slope, and elevation, 10% to 37% of corn and soybean yield was explained. In this context, Fridgen et al. [262] described software with a similar derivation of the subfield management zone. Kitchen et al. [285] used unsupervised fuzzy-k-means clustering to delineate productivity zones with ECa and elevation measurements on claypan soils. Productivity zones were also derived by Jaynes et al. [267] based on a series of profiling steps in combination with cluster analysis to determine the relationship between yield clusters and easily measured terrain attributes (i.e., slope, plane curvature, aspect, depth of depression) and ECa. In contrast to the previous investigations, Kilborn et al. [269] found no strong relationships between elevation, slope, and soil ECa with respect to biomass yield and composition. The results of Bang [106] indicate that clustering with ECa and NIR surveys could be used to delineate management zones that characterize spatial variations in soil chemical properties. However, these zones were less consistent for characterizing spatial variability in yields across temporal water content variation. Furthermore, the author reported that clustering zones developed from ECa values measured under relatively dry conditions were particularly effective in partitioning the spatial variability of SOM. It is clear that zones developed from clustering elevation and bare-soil NIR radiance were more effective than ECa alone in capturing variability in K, CEC, and SOM. Clustering on ECa with elevation and NIR provided better zones for these parameters and somewhat reduced the variability associated with measuring ECa under different soil water conditions [106].
A similar praxis was used by Schepers et al. [277]; Chang et al. [252] and Fridgen et al. [262]. Cluster analysis of an ECa map alone or with auxiliary data, such as terrain attributes and bare-soil images, has been widely used to delineate soil-based management zones. The relationship between ECa measurements, soil properties and sugar beet yields in salt-affected soils was studied by Kaffka et al. [20]. In these soils, yield was most highly correlated with salinity. This work demonstrated the utility of relationships between ECa and crop yield to answer resource input questions. Rampant and Abuzar [286] predicted yield zones from a combination of geophysical (i.e., EM38, EM31, airborne gamma radiometrics) and terrain attributes with a decision tree classifier. Individually, the geophysical data were relatively poor predictors of the yield zones. The combination of all sensors and terrain data could predict yield zones quite well, misclassifying only 5% of the area. The predictions of yield for an individual year were always worse for yield zones.
The purpose of the application of the EM38 by Guretzky et al. [263] was to examine the relationship of the relief parameter “slope”, ECa, and legume distribution in pastures. The authors concluded that slope and ECa data were useful in selecting sites in pastures with higher legume yield and showed a potential for use in site-specific management of pastures. Dang et al. [12] used an interesting procedure for identifying management zones on a salinity-affected field. Two surveys of ECa measurements were carried out; the first used a relatively wet soil profile (April–May 2009) to represent the drained upper limit of soil water, and the second used a relatively dry profile (October–November 2009) to represent the lower limit of soil water content extraction following the harvest of the winter crop. The authors developed a framework to estimate the monetary value of site-specific management options through: (1) identification of potential management classes formed from ECa at lower limit of soil water content; (2) measurement of soil attributes generally associated with soil constraints in the region; (3) grain yield monitoring; and (4) simple on-farm experiments.
Islam et al. [264] estimated key properties to identify management zones on loess and sandy soils. The authors identified ECa, topsoil pH, and elevation as key properties, which were used to delineate management classes and to construct an excellent multiple regression model between yield and the key properties. Additionally, Islam et al. [265] described the construction of waterproofed housing for the EM38, which was built using PVC pipes for swimming in a paddy rice field. The ECa data were classified into three classes with the fuzzy k-means classification method. The variation among the classes was related to differences in subsoil bulk density. The smallest ECa values representing the lowest yield and also the lowest bulk density.
There was also a significant difference in rice yield among the ECa classes, with Vanderlinden et al. [280] carried out a procedure for characterizing a management system. ECa patterns expressed as relative differences (ϑij) were associated with topography, soil depth and soil structure, and the authors derived management zones with principal component analysis.
A very detailed insight into the relationship between ECa and yield was given by Robinson et al. [275] for sites in Victoria, Australia. However, the multi-year measurements of yield and ECa delivered an inconsistent picture. Significant influences of ECa on yield were found for all measurements, but they evidenced alternating directions in semi-arid and rainy environments. (1) Decreasing yield was combined with increasing ECa-v when texture-contrast and gradational soils with shallow topsoils occurred along with increasing clay content and physio-chemical constraints; (2) In soils without significant texture-contrast, in which physio-chemical conditions were more favourable for water in the subsoil, higher yields resulted; (3) Positive trends of ECa and yield were attributed to the occurrence of higher plant-available water in the root zone in high and moderate yield zones. However, the R2 did not exceed 0.15 for all calculations.
Additionally, the EM38 has been applied in vineyards for describing soil variability to an increasing extent [5,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,62,70,71,72,73,74,75,76,77,78,79,80,81,82,86,90,91,92,93,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270]. Bramley et al. [250] described a close relationship between ECa readings from stony shallow soils and trunk circumference. However, sufficient predictors for vine vigour were not found in these investigations.
EM38 has more rarely been applied to apple orchards. Türker et al. [279] produced ECa maps and compared them with yield and pomological characteristic maps. As a result, the highest value of a non-linear regression between ECa and apple yield was determined with an R2 of 0.94.

5.2. Improvement of the Efficiency of Agricultural Field Experimentation

Only a few publications reported about the application of ECa readings to improve the efficiency of field experiments. An accurate comparison of treatments within agricultural field experiments is the primary objective of these evaluations. Spatial soil variability can have adverse effects on the accuracy and efficiency of such trials (Table 8).
Kravchenko et al. [289] used ECa as a covariate to improve the accuracy of P values on field with different levels of manure applications. Standard errors for the means of P with ECa as a covariate were smaller than those for which ECa was not used as a covariate. In soils with medium and high ECa values, the control treatment (no manure) had a significantly lower P concentration.
Johnson et al. [204] applied field wide ECa readings as a classification parameter for a block design. Blocks were located in homogeneous areas based upon measurements of soil parameters that are significant for yield. The authors noted that ECa classification can be used as a basis for blocking only when ECa and yield are correlated. On these sites, which were described by Johnson et al. [204], the dominating factors were salinity and clay content. The authors described the application of ECa as a “compelling tool in statistical design”.
The initial point of the publication of Lawes and Bramley [288] is the fact that farmers and their advisers are often not able to implement methods that are necessary for evaluation trials on their farms. The authors explore a new and simple approach to the analysis of farmer strip trials and the spatial variability of treatment response. Yield data descriptions with a linear model that accounted for the spatial autocorrelation in the data and a moving pairwise comparison of treatments were applied by the authors. The results suggest that the pairwise comparison adequately identified treatment differences and their significance. This method can be readily implemented and expanded with ECa readings, and it offers an important advance to facilitate on-farm experimentation using precision agriculture technologies.
Brevik et al. [173] indicated a need to investigate the application of ECa techniques in fields with more homogenous soil properties. For these investigations, the authors selected a field with lacustrine-derived soils that exhibited only weak spatial variability in soil properties. The highly uniform ECa readings obtained did not allow differentiation of soil map units with the ECa data. However, the results did confirm the uniform nature of the soils in the field, a critical criterion for precision agriculture applications. An example of the application of conductivity values is given in Table 9 [4].
The relationships presented in Section 5.1 between ECa and yield are here integrated in a variance of analysis (ANOVA) and an analysis of covariance (ANCOVA) with the target to model the multi-annual yield of the long-term experiment Dürnast 020. In the ANOVA only the factors “fertilizing level“ and “the form of fertilizer” have been considered. To enhance the accuracy of the simulation the covariates ECa as well as topographical parameters have been added. The ANOVA procedure delivers with the fertilization level as the single influencing factor only a weak result (R2 = 0.185, RMSE = 3.26 dt ha−1). In contrast to this result the application of the ANCOVA introduced the factors fertilization level and fertilization no. and the covariate ECa (EM38-h and EM38-v) in the simulation. The R2 of 0,875 and a RMSE with 1.29 dt ha−1 indicate a severe enhancement in comparison to the ANOVA. The partial eta-square illustrates that the introduction of the ECa readings was the main reason of this improvement. The topographical parameter channelnet (channel network base level (-)) and TWI (topographical wetness index (-)) had only minor meaning.
Here, ECa has been shown to be a useful indicator of soil variability. Compared to the standard analysis ANOVA, an ANCOVA with ECa as covariate (and also topographical parameters) reduced RMSE and enhanced R2 for treatment means and improved the accuracy of this field experiment.

5.3. Additional Application of EM38 in Agriculture and Horticulture

Additionally, some publications describe the use of ECa to assess environmental susceptibility and/or effects (Table 10).
Jaynes et al. [292] correlated ECa readings with herbicide partition coefficients. The maps are useful for determining areas with a higher leaching potential for herbicide (atrazine) application. Olesen et al. [293] developed two different algorithms (an empirical model and a causal model) for spatially varying fungicide applications. Both models make use of a ratio vegetation index and EM38 measurements. ECa maps describe the soil characteristics, in particular the soil clay content.
Hbirkou et al. [291] used ECa maps for constructing relationships between ECa and the beet cyst nematode, Heterodera schachtii. This nematode prefers deep soil with medium to light soil and non-stagnic water conditions. Correlations between ECa and nematode population density were moderate (R2 = 0.47) and strong (R2 = 0.74). Management maps based on ECa and soil taxation maps indicated areas with different soil-related living conditions for H. schachtii. These maps could make farmers able to improve site-specific management strategies on nematode-infested fields.
Grigera et al. [131] created four ECa zones from ECa readings, based on ranges of both configurations using an unsupervised classification. Soil microbial groups were correlated with different soil C fractions in the upper soil (−15 cm) and were similar across ECa zones. Zone distribution and biomarkers correlated in dependence of the fractions of particulate organic matter (fine particulate organic matter: bacterial (R = 0.85), actinomycetes (R = 0.71) biomarker concentrations; coarse particulate organic matter: bacteria R = 0.69, actinomycetes R = 0.48). In contrast, fungal (R = 0.77) and mycorrhizal (R = 0.48) biomarker concentrations were correlated only with coarse organic matter.

6. Application of EM38 in Archaeology

The application of the EM38 device is not restricted to soil properties; it also detects extrinsic components (Table 11).
Ferguson [298] applied ECa values to find metal objects in a settlement area from the 18th century. Measurements of ECa also appear to be suitable to search for graves [303]. Low values can indicate a proximity to metal, but high conductivity has been associated with grave shafts at one cemetery.
A more sophisticated procedure for archaeological detections was described by Dalan and Bevan [296]. An EM38 meter, which was operated in the inphase mode, measured the susceptibility of the top half-meter of soil. This susceptibility sounding was performed using a series of heights from 2 m to the surface, with readings taken at intervals of 5 cm. These measurements were analysed with the aid of the depth sensitivity function of McNeill [304]. In this manner, the authors could detect magnetic layers to a depth of 50 cm.
Viberg et al. [302] combined the EM38 with the MS2D (Bartington MS2 magnetic susceptibility meter). The anomalies contained in the survey data were explained by the subsequent archaeological excavation. A rubbish pit which consist mainly of organic material and fire-cracked stones was detected in both the MS2D and EM-38 data. This study of Simpson et al. [301] used additionally a fluxgate gradiometer measurements on an archaeological site. The results of the first survey showed very strong magnetic anomalies in the central field, which were caused by the brick remains of the castle. The most useful results with the EM38 were obtained from the magnetic susceptibility. Its anomalies corresponded well with the gradiometer anomalies. To enhance ECa maps, Santos et al. [299] recommended a simple procedure to eliminate the effect of elevation on ECa. In the experience of the authors, soil anomalies are partly changed by changing the elevation within an investigation area according to the water table depth or the conductive sediment layer. With a linear dependence between conductivity and the site elevation the influence of topography was removed. Corrected ECa maps substantially improved the recognition of anomalies. These maps also show a greater similarity with magnetic susceptibility maps, with both identifying archaeological structures of interest: a well-structured fireplace and a concentration of ceramic fragments.

7. Conclusions and Closing Remarks

There is no doubt that EM38 measurements have an increasing importance in exploration of areas, but weaknesses/unclarities of the method are also described in the literature:
  • The interpretation and utility of ECa readings are highly location and soil-specific; the soil properties contributing to ECa measurements must be clearly understood. From the various calibration results, it appears that regression constants for relationships between ECa, ECe, soil texture, yield, etc. are not necessarily transferable from one region to another. Several factors affect the strength of the signal and therefore, the relationships. In addition to texture, salt concentration and other physicochemical properties, calibrations are further affected by the relative response of the signal according to depth, the non-linearity of the signal and the collinearity between horizontal and vertical readings. The soil parameter with the greatest influence on ECa is also the best derivable.
  • Only a few authors [108,196] account for the influence of the farming system, crop biomass, applications of fertilizer at the time of measurement on ECa distributions. Most of the identified soil parameters that influence ECa have significant interdependency and can thus provide multivariate effects on ECa.
  • The modelling of ECa, soil properties, climate and yield are important for identifying the geographic extent to which specific applications of ECa technology (e.g., ECa – texture relationships) can be appropriately applied.
  • In the case of detecting salinity, obviously better results are achieved if both EM38 readings (vertical and horizontal) are combined with ECe values from different depth ranges. Nevertheless, Vlotman et al. [37] posed the question about the need for converting the ECe from ECa. As McKenzie [24,25] showed, a classification of salinity tolerance level of different crops is also possible only with EM38 readings. A partitioning in areas of low, medium and high salinity with measurements in a single mode or with a combination of v- and h-mode is often a sufficient inventory of the salinity distribution. But it is necessary to take into account, that on the one field e.g., 60 mS m−1 has salt problems while another field with the same reading does not have such problems. Therefore ECe will continue to be important at least in the near future.
  • The quality of a regression is often determined by a sufficient range of dependent and independent variables. Delin and Söderström [124] noted that when the ECa data were correlated with the clay content over the whole farm, the result was much better then when the correlation was restricted to single zones. This quality is also better if the target variable is also the dominant ECa-influencing factor.
  • The construction of soil sampling designs with ECa readings is limited to those properties that correlate with ECa. Other parameters require some other sampling approach such as random, grid, or stratified random sampling.
The world-wide application of the EM38 (and also of other soil sensors) is very varying:
  • It seems that the detection of salinity is still the main area of application.
  • Site-specific management in agriculture with the application of ECa is still in Germany in an initial phase of adoption among farmers. Predicting the future is difficult. Nonetheless, a greater presence of site-specific crop management based on soil detection is to be hoped for.
  • Furthermore in Germany increases the investigations in improving soil maps and in detecting soil functions, including: plant available water, sorption capacity, binding strength for heavy metals, filtering of unbound substances and natural soil fertility. Additionally, soil protection measures are also indicators for erosion prevention, retention of nutrients, and conservation/enhancement of carbon contents (based on good agricultural practice after Article 17, German Soil Protection Act). The selection of soil functions is based on the German Soil Protection Act (LABO—Bund-Länder-Arbeitsgemeinschaft Bodenschutz). Here it is not common sense to carry out this also with EM38. Until now it is not well known that, compared to traditional soil survey methods, EM38 readings can more effectively characterize diffuse soil boundaries and identify areas of similar soils within mapped soil units. This gives soil scientists greater confidence in their soil mapping.
  • The application in forests is world-wide rather seldom. But also here is an enormous potential to improve the existing site maps and to test the water distribution between the trees.
  • The improvement of evaluation of field experiments with ECa readings as covariate is more rarely used. The spatial variability of soil properties can have adverse effects on the accuracy and efficiency of field experiments. Here is a great potential to take into account the soil conditions by using ECa readings.
  • The fusion of the data of other sensors also shows great potential. The idea behind the combination of proximal soil sensors is that the accuracy of a single sensor is often not sufficient. The reading of one sensor is affected by more than one soil property of interest. The fusion of sensor data can overcome this weakness by extracting complementary information from multiple sensors or sources. Until now to an increasing extent, the readings of EM38 are evaluated in combination mainly with VIS–NIR and a gamma-ray-spectrometer.
  • Many of the instruments measure at the point or sample scale, such as soil moisture probes and tensiometers, while remote sensing devices determine regional patterns. But these techniques are limited in the depth of penetration into the subsurface.
Here geophysical methods have a positive impact, obtaining data at a range of spatial scales across fields. This survey has shown that considerable progress has been made in detection and understanding of soil functions within the last decades. Applications of practical sensors such as the EM38 are needed to achieve sustainable agriculture, to optimize economic return and to protect the environment, especially the soil.

Acknowledgments

The BMBF Project CROP.SENSE.net No. 0315530C and the BMBF Project Bonares No. 031A564E funded this research.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

CECCation exchange capacity
ECaApparent electrical conductivity
ECavApparent electrical conductivity, measured in vertical mode
ECahApparent electrical conductivity, measured in horizontal mode
ECeElectrical conductivity of aqueous soil extracts EC1:5, EC1:2 or EC1:1, soil/water suspensions)
ECpECa calculated by using predictive equations
ECrefQuotient of the measured ECa and the EC
θv, θwWeighted water content after vertical and horizontal mode
ZSoil depth

References

  1. Sudduth, K.A.; Drummond, S.T.; Kitchen, N.R. Accuracy issues in electromagnetic induction sensing of soil electrical conductivity for precision agriculture. Comput. Electron. Agric. 2001, 31, 239–264. [Google Scholar] [CrossRef]
  2. Doolittle, J.A.; Brevik, E.C. The use of electromagnetic induction techniques in soils studies. Geoderma 2014, 223–225, 33–45. [Google Scholar] [CrossRef]
  3. Sudduth, K.A.; Kitchen, N.R.; Wiebold, W.J.; Batchelor, W.D. Relating apparent electrical conductivity to soil properties across the north-central USA. Comput. Electron. Agric. 2005, 46, 263–283. [Google Scholar] [CrossRef]
  4. Heil, K.; Schmidhalter, U. Comparison of the EM38 and EM38-MK2 electromagnetic induction-based sensors for spatial soil analysis at field scale. Comput. Electron. Agric. 2015, 110, 267–280. [Google Scholar] [CrossRef]
  5. McNeill, J. Electromagnetic Terrain Conductivity Measurement at Low Induction Numbers. Available online: http://www.geonics.com/pdfs/technicalnotes/tn6.pdf (accessed on 1 November 2017).
  6. Geonics Limited. Ground Conductivity Meters. Available online: http://www.geonics.com/html/conductivitymeters.html (accessed on 15 August 2017).
  7. Corwin, D.L.; Lesch, S.M. Apparent soil electrical conductivity measurements in agriculture. Comput. Electron. Agric. 2005, 46, 11–43. [Google Scholar] [CrossRef]
  8. Cassel, F.; Goorahoo, D.; Sharmasarkar, S. Salinization and Yield Potential of a Salt-Laden Californian Soil: An In Situ Geophysical Analysis. Water Air Soil Pollut. 2015, 226, 1–8. [Google Scholar] [CrossRef]
  9. Cook, P.G.; Walker, G.R. Depth Profiles of Electrical-Conductivity from Linear-Combinations of Electromagnetic Induction Measurements. Soil Sci. Soc. Am. J. 1992, 56, 1015–1022. [Google Scholar] [CrossRef]
  10. Corwin, D.; Rhoades, J. An improved technique for determining soil electrical conductivity-depth relations from above-ground electromagnetic measurements. Soil Sci. Soc. Am. J. 1982, 46, 517–520. [Google Scholar] [CrossRef]
  11. Corwin, D.; Rhoades, J. Measurement of inverted electrical conductivity profiles using electromagnetic induction. Soil Sci. Soc. Am. J. 1984, 48, 288–291. [Google Scholar] [CrossRef]
  12. Dang, Y.P.; Dalal, R.C.; Pringle, M.J.; Biggs, A.J.W.; Darr, S.; Sauer, B.; Moss, J.; Payne, J.; Orange, D. Electromagnetic induction sensing of soil identifies constraints to the crop yields of north-eastern Australia. Soil Res. 2011, 49, 559–571. [Google Scholar] [CrossRef]
  13. Doolittle, J.; Petersen, M.; Wheeler, T. Comparison of two electromagnetic induction tools in salinity appraisals. J. Soil Water Conserv. 2001, 56, 257–262. [Google Scholar]
  14. Dunn, G.; Taylor, D.; Nester, M.; Beetson, T. Performance of twelve selected Australian tree species on a saline site in southeast Queensland. For. Ecol. Manag. 1994, 70, 255–264. [Google Scholar] [CrossRef]
  15. Ghany, A.H.; Omara, A.M.; El Nagar, M.A. Testing Electromagnetic Induction Device (EM 38) Under Egyptian Conditions; Vlotman, W.F., Ed.; EM38 Workshop: New Delhi, India, 2000. [Google Scholar]
  16. Gill, H.S.; Yee, M. EM-38 for Assessing Surface and Sub-Soil Salinity and Its Relationship to Establishment and Growth of Selected Perennial Pasture Species. In Proceedings of the SuperSoil 2004—3rd Australian New Zealand Soils Conference, Sydney, Australia, 5–9 December 2004. [Google Scholar]
  17. Herrero, J.; Ba, A.; Aragüés, R. Soil salinity and its distribution determined by soil sampling and electromagnetic techniques. Soil Use Manag. 2003, 19, 119–126. [Google Scholar] [CrossRef]
  18. Li, H.; Li, F.; Shi, Z.; Huang, M. Three Dimensional Variability of Soil Electrical Conductivity Based on Electromagnetic Induction Approach. In Proceedings of the 2010 International Conference on Artificial Intelligence and Computational Intelligence (AICI), Sanya, China, 23–24 October 2010; pp. 219–223. [Google Scholar]
  19. Johnston, M.A.; Savage, M.J.; Moolman, J.H.; du Plessis, H.M. Evaluation of calibration methods for interpreting soil salinity from electromagnetic induction measurements. Soil Sci. Soc. Am. J. 1997, 61, 1627–1633. [Google Scholar] [CrossRef]
  20. Kaffka, S.R.; Lesch, S.M.; Bali, K.M.; Corwin, D.L. Site-specific management in salt-affected sugar beet fields using electromagnetic induction. Comput. Electron. Agric. 2005, 46, 329–350. [Google Scholar] [CrossRef]
  21. Lesch, S.M.; Rhoades, J.D.; Lund, L.J.; Corwin, D.L. Mapping Soil-Salinity Using Calibrated Electromagnetic Measurements. Soil Sci. Soc. Am. J. 1992, 56, 540–548. [Google Scholar] [CrossRef]
  22. McNeill, J.D. Rapid, accurate mapping of soil salinity by electromagnetic ground conductivity meters. Adv. Meas. Soil Phys. Prop. Bring. Theory Pract. 1992, 209–229. [Google Scholar]
  23. McKenzie, R.C.; Chomistek, W.; Clark, N.F. Conversion of Electromagnetic Inductance Readings to Saturated Paste Extract Values in Soils for Different Temperature, Texture, and Moisture Conditions. Can. J. Soil Sci. 1989, 69, 25–32. [Google Scholar] [CrossRef]
  24. McKenzie, R.C.; Mathers, H.M.; Woods, S.A. Salinity and Crop Tolerance of Ornamental Trees and Shrubs; Alberta Special Crops and Horticultural Research Center: Brooks, AB, Canada, 1993.
  25. McKenzie, R.C.; George, R.J.; Woods, S.A.; Cannon, M.E.; Bennett, D.L. Use of the Electromagnetic-Induction Meter (EM38) as a Tool in Managing Salinisation. Hydrogeol. J. 1997, 5, 37–50. [Google Scholar] [CrossRef]
  26. McKenzie, R.C. Salinity: Mapping and Determining Crop Tolerance with an Electromagnetic Induction Meter (Canada). Available online: http://www2.alterra.wur.nl/Internet/webdocs/ilri-publicaties/special_reports/Srep13/Srep13-h6.pdf (accessed on 1 November 2017).
  27. Nettleton, W.D.; Bushue, L.; Doolittle, J.A.; Endres, T.J.; Indorante, S.J. Sodium-Affected Soil Identification in South-Central Illinois by Electromagnetic Induction. Soil Sci. Soc. Am. J. 1994, 58, 1190–1193. [Google Scholar] [CrossRef]
  28. Nogues, J.; Robinson, D.A.; Herrero, J. Incorporating electromagnetic induction methods into regional soil salinity survey of irrigation districts. Soil Sci. Soc. Am. J. 2006, 70, 2075–2085. [Google Scholar] [CrossRef]
  29. Norman, C.P. Kyvalley [Victoria] EM38 Salinity Survey. Available online: http://agris.fao.org/agris-search/search.do?recordID=AU9430080 (accessed on 1 November 2017).
  30. Norman, C.P. Training Manual on the Use of the EM38 for Soil Salinity Appraisal; Victorian Department of Agriculture and Rural Affairs: Victoria, Australia, 1990.
  31. Rhoades, J.; Corwin, D. Determining soil electrical conductivity-depth relations using an inductive electromagnetic soil conductivity meter. Soil Sci. Soc. Am. J. 1981, 45, 255–260. [Google Scholar] [CrossRef]
  32. Rhoades, J.D.; Manteghi, N.A.; Shouse, P.J.; Alves, W.J. Soil Electrical-Conductivity and Soil-Salinity—New Formulations and Calibrations. Soil Sci. Soc. Am. J. 1989, 53, 433–439. [Google Scholar] [CrossRef]
  33. Rhoades, J.D.; Chanduvi, F.; Lesch, S. Soil Salinity Assessment: Methods and Interpretation of Electrical Conductivity Measurements; FAO—Land and Water Development Divison: Rome, Italy, 1999; p. 166. [Google Scholar]
  34. Rhoades, J.D.; Corwin, D.L.; Lesch, S.M. Geospatial measurements of soil electrical conductivity to assess soil salinity and diffuse salt loading from irrigation. In Assessment of Non-Point Source Pollution in the Vadose Zone; Amer Geophysical Union: Washington, DC, USA, 1999; pp. 197–215. [Google Scholar]
  35. Rahimian, M.H.; Hasheminejhad, Y. Calibration of electromagnetic induction device (EM38) for soil salinity assessment. Iran. J. Soil Res. 2011, 24, 243–252. [Google Scholar]
  36. SriRanjan, R.; Karthigesu, T. Evaluation of an Electromagnetic Method for Detecting Lateral Seepage Around Manure Storage Lagoons; ASAE paper No. 952440; American Society of Agricultural Engineers: St. Joseph, MI, USA, 1995. [Google Scholar]
  37. Sharma, D.P.; Gupta, S.K. Application of EM38 for Soil Salinity Appraisal: An Indian Experience. Available online: http://content.alterra.wur.nl/Internet/webdocs/ilri-publicaties/special_reports/Srep13/Srep13-h3.pdf (accessed on 1 November 2017).
  38. Slavich, P.; Petterson, G. Estimating average rootzone salinity from electromagnetic induction (EM-38) measurements. Soil Res. 1990, 28, 453–463. [Google Scholar] [CrossRef]
  39. Slavich, P. Determining ECa-depth profiles from electromagnetic induction measurements. Soil Res. 1990, 28, 443–452. [Google Scholar] [CrossRef]
  40. Soliman, A.S.; Farshad, A.; Sporry, R.J.; Shrestha, D.P. Predicting salinization in its early stage, using electro magnetic data and geostatistical techniques: A case study of Nong Suang district, Nakhon Ratchasima, Thailand. In Proceedings of the 25th Asian Conference on Remote Sensing, Chiang Mai, Thailand, 22–26 November 2004. [Google Scholar]
  41. Sheets, K.R.; Taylor, J.P.; Hendrickx, J.M.H. Rapid Salinity Mapping by Electromagnetic Induction for Determining Riparian Restoration Potential. Restor. Ecol. 1994, 2, 242–246. [Google Scholar] [CrossRef]
  42. Triantafilis, J.; Huckel, A.I.; Mcbratney, A.B. Use of a Mobile Electromagnetic Sensing System for Assessment of Soil Salinity and Irrigation Efficiency. In Proceedings of the 16th World Conference of Soil Science, Montpelliers, France, 20–26 August 1998. [Google Scholar]
  43. Triantafilis, J.; Laslett, G.M.; McBratney, A.B. Calibrating an electromagnetic induction instrument to measure salinity in soil under irrigated cotton. Soil Sci. Soc. Am. J. 2000, 64, 1009–1017. [Google Scholar] [CrossRef]
  44. Triantafilis, J.; Huckel, A.I.; Odeh, I.O.A. Comparison of statistical prediction methods for estimating field-scale clay content using different combinations of ancillary variables. Soil Sci. 2001, 166, 415–427. [Google Scholar] [CrossRef]
  45. Triantafilis, J.; Ahmed, M.F.; Odeh, I.O.A. Application of a mobile electromagnetic sensing system (MESS) to assess cause and management of soil salinization in an irrigated cotton-growing field. Soil Use Manag. 2002, 18, 330–339. [Google Scholar] [CrossRef]
  46. Triantafilis, J.; Huckel, A.I.; Odeh, I.O.A. Field-scale assessment of deep drainage risk. Irrig. Sci. 2003, 21, 183–192. [Google Scholar]
  47. Triantafilis, J.; Odeh, I.O.A.; Jarman, A.L.; Short, M.G.; Kokkoris, E. Estimating and mapping deep drainage risk at the district level in the lower Gwydir and Macquarie valleys, Australia. Aust. J. Exp. Agric. 2004, 44, 893–912. [Google Scholar] [CrossRef]
  48. Vaughan, P.J.; Lesch, S.M.; Corwin, D.L.; Cone, D.G. Water-Content Effect on Soil-Salinity Prediction—A Geostatistical Study Using Cokriging. Soil Sci. Soc. Am. J. 1995, 59, 1146–1156. [Google Scholar] [CrossRef]
  49. Vlotman, W.F. Calibrating the EM38. Available online: http://agris.fao.org/agris-search/search.do?recordID=NL2001003220 (accessed on 1 November 2017).
  50. Whiteley, R.J. Environmental geophysics: Challenges and perspectives. Explor. Geophys. 1994, 25, 189–196. [Google Scholar] [CrossRef]
  51. Williams, B.G.; Baker, G. An electromagnetic induction technique for reconnaissance surveys of soil salinity hazards. Soil Res. 1982, 20, 107–118. [Google Scholar] [CrossRef]
  52. Williams, B.G.; Fidler, F.T. The Use of Electromagnetic Induction for Locating Subsurface Saline Material. IAHS 1985, 189–196. [Google Scholar]
  53. Williams, B.; Braunach, M. The Detection of Subsurface Salinity within the Northern Slopes Region of Victoria, Australia. In Salinity in Watercourses and Reservoirs: Proceedings of the 1983 International Symposium on State-of-the-Art Control of Salinity, July 13-15, 1983, Salt Lake City, Utah; French, R.H., Ed.; Butterworth Publishers: Stoneham, MA, USA, 1984; pp. 515–524. [Google Scholar]
  54. Wittler, J.M.; Cardon, G.E.; Gates, T.K.; Cooper, C.A.; Sutherland, P.L. Calibration of electromagnetic induction for regional assessment of soil water salinity in an irrigated valley. J. Irrig. Drain. Eng.—ASCE 2006, 132, 436–444. [Google Scholar] [CrossRef]
  55. Wollenhaupt, N.C.; Richardson, J.L.; Foss, J.E.; Doll, E.C. A Rapid Method for Estimating Weighted Soil-Salinity from Apparent Soil Electrical-Conductivity Measured with an Aboveground Electromagnetic Induction Meter. Can. J. Soil Sci. 1986, 66, 315–321. [Google Scholar] [CrossRef]
  56. Yao, R.J.; Yang, J.S.; Liu, G.M. Calibration of soil electromagnetic conductivity in inverted salinity profiles with an integration method. Pedosphere 2007, 17, 246–256. [Google Scholar] [CrossRef]
  57. Zhang, H.; Schroder, J.L.; Pittman, J.J.; Wang, J.J.; Payton, M.E. Soil salinity using saturated paste and 1:1 soil to water extracts. Soil Sci. Soc. Am. J. 2005, 69, 1146–1151. [Google Scholar] [CrossRef]
  58. Chaudhry, M.R.B.A. Electromagnetic Induction Device (EM38) Calibration and Monitoring Soil Salinity/Environment (Pakistan); Vlotman, W.F., Ed.; EM38 Workshop: New Dehli, India, 2000; pp. 37–48. [Google Scholar]
  59. Amezketa, E. Soil salinity assessment using directed soil sampling from a geophysical survey with electromagnetic technology: A case study. Span. J. Agric. Res. 2007, 5, 91–101. [Google Scholar] [CrossRef]
  60. Arndt, J.L.; Prochnow, N.D.; Richardson, J.L. Estimating Weighted Soil Salinity of Medium Textured Soils in Eastern North DAKOTA with an Aboveground Electromagnetic Induction Meter; Department of Soil Science, North Dakota State University: Fargo, ND, USA, 1987. [Google Scholar]
  61. Bakker, D.; Hamilton, G.; Hetherington, R.; Spann, C. Productivity of waterlogged and salt-affected land in a Mediterranean climate using bed-furrow systems. Field Crops Res. 2010, 117, 24–37. [Google Scholar] [CrossRef]
  62. Corwin, D.L.; Lesch, S.M. A simplified regional-scale electromagnetic induction—Salinity calibration model using ANOCOVA modeling techniques. Geoderma 2014, 230–231, 288–295. [Google Scholar] [CrossRef]
  63. Akramkhanov, A.; Brus, D.J.; Walvoort, D.J.J. Geostatistical monitoring of soil salinity in Uzbekistan by repeated EMI surveys. Geoderma 2014, 213, 600–607. [Google Scholar] [CrossRef]
  64. Barbiero, L.; Cunnac, S.; Mane, L.; Laperrousaz, C.; Hammecker, C.; Maeght, J.L. Salt distribution in the Senegal middle valley—Analysis of a saline structure on planned irrigation schemes from N’Galenka creek. Agric. Water Manag. 2001, 46, 201–213. [Google Scholar]
  65. Bennett, D.L.; George, R.J.; Whitfield, B. The use of ground EM systems to accurately assess salt store and help define land management options, for salinity management. Explor. Geophys. 2000, 31, 249–254. [Google Scholar] [CrossRef]
  66. Broadfoot, K.; Morris, M.; Stevens, D.; Heuperman, A. The role of EM38 in land and water management planning on the Tragowel Plains in Northern Victoria. Explor. Geophys. 2002, 33, 90–94. [Google Scholar] [CrossRef]
  67. Cameron, D.R.; Dejong, E.; Read, D.W.L.; Oosterveld, M. Mapping Salinity Using Resistivity and Electromagnetic Inductive Techniques. Can. J. Soil Sci. 1981, 61, 67–78. [Google Scholar] [CrossRef]
  68. Bourgault, G.; Journel, A.G.; Rhoades, J.D.; Corwin, D.L.; Lesch, S.M. Geostatistical analysis of a soil salinity data set. Adv. Agronomy 1996, 58, 241–292. [Google Scholar]
  69. De Clercq, W.; Rozanov, A. Using Iodine as a Tracer in the Field and the Detection Thereof to Reflect on Water and Salt Movement in These Soils. In Proceedings of the 3rd Global Workshop on Proximal Soil Sensing, Postdam, Germany, 26–29 May 2013; p. 252. [Google Scholar]
  70. Fitzpatrick, R.W.; Thomas, M.; Davies, P.J.; Williams, B.G. Dry Saline Land: An Investigation Using Ground-Based Geophysics, Soil Survey and Spatial Methods near Jamestown, South Australia; CSIRO Land and Water: Glen Osmond, Australia, 2003. [Google Scholar]
  71. Hendrickx, J.M.H.; Baerends, B.; Raza, Z.I.; Sadig, M.; Chaudhry, M.A. Soil-Salinity Assessment by Electromagnetic Induction of Irrigated Land. Soil Sci. Soc. Am. J. 1992, 56, 1933–1941. [Google Scholar] [CrossRef]
  72. Hopkins, D.G.; Richardson, J.L. Detecting a salinity plume in an unconfined sandy aquifer and assessing secondary soil salinization using electromagnetic induction techniques, North Dakota, USA. Hydrogeol. J. 1999, 7, 380–392. [Google Scholar] [CrossRef]
  73. Huang, J.; Subasinghe, R.; Malik, R.; Triantafilis, J. Salinity hazard and risk mapping of point source salinisation using proximally sensed electromagnetic instruments. Comput. Electron. Agric. 2015, 113, 213–224. [Google Scholar] [CrossRef]
  74. Huang, J.; Mokhtari, A.; Cohen, D.; Monteiro Santos, F.; Triantafilis, J. Modelling soil salinity across a gilgai landscape by inversion of EM38 and EM31 data. Eur. J. Soil Sci. 2015, 66, 951–960. [Google Scholar] [CrossRef]
  75. Lesch, S.M.; Rhoades, J.D.; Corwin, D.L. Statistical Modeling and Prediction Methodologies for Large Scale Spatial Sil Salinity Characterization: A Case Study Using Calibrated Electromagnetic Measurements Within the Broadview Water District. Available online: https://pdfs.semanticscholar.org/d741/83bae61378577de128283ea0139f9f0a73dc.pdf (accessed on 1 November 2017).
  76. Lesch, S.M.; Strauss, D.J.; Rhoades, J.D. Spatial Prediction of Soil-Salinity Using Electromagnetic Induction Techniques 1. Statistical Prediction Models—A Comparison of Multiple Linear-Regression and Cokriging. Water Resour. Res. 1995, 31, 373–386. [Google Scholar] [CrossRef]
  77. Lesch, S.M.; Herrero, J.; Rhoades, J.D. Monitoring for temporal changes in soil salinity using electromagnetic induction techniques. Soil Sci. Soc. Am. J. 1998, 62, 232–242. [Google Scholar] [CrossRef]
  78. Mankin, K.R.; Ewing, K.L.; Schrock, M.D.; Kluitenberg, G.J. Field measurement and mapping of soil salinity in saline seeps. In Proceedings of the ASAE International Meeting, Minneapolis, MN, USA, 10–14 August 1997. [Google Scholar]
  79. Mankin, K.; Koelliker, J. Hydrologic balance approach to saline seep remediation design. Appl. Eng. Agric. 2000, 16, 129–133. [Google Scholar] [CrossRef]
  80. Mankin, K.R.; Karthikeyan, R. Field assessment of saline seep remediation using electromagnetic induction. Trans. ASAE 2002, 45, 99–107. [Google Scholar] [CrossRef]
  81. Turnham, C. Using Electromagnetic Induction Methods to Measure Just a Bullet Point Agricultural Soil Salinity and Its Effects on Adjacent Native Vegetation in Western Australia. BSc. Thesis, Lancaster University, Lancashire, UK, 2003. [Google Scholar]
  82. Lelij, A.V.D. Use of an Electromagnetic Induction Measurement (Type EM-38) for Mapping of Soil Salinity; Water Resource Commission: Murrumbidgee, Australia, 1983; p. 22. [Google Scholar]
  83. Barr, N.F. Salinity Control, Water Reform and Structural Adjustment: The Tragowel Plains Irrigation District. Ph.D. Thesis, University of Melbourne, Melbourne, Australia, 1999. [Google Scholar]
  84. Bennett, D.; George, R. Using the EM38 to measure the effect of soil salinity on Eucalyptus globulus in south-western Australia. Agric. Water Manag. 1995, 27, 69–85. [Google Scholar] [CrossRef]
  85. Bouksila, F.; Persson, M.; Bahri, A.; Berndtsson, R. Electromagnetic induction prediction of soil salinity and groundwater properties in a Tunisian Saharan oasis. Hydrol. Sci. J. 2012, 57, 1473–1486. [Google Scholar] [CrossRef]
  86. Slavich, P.; Johnston, S. Sources of acidity and pathways of transport to the Belongil drainage system. Available online: http://www.wetlandcare.com.au/Content/templates/..%5C..%5Cdocs%5Creports%5CBelongil%20Working%20Papers.pdf#page=10 (accessed on 1 November 2017).
  87. Chaali, N.; Coppola, A.; Comegna, A.; Dragonetti, G. Assessment of Soil Electromagnetic Parameters and Their Variation with Soil Water, Salts: A Comparison among EMI and TDR Measuring Methods. In Proceedings of the EGU General Assembly Conference Abstracts, Vienna, Austria, 12–17 April 2015. [Google Scholar]
  88. Corwin, D.L.; Lesch, S.M.; Shouse, P.J.; Soppe, R.; Ayars, J.E. Identifying soil properties that influence cotton yield using soil sampling directed by apparent soil electrical conductivity. Agron. J. 2003, 95, 352–364. [Google Scholar] [CrossRef]
  89. Smitt, C.; Cox, J.; McEwan, K.; Davies, P.; Herczeg, A.; Walker, G. Salt Transport in the Bremer Hills, interpretation of Spatial Datasets for Salt Distribution; Technical Report 49/03; CSIRO Land and Water: Canberra, Austrialia; Available online: http://www.clw.csiro.au/publications/technical2003/tr49-03.pdf (accessed on 1 November 2017).
  90. Evans, T. Mapping vineyard salinity using electromagnetic surveys. Aust. Grapegrow. Winemak. 1998, 415, 20–21. [Google Scholar]
  91. Hanson, B.R.; Kaita, K. Response of electromagnetic conductivity meter to soil salinity and soil-water content. J. Irrig. Drain. Eng.—ASCE 1997, 123, 141–143. [Google Scholar] [CrossRef]
  92. Horney, R.D.; Taylor, B.; Munk, D.S.; Roberts, B.A.; Lesch, S.M.; Plant, R.E. Development of practical site-specific management methods for reclaiming salt-affected soil. Comput. Electron. Agric. 2005, 46, 379–397. [Google Scholar] [CrossRef]
  93. Spies, B.; Woodgate, P. Salinity Mapping Methods in the Australian Context. Available online: https://www.researchgate.net/profile/Peter_Woodgate/publication/242771344_Salinity_mapping_methods_in_the_Australian_context/links/5418c99f0cf2218008bf4575.pdf (accessed on 1 November 2017).
  94. Cannon, M.E.; McKenzie, R.C.; Lachapelle, G. Soil-Salinity Mapping with Electromagnetic Induction and Satellite-Based Navigation Methods. Can. J. Soil Sci. 1994, 74, 335–343. [Google Scholar] [CrossRef]
  95. Yao, R.; Yang, J. Quantitative evaluation of soil salinity and its spatial distribution using electromagnetic induction method. Agric. Water Manag. 2010, 97, 1961–1970. [Google Scholar] [CrossRef]
  96. Yao, Y.; Zhang, F.; Jiang, H. Research on model of soil salinization monitoring based on hyperspectral index and EM38. Spectrosc. Spectr. Anal. 2013, 33, 1658–1664. [Google Scholar]
  97. Corwin, D.L.; Lesch, S.M. Application of Soil Electrical Conductivity to Precision Agriculture: Theory, Principles, and Guidelines. Agron. J. 2003, 95, 455–471. [Google Scholar] [CrossRef]
  98. Soliman, A.S. Detecting Salinity in Early Stages Using Electromagnetic Survey and Multivariate Geostatistical Technique: A Case Study of Nong Suang District, Nakhon Ratchasima, Thailand. Available online: https://www.semanticscholar.org/paper/Detecting-salinity-in-early-stages-using-electroma-Soliman/b7c1d14212dcd1ab4f55b473aad5489771d7c56d (accessed on 1 November 2017).
  99. Norman, C.P.; Lyle, C.W.; Heuperman, A.F. Pyramid Hill Irrigation Area: Soil Salinity Survey May–June, 1988 (Victoria). Available online: http://agris.fao.org/agris-search/search.do?recordID=AU9430078 (accessed on 1 November 2017).
  100. Lesch, S.M.; Corwin, D.L.; Robinson, D.A. Apparent soil electrical conductivity mapping as an agricultural management tool in arid zone soils. Comput. Electron. Agric. 2005, 46, 351–378. [Google Scholar] [CrossRef]
  101. Salama, R.; Bartle, G.; Farrington, P.; Wilson, V. Basin geomorphological controls on the mechanism of recharge and discharge and its effect on salt storage and mobilization—Comparative study using geophysical surveys. J. Hydrol. 1994, 155, 1–26. [Google Scholar] [CrossRef]
  102. Zhu, Q.; Lin, H.; Doolittle, J. Repeated Electromagnetic Induction Surveys for Improved Soil Mapping in an Agricultural Landscape. Soil Sci. Soc. Am. J. 2010, 74, 1763–1774. [Google Scholar] [CrossRef]
  103. Bobert, J.; Schmidt, F.; Gebbers, R.; Selige, T.; Schmidhalter, U. Estimating Soil Moisture Distribution for Crop Management with Capacitance Probes, EM-38 and Digital Terrain Analysis. In Proceedings of the 3rd European Conference on Precision Agriculture, Montpellier, France, 16–20 June 2001; pp. 349–359. [Google Scholar]
  104. Vitharana, U.W.A.; Van Meirvenne, M.; Cockx, L.; Bourgeois, J. Identifying potential management zones in a layered soil using several sources of ancillary information. Soil Use Manag. 2006, 22, 405–413. [Google Scholar] [CrossRef]
  105. Lück, E.; Eisenreich, M.; Domsch, H.; Blumenstein, O. Geophysik für Landwirtschaft und Bodenkunde. In Geophysik für Landwirtschaft und Bodenkunde; Selbstverl. der Arbeitsgruppe Stoffdynamik in Geosystemen: Potsdam, Gemany, 2000. [Google Scholar]
  106. Bang, J. Characterization of Soil Spatial Variability for Sitespecific Management Using Soil Electrical Conductivity and Other Remotely Sensed Data. Ph.D. Thesis, North Carolina State University, Raleigh, NC, USA, 2005. [Google Scholar]
  107. Luck, E.; Gebbers, R.; Ruehlmann, J.; Spangenberg, U. Electrical conductivity mapping for precision farming. Near Surf. Geophys. 2009, 7, 15–25. [Google Scholar] [CrossRef]
  108. Heil, K.; Schmidhalter, U. Characterisation of soil texture variability using the apparent soil electrical conductivity at a highly variable site. Comput. Geosci.—UK 2012, 39, 98–110. [Google Scholar] [CrossRef]
  109. Mertens, F.M.; Paetzold, S.; Welp, G. Spatial heterogeneity of soil properties and its mapping with apparent electrical conductivity. J. Plant Nutr. Soil Sci. 2008, 171, 146–154. [Google Scholar] [CrossRef]
  110. Zhu, Q.; Lin, H.; Doolittle, J. Repeated electromagnetic induction surveys for determining subsurface hydrologic dynamics in an agricultural landscape. Soil Sci. Soc. Am. J. 2010, 74, 1750–1762. [Google Scholar] [CrossRef]
  111. Zhu, Q.; Lin, H. Comparing ordinary kriging and regression kriging for soil properties in contrasting landscapes. Pedosphere 2010, 20, 594–606. [Google Scholar] [CrossRef]
  112. Corwin, D.L.; Lesch, S.M. Characterizing soil spatial variability with apparent soil electrical conductivity: Part II. Case study. Comput. Electron. Agric. 2005, 46, 103–133. [Google Scholar] [CrossRef]
  113. McBratney, A.B.; Minasny, B.; Whelan, B.M. Obtaining ‘Useful’ High-Resolution Soil Data from Proximally-Sensed Electrical Conductivity/Resistivity (PSEC/R) Surveys. Precis. Agric. 2005, 5, 503–510. [Google Scholar]
  114. Minasny, B.; McBratney, A.B. Estimating the water retention shape parameter from sand and clay content. Soil Sci. Soc. Am. J. 2007, 71, 1105–1110. [Google Scholar] [CrossRef]
  115. Waine, T.W.; Blackmore, B.S.; Godwin, R.J. Mapping available water content and estimating soil textural class using electro-magnetic induction. Proc. EurAgEng 2000, Paper 00-SW-44. [Google Scholar]
  116. Domsch, H.; Giebel, A. Estimation of Soil Textural Features from Soil Electrical Conductivity Recorded Using the EM38. Precis. Agric. 2004, 5, 389–409. [Google Scholar] [CrossRef]
  117. Doolittle, J.A.; Indorante, S.J.; Potter, D.K.; Hefner, S.G.; McCauley, W.M. Comparing three geophysical tools for locating sand blows in alluvial soils of southeast Missouri. J. Soil Water Conserv. 2002, 57, 175–182. [Google Scholar]
  118. Harvey, O.R.; Morgan, C.L.S. Predicting Regional-Scale Soil Variability using a Single Calibrated Apparent Soil Electrical Conductivity Model. Soil Sci. Soc. Am. J. 2009, 73, 164–169. [Google Scholar] [CrossRef]
  119. Triantafilis, J.; Lesch, S.M. Mapping clay content variation using electromagnetic induction techniques. Comput. Electron. Agric. 2005, 46, 203–237. [Google Scholar] [CrossRef]
  120. Kühn, J.; Brenning, A.; Wehrhan, M.; Koszinski, S.; Sommer, M. Interpretation of electrical conductivity patterns by soil properties and geological maps for precision agriculture. Precis. Agric. 2009, 10, 490–507. [Google Scholar] [CrossRef]
  121. Weller, U.; Zipprich, M.; Sommer, M.; Castell, W.Z.; Wehrhan, M. Mapping clay content across boundaries at the landscape scale with electromagnetic induction. Soil Sci. Soc. Am. J. 2007, 71, 1740–1747. [Google Scholar] [CrossRef]
  122. Schmidhalter, U.; Zintel, A. Schätzung der räumlichen Variationen des Ton- und Wassergehaltes mit elektromagnetischer Induktion. Mitteilungen Deutschen Bodenkundlichen Gesellschaft 1999, 91, 871–874. [Google Scholar]
  123. Schmidhalter, U.A.; Zintel, A.; Neudecker, E. Calibration of Electromagnetic Induction Measurements to Survey the Spatial Variability of Soils. In Proceedings of the 3rd European Conference on Precision Agriculture, Montpellier, France, 18–20 June 2001; pp. 479–484. [Google Scholar]
  124. Delin, S.; Soderstrom, M. Performance of soil electrical conductivity and different methods for mapping soil data from a small dataset. Acta Agric. Scand. Sect. B—Soil Plant Sci. 2003, 52, 127–135. [Google Scholar] [CrossRef]
  125. Korsaeth, A. Soil apparent electrical conductivity (ECa) as a means of monitoring changesin soil inorganic n on heterogeneous morainic soils in SE Norway during two growing seasons. Nutr. Cycl. Agroecosyst. 2005, 72, 213–227. [Google Scholar] [CrossRef]
  126. Korsaeth, A. Relations between Electrical Conductivity, Soil Texture and Chemical Properties on a Clay Soil in Southern Norway; DIAS Report, Plant Production No. 100; Apelsvoll Research Centre, The Norwegian Crop Research Institute: Kapp, Norway, September 2003; pp. 139–142. [Google Scholar]
  127. Lukas, V.; Neudert, L.; Kren, J. Mapping of soil conditions in precision agriculture. Acta Agrophys. 2009, 13, 393–405. [Google Scholar]
  128. Nehmdahl, H.; Greve, M.H. Using Soil Electrical Conductivity Measurements for Delineating Management Zones on Highly Variable Soils in Denmark. In Proceedings of the 3rd European Conference on Precision Agriculture, Montpellier, France, 18–20 June 2001; pp. 461–466. [Google Scholar]
  129. Saey, T.; Van Meirvenne, M.; Vermeersch, H.; Ameloot, N.; Cockx, L. A pedotransfer function to evaluate the soil profile textural heterogeneity using proximally sensed apparent electrical conductivity. Geoderma 2009, 150, 389–395. [Google Scholar] [CrossRef]
  130. Corwin, D.L.; Kaffka, S.R.; Hopmans, J.W.; Mori, Y.; van Groenigen, J.W.; van Kessel, C.; Lesch, S.M.; Oster, J.D. Assessment and field-scale mapping of soil quality properties of a saline-sodic soil. Geoderma 2003, 114, 231–259. [Google Scholar] [CrossRef]
  131. Grigera, M.S.; Drijber, R.A.; Eskridge, K.M.; Wienhold, B.J. Soil microbial biomass relationships with organic matter fractions in a Nebraska corn field mapped using apparent electrical conductivity. Soil Sci. Soc. Am. J. 2006, 70, 1480–1488. [Google Scholar] [CrossRef]
  132. Jung, W.K.; Kitchen, N.R.; Sudduth, K.A.; Kremer, R.J.; Motavalli, P.P. Relationship of apparent soil electrical conductivity to claypan soil properties. Soil Sci. Soc. Am. J. 2005, 69, 883–892. [Google Scholar] [CrossRef]
  133. Sudduth, K.A.; Kitchen, N.R. Mapping Soil Electrical Conductivity. Remote Sens. Agric. Environ. 2004, 188–201. [Google Scholar]
  134. Triantafilis, J.; Odeh, I.O.A.; McBratney, A.B. Five geostatistical models to predict soil salinity from electromagnetic induction data across irrigated cotton. Soil Sci. Soc. Am. J. 2001, 65, 869–878. [Google Scholar] [CrossRef]
  135. Hedley, C.B.; Yule, I.Y.; Eastwood, C.R.; Shepherd, T.G.; Arnold, G. Rapid identification of soil textural and management zones using electromagnetic induction sensing of soils. Aust. J. Soil Res. 2004, 42, 389–400. [Google Scholar] [CrossRef]
  136. Van Meirvenne, M.; Verdoodt, A.; Lenoir, H.; Saey, T.; Haputanthri, T. Response of EMI based proximal soil sensor in two contrasting tropical landscapes. In Proceedings of the 3rd Global Workshop on Proximal Soil Sensing, Postdam, Germany, 26–29 May 2013; p. 56. [Google Scholar]
  137. Dalgaard, M.; Have, H.; Nehmdahl, H. Soil clay mapping by measurement of electromagnetic conductivity. In Proceedings of the 3rd European Conference on Precision Agriculture, Montpellier, France, 18–20 June 2001; pp. 367–372. [Google Scholar]
  138. Brevik, E.C.; Fenton, T.E. Influence of soil water content, clay, temperature, and carbonate minerals on electrical conductivity readings taken with an EM-38. Soil Horiz. 2002, 43, 9–13. [Google Scholar] [CrossRef]
  139. Brevik, E.C.; Fenton, T.E.; Lazari, A. Soil electrical conductivity as a function of soil water content and implications for soil mapping. Precis. Agric. 2006, 7, 393–404. [Google Scholar] [CrossRef]
  140. Hedley, C.; Roudier, P.; Yule, I.; Ekanayake, J.; Bradbury, S. Soil water status and water table depth modelling using electromagnetic surveys for precision irrigation scheduling. Geoderma 2013, 199, 22–29. [Google Scholar] [CrossRef]
  141. Kachanoski, R.; Wesenbeeck, I.V.; Gregorich, E. Estimating spatial variations of soil water content using noncontacting electromagnetic inductive methods. Can. J. Soil Sci. 1988, 68, 715–722. [Google Scholar] [CrossRef]
  142. Kachanoski, R.; Wesenbeeck, I.V.; Jong, E.D. Field scale patterns of soil water storage from non-contacting measurements of bulk electrical conductivity. Can. J. Soil Sci. 1990, 70, 537–542. [Google Scholar] [CrossRef]
  143. Khakural, B.R.; Robert, P.C.; Hugins, D.R. Use of non-contacting electromagnetic inductive method for estimating soil moisture across a landscape. Commun. Soil Sci. Plant Anal. 1998, 29, 2055–2065. [Google Scholar] [CrossRef]
  144. Kravchenko, A.; Bollero, G.; Omonode, R.; Bullock, D. Quantitative mapping of soil drainage classes using topographical data and soil electrical conductivity. Soil Sci. Soc. Am. J. 2002, 66, 235–243. [Google Scholar] [CrossRef]
  145. Malo, D.D.; Lee, D.K.; Lee, J.H.; Christopherson, C.M. Soil Mositure, Bulk Densityy, Soil Temperature, and Soil Sensor (Veris 3100® and Geonics Em-38®) Relationships: Part1—Moody County Site; Progress report; South Dakota State University: Brookings, SD, USA, 2001. [Google Scholar]
  146. Jiang, P.; Anderson, S.H.; Kitchen, N.R.; Sudduth, K.A.; Sadler, E.J. Estimating plant-available water capacity for claypan landscapes using apparent electrical conductivity. Soil Sci. Soc. Am. J. 2007, 71, 1902–1908. [Google Scholar] [CrossRef]
  147. Morgan, C.; Norman, J.; Wolkowski, R.; Lowery, B.; Morgan, G.; Schuler, R. Two Approaches to Mapping Plant Available Water: EM-38 Measurements and Inverse Yield Modeling. In Proceedings of the 5th International Conference on Precision Agriculture, Bloomington, MI, USA, 16–19 July 2000; pp. 1–13. [Google Scholar]
  148. Reedy, R.C.; Scanlon, B.R. Soil Water Content Monitoring Using Electromagnetic Induction. J. Geotech. Geoenviron. Eng. 2003, 129, 1028–1039. [Google Scholar] [CrossRef]
  149. Reedy, R.C.; Scanlon, B.R. Assessing the Impact of Land Use on Groundwater Recharge in the Southern High Plains. In Proceedings of the 2003 AGU Fall Meeting Abstracts, San Francisco, CA, USA., 8–12 December 2003. [Google Scholar]
  150. Sheets, K.R.; Hendrickx, J.M.H. Noninvasive Soil Water Content Measurement Using Electromagnetic Induction. Water Resour. Res. 1995, 31, 2401–2409. [Google Scholar] [CrossRef]
  151. Erindi-kati, A. Remote Sensing and Root Zone Soil Moisture; McGill University: Montreal, QC, Canada, 2005. [Google Scholar]
  152. Wong, M.T.F.; Asseng, S. Determining the causes of spatial and temporal variability of wheat yields at sub-field scale using a new method of upscaling a crop model. Plant Soil 2006, 283, 203–215. [Google Scholar] [CrossRef]
  153. Misra, R.K.; Padhi, J. Assessing field-scale soil water distribution with electromagnetic induction method. J. Hydrol. 2014, 516, 200–209. [Google Scholar] [CrossRef]
  154. De Benedetto, D.; Castrignanò, A.; Rinaldi, M.; Ruggieri, S.; Santoro, F.; Figorito, B.; Gualano, S.; Diacono, M.; Tamborrino, R. An approach for delineating homogeneous zones by using multi-sensor data. Geoderma 2013, 199, 117–127. [Google Scholar] [CrossRef]
  155. Malo, D.D.; Lee, D.K.; Lee, J.H.; Christopherson, S.M.; Cole, C.M.; Kleinjan, J.L.; Carlson, C.G.; Clay, D.E.; Chang, J.; Reese, C.L.; et al. Soil Moisture, Bulk Density, Soil Temperature, and Soil Sensor (Veris 3100® And Geonics EM-38®) Moody County Site; Annual Report Soil PR00-41; South Dakota State Univertisy: Brookings, SD, USA, 2000. [Google Scholar]
  156. Buchanan, S.; Triantafilis, J. Mapping Water Table Depth Using Geophysical and Environmental Variables. Ground Water 2009, 47, 80–96. [Google Scholar] [CrossRef] [PubMed]
  157. Doolittle, J.; Noble, C.; Leinard, B. An electromagnetic induction survey of a riparian area in southwest Montana. Soil Horiz. 2000, 41, 27–36. [Google Scholar] [CrossRef]
  158. Fenton, T.E.; Lauterbach, M.A. Soil Map Unit Composition and Scale of Mapping Related to Interpretations for Precision Soil and Crop Management in Iowa. In Proceedings of the 4th International Conference on Precision Agriculture, St. Paul, MN, USA, 19–22 July 1998. [Google Scholar]
  159. Hall, L.M.; Brainard, J.R.; Bowman, R.S.; Hendrickx, J.M.H. Determination of solute distributions in the vadose zone using downhole electromagnetic induction. Vadose Zone J. 2004, 3, 1207–1214. [Google Scholar] [CrossRef]
  160. Schumann, A.; Zaman, Q. Mapping water table depth by electromagnetic induction. Appl. Eng. Agric. 2003, 19, 675–688. [Google Scholar] [CrossRef]
  161. Wilson, R.C.; Freeland, R.S.; Wilkerson, J.B.; Yoder, R.E. Imaging the Lateral Migration of Subsurface Moisture using Electromagnetic Induction. In Proceedings of the American Society of Agricultural and Biological Engineers Annual International Meeting, Chicago, IL, USA, 28–31 July 2002; p. 16. [Google Scholar]
  162. Wilson, R.C.; Freeland, R.S.; Wilkerson, J.B.; Yoder, R.E. Inferring subsurface morphology from transient soil moisture patterns using electrical conductivity. Trans. ASAE 2003, 46, 1435–1441. [Google Scholar] [CrossRef]
  163. Vervoort, R.W.; Annen, Y.L. Palaeochannels in Northern New South Wales: Inversion of electromagnetic induction data to infer hydrologically relevant stratigraphy. Aust. J. Soil Res. 2006, 44, 35–45. [Google Scholar] [CrossRef]
  164. Rhoades, J.D.; Lesch, S.M.; LeMert, R.D.; Alves, W.J. Assessing irrigation/drainage/salinity management using spatially referenced salinity measurements. Agric. Water Manag. 1997, 35, 147–165. [Google Scholar] [CrossRef]
  165. Hedley, C.; Yule, I. Soil water status mapping and two variable-rate irrigation scenarios. Precis. Agric. 2009, 10, 342–355. [Google Scholar] [CrossRef]
  166. Weaver, T.; Hulugalle, N.; Ghadiri, H. Estimating drainage under cotton with chloride mass balance and an EM38. Commun. Soil Sci. Plant Anal. 2013, 44, 1700–1707. [Google Scholar] [CrossRef]
  167. Corwin, D.L.; Lesch, S.M. Characterizing soil spatial variability with apparent soil electrical conductivity I. Survey protocols. Comput. Electron. Agric. 2005, 46, 103–133. [Google Scholar] [CrossRef]
  168. Saey, T.; Simpson, D.; Vermeersch, H.; Cockx, L.; Van Meirvenne, M. Comparing the EM38DD and DUALEM-21S Sensors for Depth-to-Clay Mapping. Soil Sci. Soc. Am. J. 2009, 73, 7–12. [Google Scholar] [CrossRef]
  169. Sherlock, M.D.; McDonnell, J.J. A new tool for hillslope hydrologists: Spatially distributed groundwater level and soilwater content measured using electromagnetic induction. Hydrol. Process. 2003, 17, 1965–1977. [Google Scholar] [CrossRef]
  170. Cook, S.; Adams, M.; Corner, R. On-farm experimentation to determine site-specific responses to variable inputs. Precis. Agric. 1999, 611–621. [Google Scholar]
  171. Ammons, J.T.; Timpson, M.E.; Newton, D.L. Application of aboveground electromagnetic conductivity meter to separate Natraqalfs and Ochraqualfs in Gibson County. Soil Surv. Horiz. 1989, 30, 66–70. [Google Scholar] [CrossRef]
  172. Anderson-Cook, C.M.; Alley, M.; Roygard, J.; Khosla, R.; Noble, R.; Doolittle, J. Differentiating soil types using electromagnetic conductivity and crop yield maps. Soil Sci. Soc. Am. J. 2002, 66, 1562–1570. [Google Scholar] [CrossRef]
  173. Brevik, E.C.; Fenton, T.E.; Jaynes, D.B. The use of soil electrical conductivity to investigate soil homogeneity in Story County, Iowa, USA. Soil Horiz. 2012, 53, 50–54. [Google Scholar] [CrossRef]
  174. Dampney, P.; King, J.; Lark, R.; Wheeler, H.; Bradley, R.; Mayr, T. Automated methods for mapping patterns of soil physical properties as a basis for variable management of crops within fields. In Precision Agriculture; Stafford, J., Werner, A., Eds.; Wageningen Academic Publishers: Wageningen, The Netherlands, 2003; pp. 135–140. [Google Scholar]
  175. Greve, M.B.; Greve, M.H. Decision Support System for Classification and Representation of Fuzzy Soil Boundaries. In Proceedings of the 19th European and Scandinavian Conference for ESRI Users, Copenhagen, Denmark, 8–10 November 2004. [Google Scholar]
  176. Hinck, S.; Mueller, K.; Emeis, N. Part Field Management: Comparison of EC-value, soil texture, nutrient content and biomass in two selected fields. In Proceedings of the 3rd Global Workshop on Proximal Soil Sensing, Postdam, Germany, 26–29 May 2013; p. 270. [Google Scholar]
  177. Huang, J.; Nhan, T.; Wong, V.N.L.; Johnston, S.G.; Lark, R.M.; Triantafilis, J. Digital soil mapping of a coastal acid sulfate soil landscape. Soil Res. 2014, 52, 327–339. [Google Scholar] [CrossRef]
  178. James, I.T.; Waine, T.W.; Bradley, R.I.; Taylor, J.C.; Godwin, R.J. Determination of soil type boundaries using electromagnetic induction scanning techniques. Biosyst. Eng. 2003, 86, 421–430. [Google Scholar] [CrossRef]
  179. Rampant, P.; Abuzar, M. Geophysical Tools and Digital Elevation Models: Tools for Understanding Crop Yield and Soil Variability. In Proceedings of the SuperSoil 2004—3rd Australian New Zealand Soils Conference, Sydney, Australia, 5–9 December 2004. [Google Scholar]
  180. Triantafilis, J. Hydrostratigraphic Analysis Using Electromagnetic Induction Data and a Quasi-Three-Dimensional Electrical Conductivity Imaging. In Proceedings of the 3rd Global Workshop on Proximal Soil Sensing, Potsdam, Gemany, 26–29 May 2013; pp. 34–38. [Google Scholar]
  181. Stroh, J.C.; Archer, S.; Doolittle, J.A.; Wilding, L. Detection of edaphic discontinuities with ground-penetrating radar and electromagnetic induction. Landsc. Ecol. 2001, 16, 377–390. [Google Scholar] [CrossRef]
  182. Bönecke, E.; Franko, U. A Modelling Approach to Find Stable and Reliable Soil Organic Carbon Values for Further Regionalization. In Proceedings of the EGU General Assembly Conference Abstracts, Vienna, Austria, 12–17 April 2015. [Google Scholar]
  183. Doolittle, J.A.; Collins, M.E. A comparison of EM induction and GPR methods in areas of karst. Geoderma 1998, 85, 83–102. [Google Scholar] [CrossRef]
  184. Doolittle, J.A.; Sudduth, K.A.; Kitchen, N.R.; Indorante, S.J. Estimating Depths to Claypans Using Electromagnetic Induction Methods. J. Soil Water Conserv. 1994, 49, 572–575. [Google Scholar]
  185. Gebbers, R.; Lück, E.; Heil, K. Depth sounding with the EM38-detection of soil layering by inversion of apparent electrical conductivity measurements. Precis. Agric. 2007, 7, 95–102. [Google Scholar]
  186. Kimble, J.M.; Doolittle, J.; Taylor, R.; Windhorn, R.; Gerken, J. The Use of EMI and Electrical Instruments for Estimating Soil Properties to Help in Mapping. In Proceedings of the 2001 AGU Fall Meeting Abstract, San Francisco, CA, USA, 10–14 December 2001. [Google Scholar]
  187. Kitchen, N.R.; Sudduth, K.A.; Drummond, S.T. Mapping of sand deposition from 1993 midwest floods with electromagnetic induction measurements. J. Soil Water Conserv. 1996, 51, 336–340. [Google Scholar]
  188. Knotters, M.; Brus, D.J.; Voshaar, J.H.O. A Comparison of Kriging, Co-Kriging and Kriging Combined with Regression for Spatial Interpolation of Horizon Depth with Censored Observations. Geoderma 1995, 67, 227–246. [Google Scholar] [CrossRef]
  189. Vitharana, U.W.A.; Saey, T.; Cockx, L.; Simpson, D.; Vermeersch, H.; Van Meirvenne, M. Upgrading a 1/20,000 soil map with an apparent electrical conductivity survey. Geoderma 2008, 148, 107–112. [Google Scholar] [CrossRef]
  190. Boettinger, J.; Doolittle, J.; West, N.; Bork, E.; Schupp, E.W. Nondestructive assessment of rangeland soil depth to petrocalcic horizon using electromagnetic induction. Arid Land Res. Manag. 1997, 11, 375–390. [Google Scholar] [CrossRef]
  191. Bork, E.W.; West, N.E.; Doolittle, J.A.; Boettinger, J.L. Soil depth assessment of sagebrush grazing treatments using electromagnetic induction. J. Range Manag. 1998, 51, 469–474. [Google Scholar] [CrossRef]
  192. Brevik, E.; Fenton, T.; Horton, R. Effect of Daily Soil Temperature Fluctuations on Soil Electrical Conductivity as Measured with the Geonics® EM-38. Precis. Agric. 2004, 5, 145–152. [Google Scholar] [CrossRef]
  193. Brus, D.J.; Knotters, M.; Vandooremolen, W.A.; Vankernebeek, P.; Vanseeters, R.J.M. The Use of Electromagnetic Measurements of Apparent Soil Electrical-Conductivity to Predict the Boulder Clay Depth. Geoderma 1992, 55, 79–93. [Google Scholar] [CrossRef]
  194. Cai, C.; Lin, J.; Meng, F.; Sun, Y.; Li, D. Estimation of topsoil thickness in reclaimed field using EM38. Trans. Chin. Soc. Agric. Eng. 2010, 26, 319–323. [Google Scholar]
  195. Grellier, S.; Florsch, N.; Camerlynck, C.; Janeau, J.L.; Podwojewski, P.; Lorentz, S. The use of Slingram EM38 data for topsoil and subsoil geoelectrical characterization with a Bayesian inversion. Geoderma 2013, 200, 140–155. [Google Scholar] [CrossRef]
  196. Sudduth, K.A.; Kitchen, N.R.; Hughes, D.F.; Drummond, S.T. Electromagnetic Induction Sensing as an Indicator of Productivity on Claypan Soils. In Site-Specific Management for Agricultural Systems; American Society of Agronomy, Crop Science Society of America, Soil Science Society of America: Madison, WI, USA, 1995; pp. 671–681. [Google Scholar]
  197. Sudduth, K.A.; Kitchen, N.R.; Bollero, G.A.; Bullock, D.G.; Wiebold, W.J. Comparison of electromagnetic induction and direct sensing of soil electrical conductivity. Agron. J. 2003, 95, 472–482. [Google Scholar] [CrossRef]
  198. Sudduth, K.; Kitchen, N. Electromagnetic Induction Sensing of Claypan Depth. Am. Soc. Agric. Eng. 1993. [Google Scholar]
  199. Corwin, D.L. Delineating Site-Specific Crop Management Units: Precision Agriculture Application in GIS. In Proceedings of the 2005 ESRI International Users Conference, San Diego, CA, USA, 25–29 July 2005. [Google Scholar]
  200. Triantafilis, J.; Santos, F.M. 2-dimensional soil and vadose-zone representation using an EM38 and EM34 and a laterally constrained inversion model. Soil Res. 2010, 47, 809–820. [Google Scholar] [CrossRef]
  201. McBride, R.A.; Gordon, A.M.; Shrive, S.C. Estimating Forest Soil Quality from Terrain Measurements of Apparent Electrical-Conductivity. Soil Sci. Soc. Am. J. 1990, 54, 290–293. [Google Scholar] [CrossRef]
  202. Lück, E. Conductivity Mapping to Characterize the Spatial Variability within Large Fields. In Proceedings of the 6th International Conference on Precision Agriculture and Other Precision Resources Management, Minneapolis, MN, USA, 14–17 July 2002. [Google Scholar]
  203. Martinez, G.; Vanderlinden, K.; Ordóñez, R.; Muriel, J.L. Can Apparent Electrical Conductivity Improve the Spatial Characterization of Soil Organic Carbon? All rights reserved. No part of this periodical may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Vadose Zone J. 2009, 8, 586–593. [Google Scholar]
  204. Johnson, C.K.; Eskridge, K.M.; Corwin, D.L. Apparent soil electrical conductivity: Applications for designing and evaluating field-scale experiments. Comput. Electron. Agric. 2005, 46, 181–202. [Google Scholar] [CrossRef]
  205. Islam, M.M.; Saey, T.; De Smedt, P.; Van De Vijver, E.; Delefortrie, S.; Van Meirvenne, M. Modeling within field variation of the compaction layer in a paddy rice field using a proximal soil sensing system. Soil Use Manag. 2014, 30, 99–108. [Google Scholar] [CrossRef]
  206. Islam, M.M.; Meerschman, E.; Saey, T.; De Smedt, P.; Van De Vijver, E.; Delefortrie, S.; Van Meirvenne, M. Characterizing compaction variability with an electromagnetic induction sensor in a puddled paddy rice field. Soil Sci. Soc. Am. J. 2014, 78, 579–588. [Google Scholar] [CrossRef]
  207. Jung, W.K.; Kitchen, N.R.; Sudduth, K.A.; Anderson, S.H. Spatial characteristics of claypan soil properties in an agricultural field. Soil Sci. Soc. Am. J. 2006, 70, 1387–1397. [Google Scholar] [CrossRef]
  208. Krajco, J. Detection of Soil Compaction Using Soil Electrical Conductivity. MSc. Thesis, Cranfield University, Cranfield, UK, 2007. Available online: https://dspace.lib.cranfield.ac.uk/bitstream/1826/2346/2/MSc%20Thesis%20final.pdf (accessed on 1 November 2017).
  209. Slavich, P.G.; Yang, J. Estimation of Field Scale Leaching Rates from Chloride Mass Balance and Electromagnetic Induction Measurements. Irrig. Sci. 1990, 11, 7–14. [Google Scholar] [CrossRef]
  210. Tarr, A.B.; Moore, K.J.; Bullock, D.G.; Dixon, P.M.; Burras, C.L. Improving Map Accuracy of Soil Variables Using Soil Electrical Conductivity as a Covariate. Precis. Agric. 2005, 6, 255–270. [Google Scholar] [CrossRef]
  211. Vasic, D.; Ambrus, D.; Bilas, V. Simple Linear Inversion of Soil Electromagnetic Properties from Analytical Model of Electromagnetic Induction Sensor. In Proceedings of the 2014 IEEE Sensors Applications Symposium (SAS), Queenstown, New Zealand, 18–20 February 2014; pp. 15–19. [Google Scholar]
  212. Hendrickx, J.M.H.; Borchers, B.; Corwin, D.L.; Lesch, S.M.; Hilgendorf, A.C.; Schlue, J. Inversion of soil conductivity profiles from electromagnetic induction measurements: Theory and experimental verification. Soil Sci. Soc. Am. J. 2002, 66, 673–685. [Google Scholar] [CrossRef]
  213. Kitchen, N.; Sudduth, K.; Drummond, S. Soil electrical conductivity as a crop productivity measure for claypan soils. J. Prod. Agric. 1999, 12, 607–617. [Google Scholar] [CrossRef]
  214. Noellsch, A.J. Optimizing Crop N Use Efficiency Using Polymer-Coated Urea and Other N Fertilizer Sources Across Landscapes with Claypan Soils. MSc. Thesis, University of Missouri, Columbia, MS, USA, 2008. [Google Scholar]
  215. Eigenberg, R.A.; Nienaber, J.A. Electromagnetic survey of cornfield with repeated manure applications. J. Environ. Qual. 1998, 27, 1511–1515. [Google Scholar] [CrossRef]
  216. Eigenberg, R.A.; Nienaber, J.A. Identification of Nutrient Distribution at Abandoned Livestock Manure Handling Site Using Electromagnetic Induction. In Proceedings of the 2001 American Society of Agricultural and Biological Engineers Annual Meeting, Sacramento, CA, USA, 30 July–1 August 2001. [Google Scholar]
  217. Eigenberg, R.; Korthals, R.; Nienaber, J. Geophysical electromagnetic survey methods applied to agricultural waste sites. J. Environ. Qual. 1998, 27, 215–219. [Google Scholar] [CrossRef]
  218. Eigenberg, R.A.; Doran, J.W.; Nienaber, J.A.; Ferguson, R.B.; Woodbury, B.L. Electrical conductivity monitoring of soil condition and available N with animal manure and a cover crop. Agric. Ecosyst. Environ. 2002, 88, 183–193. [Google Scholar] [CrossRef]
  219. Eigenberg, R.A.; Nienaber, J.A. Electromagnetic induction methods applied to an abandoned manure handling site to determine nutrient buildup. J. Environ. Qual. 2003, 32, 1837–1843. [Google Scholar] [CrossRef] [PubMed]
  220. Eigenberg, R.; Nienaber, J. Soil conductivity map differences for monitoring temporal changes in an agronomic field. Am. Soc. Agric. Eng. Pap. 1999, 993173. [Google Scholar]
  221. Eigenberg, R.A.; Nienaber, J.A.; Woodbury, B.L.; Ferguson, R.B. Soil conductivity as a measure of soil and crop status—A four-year summary. Soil Sci. Soc. Am. J. 2006, 70, 1600–1611. [Google Scholar] [CrossRef]
  222. Stevens, R.; O’bric, C.; Carton, O. Estimating nutrient content of animal slurries using electrical conductivity. J. Agric. Sci. 1995, 125, 233–238. [Google Scholar] [CrossRef]
  223. Doran, J.W.; Parkin, T.B. Quantitative indicators of soil quality: A minimum data set. Soil Sci. Soc. Am. 1996. [Google Scholar]
  224. Fritz, R.; Malo, D.; Schumacher, T.; Clay, D.; Carlson, C.; Ellsbury, M.; Dalsted, K. Field comparison of two soil electrical conductivity measurement systems. Precis. Agric. 1999, 1211–1217. [Google Scholar]
  225. Ponitka, J.; Pößneck, J. Untersuchungen zur Teilflächenbewirtschaftung: Untersuchungen zur Anwendung ausgewählter teilflächenspezifischer Bewirtschaftungsmethoden am Beispiel eines Auenstandortes der Elbe. Available online: http://www.qucosa.de/fileadmin/data/qucosa/documents/1847/1165589915791-5456.pdf (accessed on 1 November 2017).
  226. Jaynes, D.B.; Colvin, T.S.; Ambuel, J. Yield mapping by electromagnetic induction. In Site-Specific Management for Agricultural Systems; American Society of Agronomy, Crop Science Society of America, Soil Science Society of America: Madison, WI, USA, 1995; pp. 383–394. [Google Scholar]
  227. Delin, S.; Lindén, B. Variations in Net Nitrogen Mineralisation within an Arable Field. Acta Agric. Scand. Sect. B Soil Plant Sci. 2002, 52, 78–85. [Google Scholar]
  228. Dunn, B.W.; Beecher, H.G. Using electro-magnetic induction technology to identify sampling sites for soil acidity assessment and to determine spatial variability of soil acidity in rice fields. Aust. J. Exp. Agric. 2007, 47, 208–214. [Google Scholar] [CrossRef]
  229. Heiniger, R.W.; McBride, R.G.; Clay, D.E. Using soil electrical conductivity to improve nutrient management. Agron. J. 2003, 95, 508–519. [Google Scholar] [CrossRef]
  230. Nadler, A. Estimating the soil water dependence of the electrical conductivity soil solution/electrical conductivity bulk soil ratio. Soil Sci. Soc. Am. J. 1982, 46, 722–726. [Google Scholar] [CrossRef]
  231. Lund, E.D.; Christy, C.D.; Drummond, P.E. Practical applications of soil electrical conductivity mapping. Available online: http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.512.2890&rep=rep1&type=pdf (accessed on 1 November 2017).
  232. Heiniger, R.W.; Carl, C. Side-by-Side Comparisons of Uniform and Site-Specific Nutrient Applications. In Precision Agriculture; American Society of Agronomy: Madison, WI, USA, 1999; p. 967. [Google Scholar]
  233. Motavalli, P.P.; Hammer, R.D.; Bardhan, S. Apparent soil electrical conductivity used to determine soil phosphorus variability in poultry litter-amended pastures. Yale Rev. Educ. Sci. 2015, 287–309. [Google Scholar] [CrossRef]
  234. Bekele, A.; Hudnall, W.H.; Daigle, J.J.; Prudente, J.A.; Wolcott, M. Scale dependent variability of soil electrical conductivity by indirect measures of soil properties. J. Terramech. 2005, 42, 339–351. [Google Scholar] [CrossRef]
  235. Zimmermann, H.M.; Plöchl, M.; Luckhaus, C.; Domsch, H. Selecting the optimum locations for soil investigations. Precis. Agric. 2003, 759–764. [Google Scholar]
  236. Lesch, S.M. Sensor-directed response surface sampling designs for characterizing spatial variation in soil properties. Comput. Electron. Agric. 2005, 46, 153–179. [Google Scholar] [CrossRef]
  237. Corwin, D.; Lesch, S.; Segal, E.; Skaggs, T.; Bradford, S. Comparison of Sampling Strategies for Characterizing Spatial Variability with Apparent Soil Electrical Conductivity Directed Soil Sampling. J. Environ. Eng. Geophys. 2010, 15, 147–162. [Google Scholar] [CrossRef]
  238. Heilig, J.; Kempenich, J.; Doolittle, J.; Brevik, E.C.; Ulmer, M. Evaluation of electromagnetic induction to characterize and map sodium-affected soils in the Northern Great Plains. Soil Horiz. 2011, 52, 77–88. [Google Scholar] [CrossRef]
  239. Johnson, C.K.; Doran, J.W.; Duke, H.R.; Wienhold, B.J.; Eskridge, K.M.; Shanahan, J.F. Field-scale electrical conductivity mapping for delineating soil condition. Soil Sci. Soc. Am. J. 2001, 65, 1829–1837. [Google Scholar] [CrossRef]
  240. Lesch, S.M.; Strauss, D.J.; Rhoades, J.D. Spatial Prediction of Soil-Salinity Using Electromagnetic Induction Techniques 2. An Efficient Spatial Sampling Algorithm Suitable for Multiple Linear-Regression Model Identification and Estimation. Water Resour. Res. 1995, 31, 387–398. [Google Scholar] [CrossRef]
  241. Lesch, S.; Rhoades, J.; Corwin, D. ESAP-95 version 2.01 R: User manual and tutorial guide. Res. Rep. 2000, 146, 17. [Google Scholar]
  242. Yao, R.; Yang, J.; Zhao, X.; Chen, X.; Han, J.; Li, X.; Liu, M.; Shao, H. A New Soil Sampling Design in Coastal Saline Region Using EM38 and VQT Method. Clean—Soil Air Water 2012, 40, 972–979. [Google Scholar] [CrossRef]
  243. Shaner, D.L.; Khosla, R.; Brodahl, M.K.; Buchleiter, G.W.; Farahani, H.J. How well does zone sampling based on soil electrical conductivity maps represent soil variability? Agron. J. 2008, 100, 1472–1480. [Google Scholar] [CrossRef]
  244. Box, G.E.; Draper, N.R. Empirical Model-Building and Response Surfaces; Wiley: New York, NY, USA, 1987; Volume 424. [Google Scholar]
  245. Tarr, A.B.; Moore, K.J.; Dixon, P.M.; Burras, C.L.; Wiedenhoeft, M.H. Use of soil electroconductivity in a multistage soil-sampling scheme. Crop Manag. 2003, 2. [Google Scholar] [CrossRef]
  246. Minasny, B.; McBratney, A.B. A conditioned Latin hypercube method for sampling in the presence of ancillary information. Comput. Geosci. 2006, 32, 1378–1388. [Google Scholar] [CrossRef]
  247. Niedzwiecki, J.D.G.; Pudelko, R. Electrical conductivity analysis of field of highly variable soils. In Proceedings of the 3rd Global Workshop on Proximal Soil Sensing, Potsdam, Germany, 26–29 May 2013. [Google Scholar]
  248. Clay, D.E.; Chang, J.; Malo, D.D.; Carlson, C.G.; Reese, C.; Clay, S.A.; Ellsbury, M.; Berg, B. Factors influencing spatial variability of soil apparent electrical conductivity. Commun. Soil Sci. Plant Anal. 2001, 32, 2993–3008. [Google Scholar] [CrossRef]
  249. Neudecker, E.; Schmidhalter, U.; Sperl, C.; Selige, T. Site-Specific Soil Mapping by Electromagnetic Induction. In Proceedings of the 3rd European Conference on Precision Agriculture, Montpellier, France, 16–20 June 2001; pp. 271–276. [Google Scholar]
  250. Bramley, R.G.V.; Trought, M.C.T.; Praat, J.P. Vineyard variability in Marlborough, New Zealand: characterising variation in vineyard performance and options for the implementation of Precision Viticulture. Aust. J. Grape Wine Res. 2011, 17, 72–78. [Google Scholar] [CrossRef]
  251. Bramley, R.G.V. Precision Viticulture—Tools to Optimise Winegrape Production in a Difficult Landscape. In Proceedings of the 6th International Conference on Precision Agriculture and Other Precision Resources Management, Minneapolis, MN, USA, 14–17 July 2002; p. 33. [Google Scholar]
  252. Chang, J.Y.; Clay, D.E.; Carlson, C.G.; Clay, S.A.; Malo, D.D.; Berg, R.; Kleinjan, J.; Wiebold, W. Different techniques to identify management zones impact nitrogen and phosphorus sampling variability. Agron. J. 2003, 95, 1550–1559. [Google Scholar] [CrossRef]
  253. Cockx, L.; Meirvenne, M.V.; Hofman, G. The Use of Electromagnetic Induction in Delineating Nitrogen Management Zones. In Proceedings of the 7th International Conference on Precision Agriculture and Other Precision Resources Management, Minneapolis, MN, USA, 25–28 July 2004. [Google Scholar]
  254. Cockx, L.; Van Meirvenne, M.; Hofman, G. Characterization of nitrogen dynamics in a pasture soil by electromagnetic induction. Biol. Fertil. Soils 2005, 42, 24–30. [Google Scholar] [CrossRef]
  255. Corwin, D. Past, present and future trends of soil soil electrical conductivity measurement using geophysical methods. In Handbook of Agricultural Geophysics; CRC Press: Boca Raton, FL, USA, 2008; pp. 17–44. [Google Scholar]
  256. Delin, S. Site-specific Nitrogen Fertilization Demand in Relation to Plant Available Soil Nitrogen and Water. Available online: https://pub.epsilon.slu.se/730/ (accessed on 1 November 2017).
  257. Domsch, H.; Kaiser, T.; Witzke, K.; Zauer, O. Empirical methods to detect management zones with respect to yield. Precis. Agric. 2003, 187–192. [Google Scholar]
  258. Fleming, K.L.; Westfall, D.G.; Bausch, W.C. Evaluating Management Zone Technology and Grid Soil Sampling for Variable Rate Nitrogen Application. In Proceedings of the 5th International Conference on Precision Agriculture, Bloomington, Minnesota, USA, 16–19 July 2000; p. 179. [Google Scholar]
  259. Fountas, S.; Anastasiou, E.; Xanthopoulos, G.; Lambrinos, G.; Manolopoulou, E.; Apostolidou, S.; Lentzou, D.; Tsiropoulos, Z.; Balafoutis, A. Precision agriculture in watermelons. In Precision Agriculture’15; Wageningen Academic Publishers: Wageningen, The Netherlands, 2015; pp. 399–403. [Google Scholar]
  260. Fraisse, C.W.; Sudduth, K.A.; Kitchen, N.R. Delineation of site-specific management zones by unsupervised classification of topographic attributes and soil electrical conductivity. Trans. ASAE 2001, 44, 155–166. [Google Scholar] [CrossRef]
  261. Franzen, D.; Kitchen, N. Developing Management Zones to Target Nitrogen Applications; Potash & Phosphate Institute: Atlanta, GA, USA, 1999. [Google Scholar]
  262. Fridgen, J.J.; Kitchen, N.R.; Sudduth, K.A.; Drummond, S.T.; Wiebold, W.J.; Fraisse, C.W. Management Zone Analyst (MZA): Software for subfield management zone delineation. Agron. J. 2004, 96, 100–108. [Google Scholar] [CrossRef]
  263. Guretzky, J.A.; Moore, K.J.; Burras, C.L.; Brummer, E.C. Distribution of legumes along gradients of slope and soil electrical conductivity in pastures. Agron. J. 2004, 96, 547–555. [Google Scholar] [CrossRef]
  264. Islam, M.M.; Saey, T.; Meerschman, E.; De Smedt, P.; Meeuws, F.; Van De Vijver, E.; Van Meirvenne, M. Delineating water management zones in a paddy rice field using a Floating Soil Sensing System. Agric. Water Manag. 2011, 102, 8–12. [Google Scholar] [CrossRef]
  265. Islam, M.; Cockx, L.; Meerschman, E.; De Smedt, P.; Meeuws, F.; Van Meirvenne, M. A floating sensing system to evaluate soil and crop variability within flooded paddy rice fields. Precis. Agric. 2011, 12, 850–859. [Google Scholar] [CrossRef]
  266. Jaynes, D.; Colvin, T.; Ambuel, J. Soil Type and Crop Yield Determination from Ground Conductivity Surveys. Am. Soc. Agric. Eng. 1993.
  267. Jaynes, D.B.; Colvin, T.S.; Kaspar, T.C. Identifying potential soybean management zones from multi-year yield data. Comp. Electron. Agric. 2005, 46, 309–327. [Google Scholar] [CrossRef]
  268. Kern, A.; Brevik, E.C.; Fenton, T.E.; Vincent, P.C. Comparisons of soil ECa maps to an order 1 soil survey for a Central Iowa field. Soil Horiz. 2008, 49, 36–39. [Google Scholar] [CrossRef]
  269. Kilborn, D.A.; Moore, K.J.; Hintz, R.L.; Tarr, A.B. Chariton Valley Biomass Project Task 5.10.0. Available online: http://www.iowaswitchgrass.com/__docs/pdf/5-10-0%20final%20report.pdf (accessed on 1 November 2017).
  270. Lamb, D.; Bramley, R.; Hall, A. Precision Viticulture-an Australian Perspective. In Proceedings of the XXVI International Horticultural Congress: Viticulture-Living with Limitations 640, Toronto, ON, Canada, 11–17 August 2002; pp. 15–25. [Google Scholar]
  271. Luchiari, A., Jr.; Shanahan, J.; Francis, D.; Schlemmer, M.; Schepers, J.; Liebig, M.; Schepers, A.; Payton, S. Strategies for Establishing Management Zones for Site Specific Nutrient Management. In Proceedings of the 5th International Conference on Precision Agriculture, Bloomington, MN, USA, 16–19 July 2000. [Google Scholar]
  272. McCormick, S.; Bailey, J.S.; Jordan, C.; Higgins, A. A potential role for electrical conductivity mapping in the site-specific management of grassland. Cent. Agric. Landsc. Land Use Res. 2003, 393. [Google Scholar]
  273. Oliver, Y.M.; Wong, M.T.F.; Robertson, M.J. Targeting the subsoil to better manage acidity spatially. In Proceedings of the 17th ASA conference, Hobart, Australia, 20–24 September 2015; Available online: http://2015.agronomyconference.com/papers/agronomy2015final00049.pdf (accessed on 1 November 2017).
  274. Proffitt, T.; Bramley, R. Further developments in precision viticulture and the use of spatial information in Australian vineyards. Aust. Vitic. 2010, 14, 31–39. [Google Scholar]
  275. Robinson, N.J.; Rampant, P.C.; Callinan, A.P.L.; Rab, M.A.; Fisher, P.D. Advances in precision agriculture in south-eastern Australia. II. Spatio-temporal prediction of crop yield using terrain derivatives and proximally sensed data. Crop Pasture Sci. 2009, 60, 859–869. [Google Scholar] [CrossRef]
  276. Saleh, A.; Belal, A.A. Delineation of site-specific management zones by fuzzy clustering of soil and topographic attributes: A case study of East Nile Delta, Egypt. IOP Conf. Ser.: Earth Environ. Sci. 2014, 18, 012046. [Google Scholar] [CrossRef]
  277. Schepers, A.R.; Shanahan, J.F.; Liebig, M.A.; Schepers, J.S.; Johnson, S.H.; Luchiari, A. Appropriateness of management zones for characterizing spatial variability of soil properties and irrigated corn yields across years. Agron. J. 2004, 96, 195–203. [Google Scholar] [CrossRef]
  278. Sun, Y.; Cheng, Q.; Lin, J.; Schellberg, J.; Schulze Lammers, P. Investigating soil physical properties and yield response in a grassland field using a dual-sensor penetrometer and EM38. J. Plant Nutr. Soil Sci. 2013, 176, 209–216. [Google Scholar] [CrossRef]
  279. Türker, U.; Talebpour, B.; Yegül, U. Determination of the relationship between apparent soil electrical conductivity with pomological properties and yield in different apple varieties. Žemdirbystė=Agric. 2011, 98, 307–314. [Google Scholar]
  280. Vanderlinden, K.; Martinez, G.; Giráldez, J.V.; Muriel, J.L. Characterizing Soil Management Systems using Electromagnetic Induction. In Proceedings of the 19th World Congress of Soil Science, Soil Solutions for a Changing World, Brisbane, Australia, 1–6 August 2010. [Google Scholar]
  281. Vitharana, U.W.A.; Van Meirvenne, M.; Simpson, D.; Cockx, L.; De Baerdemaeker, J. Key soil and topographic properties to delineate potential management classes for precision agriculture in the European loess area. Geoderma 2008, 143, 206–215. [Google Scholar] [CrossRef]
  282. Zhang, R.; Wienhold, B.J. The effect of soil moisture on mineral nitrogen, soil electrical conductivity, and pH. Nutr. Cycl. Agroecosyst. 2002, 63, 251–254. [Google Scholar] [CrossRef]
  283. McBratney, A.; Gruijter, J.D. A continuum approach to soil classification by modified fuzzy k-means with extragrades. Eur. J. Soil Sci. 1992, 43, 159–175. [Google Scholar] [CrossRef]
  284. Corwin, D.; Carrillo, M.; Vaughan, P.; Rhoades, J.; Cone, D. Evaluation of a GIS-linked model of salt loading to groundwater. J. Environ. Qual. 1999, 28, 471–480. [Google Scholar] [CrossRef]
  285. Kitchen, N.R.; Sudduth, K.A.; Myers, D.B.; Drummond, S.T.; Hong, S.Y. Delineating productivity zones on claypan soil fields using apparent soil electrical conductivity. Comput. Electron. Agric. 2005, 46, 285–308. [Google Scholar] [CrossRef]
  286. Abuzar, M.; Rampant, P.; Fisher, P. Measuring spatial variability of crops and soils at sub-paddock scale using remote sensing technologies. In Proceedings of the 2004 IEEE International Geoscience and Remote Sensing Symposium, Anchorage, AK, USA, 20–24 September 2004; pp. 1633–1636. [Google Scholar]
  287. Kravchenko, A.N.; Harrigan, T.M.; Bailey, B.B. Soil Electrical Conductivity as a Covariate to Improve the Efficiency of Field Experiments. Trans. ASAE 2005, 48, 1353–1357. [Google Scholar] [CrossRef]
  288. Lawes, R.A.; Bramley, R.G.V. A Simple Method for the Analysis of On-Farm Strip Trials. Agron. J. 2012, 104, 371–377. [Google Scholar] [CrossRef]
  289. Kravchenko, A.N.; Robertson, G.P.; Thelen, K.D.; Harwood, R.R. Management, Topographical, and Weather Effects on Spatial Variability of Crop Grain Yields. Agron. J. 2005, 97, 514–523. [Google Scholar] [CrossRef]
  290. Cosby, A.; Trotter, M.; Falzon, G.; Stanley, J.; Powell, K.; Schneider, D.; Lamb, D. Mapping redheaded cockchafer infestations in pastures—Are PA tools up to the job? In Precision Agriculture’13; Wageningen Academic Publishers: Wageningen, The Netherlands, 2013; pp. 585–592. [Google Scholar]
  291. Hbirkou, C.; Welp, G.; Rehbein, K.; Hillnhütter, C.; Daub, M.; Oliver, M.; Pätzold, S. The effect of soil heterogeneity on the spatial distribution of Heterodera schachtii within sugar beet fields. Appl. Soil Ecol. 2011, 51, 25–34. [Google Scholar] [CrossRef]
  292. Jaynes, D.B.; Novak, J.M.; Moorman, T.B.; Cambardella, C. Estimating herbicide partition coefficients from electromagnetic induction measurements. J. Environ. Qual. 1995, 24, 36–41. [Google Scholar] [CrossRef]
  293. Olesen, J.E.; Jørgensen, L.N.; Jensen, P.K.; Thomsen, A.G.; Jensen, J.E. Sensor-Based Graduation of Fungicide Application in Winter Wheat. Available online: https://www2.mst.dk/udgiv/publications/2008/978-87-7052-701-9/pdf/978-87-7052-702-6.pdf (accessed on 1 November 2017).
  294. Ritter, C.; Dicke, D.; Weis, M.; Oebel, H.; Piepho, H.P.; Büchse, A.; Gerhards, R. An on-farm approach to quantify yield variation and to derive decision rules for site-specific weed management. Precis. Agric. 2008, 9, 133–146. [Google Scholar] [CrossRef]
  295. Bevan, B. The search for graves. Geophysics 1991, 56, 1310–1319. [Google Scholar] [CrossRef]
  296. Dalan, R.A.; Bevan, B.W. Geophysical indicators of culturally emplaced soils and sediments. Geoarchaeol. Int. J. 2002, 17, 779–810. [Google Scholar] [CrossRef]
  297. Dalan, R.A. Remote Sensing in Archaeology: An Explicitly North American Perspective. In Magnetic Susceptibility; Johnson, J.K., Ed.; University of Alabama Press: Tuscaloosa, AL, USA, 2006; pp. 161–203. [Google Scholar]
  298. Ferguson, R.B. The search for port la joye: Archaeology at Ile Saint-Jeans first French settlement. Island Mag. 1990, 27, 3–8. [Google Scholar]
  299. Santos, V.R.N.; Porsani, J.L.; Mendonça, C.A.; Rodrigues, S.I.; DeBlasis, P.D. Reduction of topography effect in inductive electromagnetic profiles: Application on coastal sambaqui (shell mound) archaeological site in Santa Catarina state, Brazil. J. Archaeol. Sci. 2009, 36, 2089–2095. [Google Scholar] [CrossRef]
  300. Simpson, D.; Lehouck, A.; Verdonck, L.; Vermeersch, H.; Van Meirvenne, M.; Bourgeois, J.; Thoen, E.; Docter, R. Comparison between electromagnetic induction and fluxgate gradiometer measurements on the buried remains of a 17th century castle. J. Appl. Geophys. 2009, 68, 294–300. [Google Scholar] [CrossRef]
  301. Simpson, D.; Lehouck, A.; Van Meirvenne, M.; Bourgeois, J.; Thoen, E.; Vervloet, J. Geoarchaeological prospection of a Medieval manor in the Dutch polders using an electromagnetic induction sensor in combination with soil augerings. Geoarchaeol.—Int. J. 2008, 23, 305–319. [Google Scholar] [CrossRef]
  302. Viberg, A.; Trinks, I.; Lidén, K. Archaeological Prospection in the Swedish Mountain Tundra Region; Presses Universitaires de Rennes: Rennes, France, 2009. [Google Scholar]
  303. Bevan, B.W. The Search for Graves. Geophysics 1991, 56, 1310–1319. [Google Scholar] [CrossRef]
  304. McNeill, J.D. The application of electromagnetic techniques to environmental geophysical surveys. Geol. Soc. Lond. Eng. Geol. Spec. Publ. 1997, 12, 103–112. [Google Scholar] [CrossRef]
Figure 1. (Left) Relative cumulative contribution vs depth for vertically (RV(z)) and horizontally (RH(z)) orientated dipoles; (Right) Comparison of the relative responses for vertically (FV(z)) and horizontally (FH(z)) oriented dipoles.
Figure 1. (Left) Relative cumulative contribution vs depth for vertically (RV(z)) and horizontally (RH(z)) orientated dipoles; (Right) Comparison of the relative responses for vertically (FV(z)) and horizontally (FH(z)) oriented dipoles.
Sensors 17 02540 g001
Figure 2. Mounting of the EM38 on a metal-free sledge pulled by a tractor (constructed after Corwin and Lesch [7]).
Figure 2. Mounting of the EM38 on a metal-free sledge pulled by a tractor (constructed after Corwin and Lesch [7]).
Sensors 17 02540 g002
Figure 3. Procedure of interpolation of ECa across field boundaries. (a) Lanes of ECa –measurements with EM38 on arable farmland (16.9 ha); (b) Lanes of ECa –measurements (field-by-field means (mfield) were subtracted from individual observations); (c) Interpolation 5 m × 5 m grid of ECa (residuals); (d) Interpolation 5 m × 5 m grid of ECa (residuals+local means).
Figure 3. Procedure of interpolation of ECa across field boundaries. (a) Lanes of ECa –measurements with EM38 on arable farmland (16.9 ha); (b) Lanes of ECa –measurements (field-by-field means (mfield) were subtracted from individual observations); (c) Interpolation 5 m × 5 m grid of ECa (residuals); (d) Interpolation 5 m × 5 m grid of ECa (residuals+local means).
Sensors 17 02540 g003aSensors 17 02540 g003b
Figure 4. Relationships between ECa and multi-annual mean of yield (wheat) of the long-term experiment Dürnast 020 in dependence of fertilization level (control plots: no fertilizer, fertilized plots (low): 100–140 kg ha−1 N, fertilized plots (high): 150–180 kg ha−1 N).
Figure 4. Relationships between ECa and multi-annual mean of yield (wheat) of the long-term experiment Dürnast 020 in dependence of fertilization level (control plots: no fertilizer, fertilized plots (low): 100–140 kg ha−1 N, fertilized plots (high): 150–180 kg ha−1 N).
Sensors 17 02540 g004
Table 1. Overview with literature of relationships between EM38-ECa and salinity.
Table 1. Overview with literature of relationships between EM38-ECa and salinity.
StudyParametersLocation of Investigation
Derivation of salinity with ECa and ECe
[8]ECa and ECe relationships: classifying salt affected areaCalifornia, USA
[9]Descriptions and formulations of ECe and ECa; mathematical coefficients;South Australia
[10,11]Descriptions and formulations of ECe and ECa; inverted salinity profiles;South California, USA
[12]ECa and ECsaturated extract, Na, Cl, Salinity maps with relation to yield Barley)North-east Australia
[13]Calibration ECe and ECav, ECahMissouri, USA
[14]ECa and EC1:5 relationships to perform growth of Australian tree species on saline sitesQueensland, Australia
[15]Formulations of ECe and ECaEgypt
[16]Relationship ECa and ECe, ECa observations on establishing and growth of perennial pasture speciesAustralia
[17]Salinity contour maps with ECe and ECav, ECahNnortheast Spain
[18]Salinity classification system based on EC1:5 with groups of degradesHenan, China
[19]Formulations of ECe and ECaCalifornia, USA
[20]ECa, ECe to apply site specific management tech. on saline sitesCalifornia, USA
[21]ECe and ECav, ECah advanced calibrations reduce soil sampling from 200-300 to 36,California, USA
[5,22]Site calibration ECe and ECav, ECahSaskatchewan, Canada
[23,24,25,26]Formulations of ECa and ECe; Salt tolerance of trees, forages, crops and turf grasses; survival and growth of eucalyptus and pastures in saline soils.Alberta, Canada
[27]Exchangeable sodium percentage and ECe in relation to ECaIllinois, USA
[28]Soil survey with salinity regions; relationship ECe and ECa to detect salinity of irrigated districtsAragon, Spain
[29]Ranges of ECa as classification system of saline areasVictoria, Australia
[30]Salinity classification system based on ranges of total dissolved salt concentrations, EC1:5 with groups of crops with different tolerances to rootzone salinityVictoria, Australia
[31,32,33,34]Descriptions and formulations of ECa, ECe, ECp and EC ratios; multiple regression coefficients;California, USA
[35]Relationships of ECe and ECa, Soil salinity maps of different depth intervals and salinity profile maps at upstream and downstream of the field bordersYazd Province, Iran
[36]Monitoring spill of liquid manure occurred a few years agoManitoba, Canada
[37]Formulations of ECe and ECa (India)India (different regions)
[38,39]Descriptions and formulations of ECe and ECa; modeled coefficients;NSW, Australia
[40]Comparison EC1:5 - ECe and ECa to detect salinity in an early stageNakhon Ratchasima, Thailand
[41]Comparison ECe and ECa to detect salinityNew Mexico, USA
[42,43,44,45]Determination ECe profiles with ECa (EM38 and EM31); geostatistical methods to predict salinity from ECa (EM38 and EM31), comparison calibration approaches;NSW, Queensland, Australia
[46,47]Ratio (EM38/EM31) sampling points to determine deep drainage and leaching fraction, ECa and ECe; ECa and clay; ECa and deep drainage;NSW, Australia
[48]ECe, water content and ECah, combined with cokrigingCalifornia, USA
[49]Descriptions, formulations, classifications of ECa, ECe, ECp and EC ratios
[50]Overview salinity and determination
[51,52,53]Detection subsurface saline materialVictoria, Australia
[54]Calibration models ECe and ECa and water content over regional scaleColorado, USA
[55]Descriptions and formulations of ECe and ECa, simple depth weighted coefficients;North Dakota, USA
[56]Depthwise calibration models ECav, ECah and ECe and EC1:5 to construct inverted salinity profilesJiangsu, China
[57]Comparison saturated paste and 1:1 soil to water extractsOklahoma, Texas, USA
[58]Formulations of ECe and ECaPakistan
[59]Site calibration ECe and ECav, ECahNavarre, Spain
[60]Site calibration ECe and ECav, ECahNorth Dakota, USA
[61]Site calibration EC(1:5) and ECahWest Australia
[62]Salinity calibration model to simulate ECe from ECaCalifornia, Minnesota, USA
[57]Comparison saturated paste and 1:1 soil to water extractsOklahoma, Texas, USA
Construction of salinity maps
[63]Interpolation methods of ECa;
ECa maps as base for salinity maps/ECe)
Uzbekistan
[64]Relation ECa-topography-salinity extensionSenegal
[65]ECa-salinity areasSE Australia
[66]Salinity maps with stepwise data processingVictoria, Australia
[67]Mapping salinity with EM38, EM31 and Wenner arrayAlberta, Canada
[68]Geostatistical analysis of soil salinity data–––––––
[69]Salinity distribution within a field and combination with iodine tracer studyCape Province, South Africa
[70]Soil salinity maps with ECa, in relation to land use and soil/geologySouth Australia
[71]ECa and visual agronomic survey of salinityPunjab, Pakistan
[72]Mapping of salinity plume in a sandy aquiferNorth Dakota, USA
[73]Detecting salt stores and evaluation of the risk of salinisationNSW, Australia
[74]ECa maps by inverting data collected at various heights in the EM4SOIL softwareYazd Province, Iran
[75]Salinity characteristics with PCACalifornia, USA
[76]Comparison of multiple linear regression and cokrigingCalifornia, USA
[77]Temporal changes in salinity using ECaAragon, Spain
[78,79,80]Saline seep mapping and remediation; comparison salinity (ECe) and ECa of different conductivity tools;
saline seep mechanism in combination with hydrological modeling
Kansas, USA
[81]Comparison salinity (ECa) between different land useAustralia
[82]EM38 field wiseNSW, Australia
Salinity and field management
[83]Assessment of salinity by farmersAustralia
[84]Effect of salinity on eucalyptus treesSE Australia
[85]Soil salinity and groundwater propertiesTunisia
[86]Extension of groundwater acidityNSW, Australia
[87]EM38 and TDR: comparison of measuring methods-
[88]Assessment of soil quality properties with ECaCalifornia, USA
[89]ECa distribution in the landscape and as a consequence of evapotranspiration and phreatic riseSouth Australia
[90]Salinity in vineyardsAustralia
[91]ECa–salinity–water contentCalifornia, USA
[92]Salinity management in cotton fieldsCalifornia, USA
EM38 in combination with other sensors
[93]Comparison tools and methods detection salinityAustralia
[94]EM38 in combination with satellite-based navigation methodsAlberta, Canada
[95]Increasing precision of salinity with EM38 and EM31 (both ECah) at various layersYellow River Delta, China
[96]Hyperspectral data related to different soil salinization extent was combined with ECa order to establish a soil salinization monitoring modelWeigan River, China
Table 2. References indicating relationships between EM38-ECa and soil texture.
Table 2. References indicating relationships between EM38-ECa and soil texture.
StudyTextureTexture Content (%)ECa (mS m−1)R2Location of Investigations
Europe
[103]Clay
Silt
Silt + Clay
not describedECav: 10–1100.28/0.53 *
0.14/0.49 *
0.25/0.71 *
* with extracting
TWI-trend
Wulfen, Kassow, East Germany
[116]Clay Silt4–16
7–36
ECav: 3–30ECav: 0.55 (clay)
ECav: 0.67 (clay + silt) (after factor scoring)
Brandenburg, Berlin, Germany
[120]Clay2–60ECav: mean 13–92ECav: 0.56Saxony-Anhalt, Germany
[121]Clay2–45ECav: 2–80ECa: 0.66
ECa corr: 0.85, corrected across field boundaries with neighbors regression
Bavaria, Germany
[122]Clay6–42ECav, ECah: 6–36ECav: 0.08–0.38
ECah: 0.13–0.33
Scheyern, Germany
[123]Clay7––32ECav: 8–44
ECah: 6-41
ECav: 0.21–0.44
ECah: 0.13–0.67
Scheyern, Germany
Silt4–53ECav: 8-44
ECah: 6–41
ECav: 0.11–0.46
ECah: 0.01–0.60
Sand28–79ECav: 8–44
ECah: 6–41
ECav: 0.04–0.38
ECah: 0.13-0.69
[109]Clay
Silt
Sand
2–25
5–69
5–50
ECav: 5–65ECav: 0.76–0.76
ECav: 0.65–0.71
ECav: 0.00–0.69
3 fields around Bonn, Germany
[108]Clay3–48ECav: 2–99
ECah: 5–77
ECav: 0.76
ECah: 0.74
South Germany
Silt4–71ECav: 2–99
ECah: 5–77
ECav: 0.67
ECah: 0.67
Sand + gravel15–67ECav: 2–99
ECah: 5–77
ECav: 0.76
ECah: 0.74
[124]Clay5–30ECav: 9 (mean)ECav: 0.94Southwest Sweden
[125]Clay9–24ECav: 4
ECah: 32.2
approximate
values
two depths:
0–25 cm, 25–60 cm and 2 fields
ECav: 0.19–0.41
ECah: 0.32–0.45
South Norway
Silt28–49ECav: 0.006–0.52
ECah: 0.002–0.56
Sand33–61ECav: 0.01–0.4
ECah: 0.02–0.44
Gravel3–11ECav: 0.05–0.94
ECah: 0.08–0.94
[126]Clayabout 5–40ECah: 6–26ECah: 0.63South Norway
[127]Clay
Sand
23–44
39–67
ECav: 0–50ECav: 0.55
ECav: 0.41
Moravia, Czech Republic
[128]Clay4–24 ECav: 0.49–0.67 (different dates on the same field)Jütland, Denmark
[129]Clay2–56ECav: 9–106
ECah: 5–97
ECav: 0.81East-Flanders, Belgium
[104]Claytopsoil: 14–24
subsoil: 3–27
ECav: 18–47
ECav: 12–36
(ECav* ECah)0.5: 0.69 subsoil
(ECav* ECah)0.5: 0.16 topsoil
Flanders, Belgium
North America
[106]Clay10–46 (mean values)ECav: 1–54
ECah: 1–56
ECav–30 cm: about 0.5
ECah–30 cm: 0.3–0.56
North Carolina, USA
Silt20–35 (mean values)ECav: 1–54
ECah: 1–56
ECav–30 cm: 0.4–0.6
ECah–30 cm: −0.3–0.56
Sand40–70 (mean values)ECav: 1–54
ECah: 1–56
ECav–30 cm: about 0.4
ECah–30 cm: −0.3–−0.6
[130]Clay
Silt
Sand
24–44
26–51
8–50
ECav , ECah: about 40, salinity affected0.08
0.18
0.14
ln of geometric mean of ECav and ECahCalifornia, USA
[112]Clay3–48about ECav, ECah 10-65ECav: 0.11.
ECah: 0.08
Western California, USA
[131]Clay14–29ECav: 19–35
ECah: 14–26
ECav: 0.69
ECah: 0.66
Nebraska, USA
[118]Clay12–32ECav: 19–1180.7612 sites in Texas, USA
[132]Clay13–63ECav: 30–65
ECah: 38–83
ECav–30 cm: 0.55
ECah–30 cm: 0.55
Central Missouri, USA
Silt33–81ECav: 30–65
ECah: 38–83
ECav–30 cm: 0.55
ECah–30 cm: 0.55
Sand6–11ECav: 30–65
ECah: 38–83
ECav–30 cm: 0.27
ECah–30 cm: 0.27
[3]Clay
Silt
13–36
31–67
ECav: 7–37ECav: 0.55
ECav: 0.15 and 0.48 (2 fields)
North-central states, USA
[133]Clay
Silt
Sand
about 5–40
unknown
unknown
ECav:about 5–60ECav: 0.36–0.77
ECav: 0.27–0.71
ECav: 0.21–0.36
Midwest USA
[100]Clay
Sand
10–32
52–85
ECav: 84.8
ECah: 40.1
ECah: 0.76
ECah: 0.74
Southwest USA
Australasia
[42,45]Clayabout 30–85ECav:80–200
(salt affected)
ECav 0.62 and 0.64NSW, Australia
[134]Clayabout 40–65ECav:30–210ECav: 0.72NSW, Australia
[119]Clay15–58ECav: 5–159
ECah: 13–147
ECav: 0.66
ECah: 0.67
combination of EM34 and EM38 in different modes:0.79
NSW, Australia
[135]Clayabout 20–45about 10–36ECav: 0.72
ECah: 0.65
Manavata, New Zealand
Asia
[136]Clay
Silt
Sand
1.5–41.3
6.5–33.5
45.8–91.0
ECav: 1–40topsoil: 0.47 (on average)Sri Lanka
Unknown
[137]Clay12–20ECav: 7–20
ECah: 7–15
ECav: 0.78
ECah: 0.80
Not described
Table 3. Literature describing relationships between EM38-ECa and parameters of soil water.
Table 3. Literature describing relationships between EM38-ECa and parameters of soil water.
StudyParametersLocation of Investigations
Water content
[138]Water contentIowa, USA
[139]Water contentIowa, USA
[112]Water contentSouth California, USA
[91]Water contentCalifornia, USA
[140]Water content, water table depthNew Zealand
[141,142]Water contentOntario, Canada
[143]Water storage [mm]Minnesota, USA
[144]Soil drainage classesIllinois, USA
[145]Soil water content (θv, θw), ±3%South Dakota, USA
[146]Plant available water contentMissouri, USA
[147]Water contentColumbia County, USA
[148,149]Volumetric water contentTexas, USA
[122]Water content: ECav: 0.39; ECah: 0.26
Plant available water content: ECav: 0.31; ECah: 0.29
Bavaria, Germany
[123]Water content ECav: 0.04–0.26; ECah: 0.16–0.64Bavaria, Germany
[150]Water contentFlorida, USA
[3]Water contentNorth-central USA
[151]Water content with EM38 and ASD spectrometerQuebec, Canada
[102]Repeated ECa measurements for determining water contentPennsylvania, USA
[152]Detection of available water content from ECa, for using in the yield software ADSIMWA, Australia
[153]Repeated ECa measurements and relation to water content (irrigation)Queensland, Australia
[115]Available water content and soil water deficit from texture finess classes and ECaCambridgeshire, UK
[154]ECa in combination with GPR to predict field wide water contentSouth-east Italy
[155]Soil water content, soil bulk densitySouth Dakota, USA
Groundwater, water table depth, water drainage
[156]Water table depth using geophysical and relief variablesDarling River, Australia
[9]Groundwater rechargeSouth Australia
[157]Depth to groundwater tableMontana, USA
[158]Soil drainage classesIowa, USA
[159]Characterizing of water and solute distributions in the vadose zone with readings of EM38 and borehole conductivity meterNew Mexico, USA
[160]Water table depthFlorida, USA
[161,162]Detection of areas with different water movementsTennessee, USA
[46]Deep drainage riskAustralia
[163]Hydraulic conductivity of palaeochannel in alluvial plainsNSW, Australia
[42,45]Deep drainage (mm/year) with a 4-parameter broken-stick model fitted to ECav beyond 120 cmAustralia
Irrigation
[164]Irrigation effectiveness/drainageCalifornia, US,
[165]ECa – soil available water holding capacity on two variable-rate irrigation scenariosNew Zealand
[166]ECa for quick assessment of deep drainage under irrigated conditions in the field.Australia
Table 4. Literature indicating derivations of soil types and patterns as well as further soil parameters from EM38-ECa.
Table 4. Literature indicating derivations of soil types and patterns as well as further soil parameters from EM38-ECa.
StudyInvestigation ObjectLocation of Investigation
Soil types
[171]Separation between Natraqualf and OchraqualfTennessee, USA
[172]Soil types, yield mapsVirginia, USA
[173]ECa to derive more homogeneous lacustrine-derived soilsIowa, USA
[174]Soil pattern as basis of management zonesEngland
[175]Soil boundariesDenmark
[158]Soil map unit boundaries, detection of inclusionsIowa, USA
[2]Refine and improvement of soil maps-
[176]Soil types with clusteranalysisElbe-Weser-region, Germany
[177]Detection of areas with sulfidic sediments and coastal acid sulfate soilsNSW, Australia
[128]Soil typesJütland, Denmark
[178]Soil boundaries between clay loam and sandy loam soilsCambridge, UK
[179]Soil types, in combination with terrain parameters and other sensorsNW Victoria, Australia
[102]Repeated ECa measurements for determining soil typesPennsylvania, USA
[180]Inversion of EM38 and EM34 sigma-a data to detect the areal distribution of soil typesDarling River, Australia
[181]Distinguishing between soils with cambic pedogenic horizons and argillic horizons; boundaries of soil map unitsTexas, USA
[182]Supporting delineation of spatial distribution of C contentHarz region, Germany
Soil depth to horizons/layers/discontinuities/borders
[183]Depth to limestone bedrock and clayey residuumFlorida, Pennsylvania, USA
[184]Depth of claypan soilsMissouri, USA
[185]Soil depth soundingEast, south Germany
[5]Soil depth soundingOntario, Canada
[186]Depth to sand and gravelUnknown
[187]Depth of sand depositionMissouri, USA
[188]Layer depth, ECa as auxiliary variableNorth Netherlands
[189]Depth of the Tertiary substratumFlanders, Belgium
[190]Soil depth to petrocalcic horizonUtah, USA
[191]Soil depth to bedrock (loess above basalt)Idaho, USA
[192]Bulk density and ECaIowa, USA
[193]Boulder clay depthNorth Netherlands
[194]Linear, negative relation between ECa and topsoil layer thicknessFuxin, China
[195]Bayesian method to map the clay content of the Bt horizon associated with the control of encroaching treesSouth Africa
[1,196,197,198]Depth to claypan soilsMissouri, USA
Further soil properties
[88]Soil properties and cotton yieldCalifornia, USA
[199]Soil properties and cotton yieldCalifornia, USA
[112]Water content, cation exchange capacity, cations and anions in saturation extract and exchangeable, B, Mo, pH, C, N,West California, USA
[132]Cation exchange capacity, C, N, P, soil enzyme, microbial biomass, hydr. Sat. K., bulk densityMissouri, USA
[3]Water content, cation exchange capacityNorth-central states, USA
[45]CEC in salt affected soilsNSW, Australia
[200]CEC in dependence of EM38, EM31, 3 remotely sensed (Red, Green and Blue spectral brightness), 2 trend surface (Easting and Northing) variablesNSW, Australia
[201]Exchangeable Ca, Mg, cation exchange capacityOntario, Canada
[124]ECa as a covariable in cokriging improved the prediction of pH, clay, SOMSweden
[202]ECa in relation to water content, yield, CEC, clay silt, organic matterBrandenburg, Saxony-Anhalt, Germany
[131]C, total dissolved solids, depth of topsoilNebraska, USA
[203]Soil organic carbon and classifing with fields normalized ECaAndalucia, Spain
[204]N-dymanics for management zonesNebraska, USA
[176]Precision agriculture: combination of ECa and soil parameters (clay, yield, plant available water)Mecklenburg, Germany
[205,206]Compaction in paddy rice fields by puddlingBangladesh
[207]ECa as subsidiary variable for interpolationMissouri, USA
[208]Soil compactionSilsoe, UK
[209]Relations leaching rates to ECaNSW, Australia
[210]ECa as subsidiary variable for interpolation of P, K, pH, organic matter and water contentIowa, USA
[211]Simple linear inversion of ECa to simulate magnetic susceptibility-
Table 5. Literature describing selection of areas for soil sampling with EM38-ECa.
Table 5. Literature describing selection of areas for soil sampling with EM38-ECa.
StudyInvestigation ObjectLocation of Investigation
[59]Soil sampling pointsEbro River, Spain
[199]Sampling designWest California, USA
[237]ECa base sampling design: response surface sampling design (RSSD), stratified random sampling design (SRSD)California, USA
[228]Soil sampling design pHNSW, Australi,
[238]Mapping sodium affected soilsGreat Plains, USA
[204,239]Soil sampling design, soil unitsWest California, USA
[100,236,240,241]Soil sampling designSouthwest USA
[115]Sampling design for loacation of neutron probe access tubesCambridgeshire, UK
[242]VQT method (variance quad-tree) in combination of relief data and ECaJiangsu Province, China,
[235]Optimum locations for soil investigationsBrandenburg, Germany
Table 6. Composition of literature with derivations of yield maps, management zones and selection of areas for fertilization with EM38-ECa.
Table 6. Composition of literature with derivations of yield maps, management zones and selection of areas for fertilization with EM38-ECa.
StudyInvestigation ObjectLocation of Investigation
[172]Yield maps, Soil types and ECaVirginia, USA
[106]ECa, NIR, elevation, slope with k-means clustering to define management zonesNorth Carolina, USA
[65]Help for define management options with ECaSW, Australia
[250]Development of predictors of vine yield from ECaNew Zealand
[251]Management zones in vinicultureClare Valley, Australia
[103]Relationship ECa crop yieldNorth, east Germany
[252]Management zones on soil NO3 and P sampling variabilitySouth Dakota, USA
[253,254]N-management zonesBelgium
[130,199,255]Soil properties and cotton yieldCalifornia, USA
[174]Soil pattern as basis of management zonesEngland
[12]Identifiing management classes with ECa (measured at high and low water content)North-east Australia
[154]Multi-sensor data (EM38, GPR, FieldSpec) to delineate homogeneous zonesItaly
[256]Relationships ECa, N-fertilizing demandSouthwest Sweden
[257]Relationship ECa crop yield , management zonesBrandenburg, Germany
[258]Establishing of management zones with Corg, clay, NO3, K, Zn, ECa, corn yield dataColorado, USA
[259]Correlations ECa with yield, sugar content, piercing force, Kramer energy in a single yearPeleponnese, Greece
[260]Relationship ECa crop yield, management zonesMissouri, USA
[261]Management zones and N applicationsMissouri, USA
[262]Management zones delineation softwareMissouri, USA
[224]ECa to predict NO3-concentrationDakota, USA
[131]ECa zonesNebraska, USA
[263]Distribution of legumes in pastures in dependence of ECa and slopeIowa, USA
[176]Soil types (derived from ECa) related to yield, K, MgElbe-Weser-region, Germany
[92]Management zones salt affected sitesCalifornia, USA
[264]Development of key properties for delineation management zonesNorth Belgium
[265]Management zones in a paddy rice field with ECaBangladesh
[226,266]Relationship ECa crop yieldIowa, USA
[267]Management zones with yield, elevation and ECaIowa, USA
[132]Relationship ECa crop yieldMissouri, USA
[268]ECa-maps to derive management zonesIowa, USA
[269]Relationship ECa crop yield, terrain attributesIowa, USA
[213]Relationship and classification ECa crop yieldNorth central Missouri, USA
[270]Managing and monitoring variability in vineyardsAustralia
[271]Management zones with yield, elevation, ECa, aerial photosNebraska, USA
[272]Site-specific management of grasslandIreland
[249]Comparison ECa – German national soil inventory (Bodenzahlen)Bavaria, Germany
[273]Lime applicationto reduce subsoil acidityWestern Australia
[225]Relationships ECa, N-fertilizing zonesSaxonia, Germany
[274]Senor application in viticultureAustralia
[275]Multiyear ECa – yield relationshipVictoria, Australia
[276]Delineation of site-specific management-zones with ECa and topographic parametersNile Delta, Egypt
[277]Data fusion (Terrian attributes, ECa, yield, aerial imagers)Minnesota, USA
[179]Yield zones, yield per year, in combination with terrain parameters and other sensorsNorth West Victoria, Australia
[164]Relationship ECa crop yieldBavaria, Germany
[196]Relationship ECa crop yieldMissouri, USA
[278]Relationship ECa − volumetric water content (−35 cm) – yieldNRW, Germany
[279]ECa and yield of applesAnkara, Turkey,
[42,45]Sampling points with ratio (ECav-EM38/ECa-EM31)NSW, Australia
[245]Management zones and multilevel sampling schemeCentral Iowa, USA
[280]Management zones with ECa relative differences (ϑij , Eq. 31)SW Spain
[104]Management zones (delineated mainly with subsoil clay from ((ECav* ECah).5)) delivered from ECa)Flanders, Belgium
[281]Characterization of soil variation by key variables: pH, ECa, organic matterFlanders, Belgium
[121]Interpolation of ECa across field boundariesBavaria, Germany
[282]EC and soil inorganic N (no EM38-ECa)Nebraska, USA
Table 7. Regressions between ECa and multi-annual mean of yield (wheat) of the long-term experiment Dürnast 020 in dependence of fertilization level (see Figure 1).
Table 7. Regressions between ECa and multi-annual mean of yield (wheat) of the long-term experiment Dürnast 020 in dependence of fertilization level (see Figure 1).
Yield (dt ha−1)ConfigurationNEquationR2 Significance
Control plotsVertical12101.33 − 1.411 × ECa0.67 ***
Horizontal1264.61 − 0.758 × ECa0.81 ***
Fertilized plots (low)Vertical42106.85 − 0.81 × ECa0.36 **
Horizontal4253.466 + 1.394 × ECa − 0.025 × ECa20.76 ***
Fertilized plots (high)Vertical42111.2 − 0.811 × ECa0.22 *
Horizontal4276.853 + 0.361 × ECa − 0.012 × ECa20.67 ***
n.s. > 0.05, * 0.05 ≥ p > 0.01, ** 0.01 ≥ p > 0.001, *** p ≤ 0.001.
Table 8. Applications of EM38-ECa for improving the efficiency of field experiments.
Table 8. Applications of EM38-ECa for improving the efficiency of field experiments.
StudyInvestigation ObjectLocation of Investigation
[173]ECa to derive more homogeneous lacustrine-derived soilsIowa, USA
[204]Classification parameter for block designCalifornia, USA
[287]P-content in a field experiment with different levels of manure applicationsMichigan, USA
[288]Comparison of yield between strip trials, partly ECa; simplified evaluation methodSouth, west Australia
Table 9. Simulation of the yield (1980–2012) with ANOVA and ANCOVA with the factors fertilizing level and fertilizer-no. and the covariates ECa and relief parameters.
Table 9. Simulation of the yield (1980–2012) with ANOVA and ANCOVA with the factors fertilizing level and fertilizer-no. and the covariates ECa and relief parameters.
Target Variable, YearsModel and EffectsSignificancePartial Eta-Square Adjusted R2RMSE (dt ha−1)
Yield (dt ha −1), mean 1980, 1983, 1986, 1989, 1992, 1995, 1998, 2001, 2004, 2007, 2010, 2012Adjusted model
Constant
Fertilization level
Fertilizer no.
Fertilization level*Fertilizer no.
0.008
0.000
0.000
0.414
0.971
0.313
0.998
0.258
0.081
0.018
0.183.26
Yield (dt ha −1)3, mean (1980, 1983, 1986, 1989, 1992, 1995, 1998, 2001, 2004, 2007, 2010, 2012Adjusted model
Constant
Fertilization level
Fertilizer no.
Fertilization level*
Fertilizer no.
ECa (EM38-h)^3
lg10(ECa (EM38-v))
Channelnetwork^3
TWI^3
0.000
0.007
0.000
0.000
0.145
0.000
0.000
0.001
0.024
0.904
0.106
0.764
0.341
0.131
0.275
0.276
0.144
0.075
0.881.29
Significance: n.s. > 0.05, * 0.05 ≥ p>0.01, ** 0.01 ≥ p > 0.001, *** p ≤ 0.001; Partial eta-square: Measure of sensitivity to the correlated independent variables; Adjusted R2: adjusted R2 (coefficient of determination).
Table 10. Additional applications of EM38-ECa in agriculture and horticulture.
Table 10. Additional applications of EM38-ECa in agriculture and horticulture.
StudyInvestigation ObjectLocation of Investigation
[234]Corg, K, pH, Bray-2 P,Louisiana, USA
[290]Detecting soil properties as indicators for population density of Redheaded cockchafer (Adoryphourus couloni)Victoria, Australia
[215,217,219]Specific ions that are associated with animal wasteNebraska, USA
[220]N decomposition, organic and artificial fertilizerNebraska, USA
[221]ECa as an indicator of N gains and losses, available N sufficiency for corn in early stage and NO3-N surplus after harvestNebraska, USA
[291]ECa as indicator for soil conditions which are prefered by Heterodera schachtiiNorth Rhine-Westphalia, Germany
[292]Herbicide partition coefficientsIowa, USA
[233]Variation in soil testing PMissouri, Oklahoma, USA
[293]Part of fungicide application models in combination with ratio vegetation indexDenmark
[294]Weed distribution, herbicide injury in dependency of ECaNorth Rhine-Westphalia, Germany
[222]NH4, K in animal slurriesIreland
Table 11. Additional applications of EM38-ECa in archaeology.
Table 11. Additional applications of EM38-ECa in archaeology.
StudyInvestigation ObjectLocation of Investigation
[295]Detection of graves with inphase and quadphase readingsMaryland, USA
[296,297]Prehistoric earthworks with measurements in inphase modeOhio, USA
[298]Metal objects from the 18th centuryCanada
[299]Removing of the effect of elevation on the distribution of ECa readingsSanta Catarina State, Brazil
[300]Comparison EM38 fluxgate gradiometerBelgium
[301]Medieval manor in the dutch poldersNetherlands
[302]Area prospection with EM38 and MS2DTundra region, Sweden

Share and Cite

MDPI and ACS Style

Heil, K.; Schmidhalter, U. The Application of EM38: Determination of Soil Parameters, Selection of Soil Sampling Points and Use in Agriculture and Archaeology. Sensors 2017, 17, 2540. https://doi.org/10.3390/s17112540

AMA Style

Heil K, Schmidhalter U. The Application of EM38: Determination of Soil Parameters, Selection of Soil Sampling Points and Use in Agriculture and Archaeology. Sensors. 2017; 17(11):2540. https://doi.org/10.3390/s17112540

Chicago/Turabian Style

Heil, Kurt, and Urs Schmidhalter. 2017. "The Application of EM38: Determination of Soil Parameters, Selection of Soil Sampling Points and Use in Agriculture and Archaeology" Sensors 17, no. 11: 2540. https://doi.org/10.3390/s17112540

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop