in .,.. _,,. . . .3... aw, hm... 11m? .. .i 5.35.555 3:; fit than». .Eogfifiu . u... vauxhbasfi. .. . l. I: 2 . .3 »n. 4 {Jrz «53.3%.. ~ . THESIS 4. {\fl) 3‘ i \ ' I illllllllllllllllllll lllllllllLllllll 3129 This is to certify that the dissertation entitled ACCURACY OF SOIL PROPERTY MAPS FOR SITE-SPECIFIC MANAGEMENT presented by Thomas G. Mueller has been accepted towards fulfillment of the requirements for Ph . D . degree in So 11 Management myé? Mam/Le Major professor Date December 16, 1998 MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MTE DUE DATE DUE DATE DUE use WWI-m.“ ACCURACY OF SOIL PROPERTY MAPS FOR SITE-SPECIFIC MANAGEMENT By Thomas G. Mueller A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Crop and Soil Science 1998 ABSTRACT ACCURACY OF SOIL PROPERTY MAPS FOR SITE-SPECIFIC MANAGEMENT By Thomas G. Mueller The accuracy of soil property maps for site-specific management may be inadequate at sampling intensities recommended by commercial agriculture. Since the success of site-specific fertilizer applications depends on the quality of soil property maps, it is critical for Michigan farmers who are adopting these practices to have an understanding of the accuracy associated with different soil sampling strategies, soil sampling intensities, and interpolation techniques. This thesis evaluates how grid sampling and interpolation schemes affected map accuracy based on measures of map error. The second objective was to evaluate how different interpolation techniques that incorporate terrain attributes affects the spatial predictions of soil properties and whether relative performance of these techniques is affected by the scale of soil sampling. In addition to the soil samples used for spatial interpolation, samples were collected to assess the quality of the predictions. Grid point sampling at the industry standard intensity (100 m regular grid), grid cell sampling (100 m grid cells) and directed sampling based on soil type were not adequate to produce accurate nutrient condition maps for this field even though most of the variables were spatially structured. Prediction efficiencies were 0.5 to 10.5 % greater for inverse distance weighted interpolation than for kriging using a distance exponent of 1.5 at the 30-m grid sampling intensity. At high sampling densities (30 m regular grid), interpolation methods that utilized terrain attributes had similar prediction errors to interpolation methods that did not utilize terrain attributes. At a lower sampling intensity (61 m regular grid), methods that utilized terrain attributes, especially multiple regression, were more accurate than methods that did not. At the 61 m grid, the RMSE for multiple regression was lower (3.3 g kg") than the RMSE for ordinary kriging (4.1 g kg’l). A 100 m grid was not of sufficient intensity to be used to create accurate maps of soil properties for site-specific nutrient management, enhancing spatial estimates with terrain attributes can reduced the number of samples required to create an accurate map. It was not necessary to use complex, time consuming geostatistical techniques to use terrain attributes because multiple regression was sufficient. Copyright by Thomas G. Mueller 1 998 This dissertation is dedicated to my wife, Angela, for her love and perseverance; to our parents, Philip, Camille, Peter, and Gabriela, for their support and encouragement, and to our children, Daniel and Christina, for the joy they have given us, ACKNOWLEDGMENTS I would like to thank Dr. Francis J. Pierce believing in me and for his friendship, guidance, encouragement, and financial support. I would also like to express my gratitude to Dr. Darryl Wamcke for serving as the co-chair of my graduate committee and for his helpful advice. I am very grateful to both Dr. G. Philip Robertson and Dr. Irvin E. Widders for their advice and for serving on my committee. Again, I would like to thank my entire committee for their precious time. A special thanks goes to the farmers, John Anibal and Patrick Feldpausch, for there generosity. Also thanks to Jim Walseth at Springhill Engineering for mapping elevation at one of the research locations. Also thanks to Steve Law at the NRCS office in Clinton County for his help soil sampling. I am grateful to Dr. Oliver Schabenberger for many hours of consultation. I appreciate the invaluable advice of Dr. Pierre Goovaerts from the University of Michigan. I am especially grateful to Dr. Jose E. Cora for his friendship and collaboration. I am also grateful to for the technical assistance of Brian Long. I also would like to thank Cal Bricker for his advice and friendship. Thanks to Brian Long, Cal Bricker, and Gary Zehr for their help in the field near Kalamazoo. I appreciate the friendship and technical assistance of Brian Baer, Sven Bohm, Dr. Dave Harris, Dr. Samira Daroub, Brian Graff, and Elaine Parker. Thanks also goes to Dan Brown for teaching two very useful classes in GIS and for being so helpful. A special thanks goes to Demetrios Gatziolis for help with GIS and his friendship. I also appreciate the help of Jon Dahl and Vickie Smith from the Michigan State University soil testing laboratory for always being so helpful with research and teaching soil samples. vi Thanks for all of the hard work' to the students I have supervised: Curtis Beard, Bill Bower, Frank Boyd, Raulie Casteel, Jolene Friedland, Jeanette Makries, and Edwin Toledo. I appreciate their friendship and their hard work. Thanks to Dr. Daniel Rasse, Dr. Joseph Masabni, Mohamed Elwade, James Sallee, and Djail Santos, Laurent Gilet, Carlos Paglis and Dr. Helvecio De-Polli, and Dr. Tom Willson, and Jeff Smeenk for their friendship. vii TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. ix LIST OF FIGURES ............................................................................................................ x INTRODUCTION .............................................................................................................. 1 Prediction Errors and Efficiencies .......................................................................... 2 Spatial Interpolation ................................................................................................ 3 CHAPTER 1 ASSESSING MAP ACCURACY FOR SITE-SPECIFIC FERTILITY MANAGEMENT .................................................................. 7 INTRODUCTION .............................................................................................................. 7 MATERIALS AND METHODS ...................................................................................... 10 Site description ...................................................................................................... 10 Soil Sampling Design and Laboratory Analysis ................................................... 12 Data Analysis ........................................................................................................ 14 Phase 1: FULL data set .............................................................................. 14 Phase 11: Accuracy of SSF M sampling and interpolation ......................... 15 RESULTS AND DISCUSSION: ...................................................................................... 16 Phase 1. FULL Data Set. ....................................................................................... 16 Phase 11: Accuracy of SSFM sampling and interpolation ..................................... 26 Grid Sampling ........................................................................................... 26 CELL and directed sampling .................................................................... 36 CONCLUSIONS ............................................................................................................... 38 CHAPTER 2 COMPARISON OF TECHNIQUES TO OPTIMIZE SPATIAL ESTIMATES OF SOIL PROPERTIES USING TERRAIN ATTRIBUTES ........................................................................ 40 INTRODUCTION ............................................................................................................ 40 MATERIALS AND METHODS ...................................................................................... 45 Site description ...................................................................................................... 45 Soil Sampling Design and Laboratory Analysis ................................................... 47 DEM Creation and Terrain Analysis .................................................................... 48 Data Analysis ........................................................................................................ 48 RESULTS AND DISCUSSION ....................................................................................... 50 Nature of the Data ................................................................................................. 50 Evaluation of Interpolation Methods .................................................................... 58 CONCLUSIONS ............................................................................................................... 65 SUMMARY AND RECOMENDATIONS ...................................................................... 65 BIBLIOGRAPHY ............................................................................................................. 68 viii LIST OF TABLES Table 1.1 Map symbols, soil series or complex name, NRCS soil taxonomic Table 1.2. Table 1.3. Table 1.4. Table 2.1. Table 2.2. Table 2.3. Table 2.4. Table 2.5. description (Pregitzer, 197 8), and area occupied in the field ........... 10 Data set descriptions .......................................................................... 13 Summary statistics for the FULL data set. ........................................ 16 Directional and omnidirectional semivariogram model parameters for the FULL data set ....................................................................... 22 Applications of regression, kriging with an external drift, and cokriging for enhancing spatial estimates of soil properties using auxiliary information. ....................................................................... 41 Directional (G30) and omnidirectional (G30 and G61) semivariogram model parameters for carbon to be used for ordinary kriging ................................................................................ 55 Correlations between total carbon and terrain attributes and explained variability (G30 data set). .................................................. 56 Linear model of coregionalization (G30) used for cokriging. . ........... 57 Semivariograrn model parameters fo the direction of maximum spatial continuity (N 90 E) for kriging with a trend model and kriging with an external drift. ........................................................... 57 ix LIST OF FIGURES Figure 1.1. Modeled empirical semivariogram. .................................................... 6 Figure 1.2. A digital orthophotograph overlain by the 30 m (black square), the 100 111 (blue cross), the random validation data set (green circle) and NRCS soil types (CaA = Capac loam with 0 to 4 % slope; MeA = Metamora-Capac sandy loam with 0 to 4% slope; MoB = Morley Loam with 2 to 6% slope; and WbA = Wasepi sandy loam with 2 to 6 % slope)... .................................................. 11 Figure 1.3. Normal probability (Q-Q) plots for variables and log transformations with 95% confidence intervals. .............................. 18 Figure 1.4. Directional variogram contour maps for the FULL data set. ........... 20 Figure 1.5. Directional and omnidirectional semivariograms for the original and transformed FULL data sets. ..................................................... 21 Figure 1.6. Contour maps of kriged soil variables and fertilizer recommendations using the FULL data sets overlain by the soil type boundaries ................................................................................ 24 Figure 1.7. Omnidirectional experimental and modeled (exponential) semivariogram for the FULL, Gwmb, G30, G61, and G100 data sets ...27 Figure 1.8. The RMSE for several estimation approaches: IDW interpolation for distance exponents ranging from 0.1 to 5.0 in 0.1 increments (line plots on the left), kriging, whole field, and grid cell sampling; and zone sampling. .......................................................... 28 Figure 1.9. The relationship between the optimal distance exponent and the CV. ................................................................................................... 30 Figure 1.10. Predicted vs. measured for kriging with the Gcomb, G30, and G100 data sets: ........................................................................................... 31 Figure 1.11. The relationship between prediction efficiency for kriging the G30 data set, the range of spatial correlation of the FULL data set, and the RSV of the FULL data set. ............................................ 33 Figure 1.12. The relationship between prediction efficiency for the kriged G30 data set and the range of spatial correlation for the FULL data set for variables with RSV values greater than 70% .............................. 34 Figure 1.13. Prediction efficiency for kriging versus the prediction efficiency for IDW.’ ........................................................................................... 35 Figure 1.14. Average semivariograms (large solid circles) and semivariograms for individual pairs (hallow circles) for P using the Gm“, data set. ............................................................................. 37 Figure 2.1. Study location: (a) areal photograph and (b) soil and kinematic GPS measurement locations overlain by soil type boundaries ......... 46 Figure 2.2. Surface maps (a) of elevation and (b) slope and a contour map (c) of elevation overlain with arrows indicating aspect .................... 51 Figure 2.3. Normal probability (Q-Q) plots for variables and log transformation with 95% confidence intervals. ................................ 52 Figure 2.4. Sernivariogram surfaces for total carbon (G30) and elevation (n = 1000). ................................................................................................ 53 Figure 2.5. Total carbon (g kg") contour map created with ordinary kriging using the G30 data set overlain with elevation contours (lines). ....... 55 Figure 2.6. Prediction efficiencies and RMSE's. ................................................ 59 Figure 2.7. Predicted versus measured values for several prediction methods at two scales of measurement. ......................................................... 60 Figure 2.8. The relationship between prediction efficiency and sample point density for multiple regression interpolation .................................... 62 Figure 2.9. Predicted total carbon (g kg!) using kriging with an external drift (G30) and multiple regression (G30 and G61) ..................................... 64 xi INTRODUCTION Precision agriculture is about managing soil and crop variability in space and time in order to improve crop performance and environmental quality (Pierce and Nowak, 1999). Soil chemical and physical properties and crop yield are spatially quite variable. While some attributes are stable over time (e.g. total carbon, texture), others exhibit a great deal of temporal variability (e. g. soil N, soil water content water, grain yield, and pest infestation). The area of precision agriculture related to nutrient management is often referred to as site-specific fertility management (SSFM). In this dissertation, I am specifically concerned with the SSF M of soil nutrients that that are spatially variable but relatively temporally stable, specifically, total carbon, pH, P, K, Ca, and Mg. Several conditions are essential for successful SSF M, the most basic of which is that variation in soil properties is adequately known (Pierce and Nowak, 1999 and Sawyer, 1994). Some of the studies that have tested the validity of this premise have shown that it is not always valid (Wollenhaupt et a1, 1994; Gotway et al., 1996). Some SSFM agronomic and economic studies that have had mixed or negative results (Snyder et al., 1997; Wibawa et al., 1993; Wollenhaupt and Bucholz, 1993) might be explained by poor prediction accuracy of soil properties. Fortunately, there are good methods for measuring prediction and map accuracy. The kriging variance can not be used to estimate the map accuracy because it is independent of the data values (Deutsch and Journel, 1998) and only dependent on the covariance model and the data configuration (Goovaerts, 1997). Cross-validation is a technique where sample data points are sequentially dropped from the prediction data set, estimated from the neighboring point, and then replaced (Deutsch and Joumel, 1998). The measured values are subtracted from the predicted values to calculate the residuals which can be used to assess the accuracy of the predictions. For a regular gridded data set, cross-validation tends to over estimate prediction errors. A better approach would be to jack-knifing with an independent validation data set. This approach was used in this dissertation. The calculation of these errors is discussed in the first section of this introduction. The first objective of this dissertation was to evaluate how grid sampling and interpolation schemes affected map accuracy and prediction efficiency. The next objective was to evaluate different analytical techniques that using terrain attributes affect the spatial prediction of soil properties and whether relative performance of these techniques was affected by the scale of soil sampling. This dissertation requires some understanding of geostatistics, ordinary kriging, and inverse distance interpolation. The theory is presented in the second section of this introduction. Prediction Errors and Efficiencies In this study, measures of map error included MSE, RMSE (root mean squared error), and bias. Let vi denote the difference between predicted value and observed value at location 8;, i=1, ..., nv, where nV is the number of values in the validation data set. A map correct on average should have EN] = 0. The bias of the map is estimated as 11 V l "" Bias = ——Zvi i-l and the MSE of the map as MSE = Bias2 +—l—1—Zv(v,.—v)2 1 "V ——zvs 11v —1i=1 The RMSE is the square root of the MSE, RMSE-:JMS . The mean square error combines accuracy (biasz) with precision, the variance of the residuals. Prediction efficiency referred to as goodness by Gotway et a1. (1996) is calculated as Prediction efficiency = 100% * (MSE .30 — MSEMM XMSE .30 )“. [1] Positive prediction efficiencies can be interpreted as the percent reduction in MSE as compared with the field average approach (A30). A prediction efficiency of 15% for kriging can be interpreted as "compared to the field average approach, kriging reduced the MSE by 15%." The optimal distance exponent for IDW interpolation was determined as the one yielding the lowest RMSE. Spatial Interpolation A spatial estimator at an unobserved location 3 of an attribute Z is defined as a weighted average of the observed values of the attribute at spatial locations 5:. The weights Wsm may be restricted to non-zero values in some neighborhood N“) of the target location 5. If n(s) is the number of observed values in the neighborhood, the predicted value is calculated as ,. n(s) 2(s) = Zwsizc.) i=1 The essential difference between kriging and IDW estimation lies in the determination of the weights. The weights w for IDW are based on the distance between the point to be estimated and each of the n sample data points d(sa, 3) within the search neighborhood N(s). n(s) w. = [dew]: Eiders)? where e is the user defined distance exponent (Gotway, 1996). Smoothness of interpolated values decreases with the magnitude of the exponent. The kriging weights are calculated by solving the kriging system presented here in an expanded matrix form _ _ _ 1-1 . W1 CovH . - - Covln 1 Cov” 1 w" Covn1 . Covm 1 COVm M- _ 1 1 0_ . 1 . where )i. is a Lagrange parameter needed to satisfy certain constraints. Cov“ through Covm are the covariances among the sample data points, and C1, to Cns are the covariances between each sample data point and the unobserved location 5 (Goovaerts, 1997). For details of constrained minimization through Lagrange multipliers in this context see Isaaks and Srivastava (1989). Under weak stationarity conditions (Cressie, 1993) the covariances between two data points depends not on their actual coordinate, but matters. Since the kriging weights are functions of the covariances which in turn depend on spatial separation of data points, kriging weights are distance related weights too. The metric in which distances are assessed is not, however, Euclidean distance alone, but depends on the degree of spatial dependencies. Under second order stationarity, the covariance model are related to semivariogram models through the relationship COVij = C(O) - y(hi,~); where C(O) is the variance and hij is the Euclidean distance between locations 3; and Si the distance between sample locations. The semivariogram models are fit to empirical semivariograms 11(h) which is the average sample variance of points separated by distance h but computationally defined as V2 of the average squared difference of points separated by distance h, n(h) )=n[2 1(h)]Z[S (ti)— S(ai+h)] 2 i=1 where n(h) is the number of pairs at lag distance h or in some neighborhood of b. Figure 1.1 illustrates an isotropic semivariogram modeled with an exponential function. The plateau the variogram reaches is called the sill. The discontinuity at the origin is referred to as the nugget variance and is attributable to the additive effects of a white noise process and measurement error (Cressie, 1993). The sill less the nugget variance is referred to as the structural variance. The separation distance (lag) at which the variogram reaches the plateau (spherical models) or 95% of the sill (exponential and Gaussian models) is called the range or range of spatial correlation. 12 Variance at 95 % of sill \ _ _ _5. ° 0 ——f———e—~————e .H__u sm————-> Semivariance O) . / Range 1 [if _ _Nugget variance 0 I. 0 75 150 lag (m) Figure 1.1. Modeled empirical semivariogram. l l .(_ _ _Structural variance l I l I I l Relative structural variability (RSV) was used as a normalized measure of spatial dependence (Robertson et al., 1993; 1997) and is defined as RSV = structurallyanance = 1 _ RNE [2] $1 and is related to what is commonly referred to as the relative nugget effect (RN E). CHAPTER 1 Assessing Map Accuracy for Site-Specific Fertility Management INTRODUCTION Several conditions are essential for successful site-specific fertility management (SSFM). The most basic condition is that variation in soil properties is adequately known (Pierce and Nowak, 1999 and Sawyer, 1994). Soil property predictions across landscapes are affected by soil sampling, laboratory analysis, prediction, and cartographic errors. Poor map quality may explain why results of some SSF M agronomic and economic studies have had mixed or negative results (Snyder et al., 1997; Wibawa et al., 1993; Wollenhaupt and Bucholz, 1993). Some have suggested alternatives to grid sampling, including directed sampling (Pocknee et al., 1996). In general, condition and management maps are rarely examined for quality, which is unfortunate because methods exist to assess map accuracy. Measures of accuracy and goodness have been used in SSFM research to assess quality of maps and soil properties predictions, mostly by studying the impact of grid. sampling intensity (Wollenhaupt et al., 1994; Franzen and Peck, 1995; Gotway et al., 1996, Mohamed et al., 1996) and interpolation techniques (Wollenhaupt et al., 1994; Gotway et al., 1996, Mohamed et al., 1996) on map or prediction error. Wollenhaupt et a1. (1994) and Mohamed et al. (1996) considered map accuracy to be the percentage of areal overlap of mapping categories between maps in question and maps considered representative of the true spatial distribution of a soil property in space. These truth maps were arbitrarily defined to be contour maps created either with Delaunay triangulation of soil properties sampled at a 32-m grid (Wollenhaupt et al., 1994) or with the kriging of soil properties sampled on a 20 x 40-m grid (Mohamed et al., 1996). The correlation between predicted and observed data sets has been used as a measure of prediction accuracy (Franzen and Peck, 1995; Goovaerts, 1997). A correlation approach alone is problematic. While predicted and observed data sets may be highly correlated, they may deviate greatly from a 1:1 relationship and therefore be of low predictive value in a mapping context. Franzen and Peck (1995) assessed map error by assigning both measured values and their associated predictions to classes and then determining the percentage of measured values that were assigned to the same class as their predictions. Gotway et al. (1996) used mean square error (MSE) as a measure of prediction accuracy calculated using independent validation data sets. Map goodness was calculated by comparing the MSE for interpolation to the MSE for the field average approach. To a large extent, the appropriate grid spacing depends on the spatial structure of soil properties (Sadler et al., 1998) and the range of spatial correlation (Mohamed et al., 1996). Furthermore, the appropriate interpolation method may depend on specific coefficients of the interpolation procedure used. For example, the optimal distance exponent for the inverse distance weighted (IDW) interpolation procedure may depend upon the coefficient of variation (CV; Gotway et al., 1996). While grid sampling and interpolation approaches have limitations, a shifi from grid sampling to directed sampling schemes or grid cell based approaches may not improve map accuracy. Directed sampling is a technique wherein samples are collected and composited from specified areas within a field. Pocknee et al. (1996) suggest that the areas be delineated based on established differences within a field. Differentiating criteria might include soil map delineations (Bell et al., 1995, Moore et al., 1993, Windawa et al. 1993), management history, yield potential maps, aerial imagery (McCann et al., 1996; Pocknee et al., 1996), or electromagnetic induction (Jaynes, 1996). Directed sampling techniques require prior knowledge about the factors that regulate crop yield and nutrient availability. In general, map or prediction errors associated with directed sampling schemes have not been reported. Instead of sampling on grid points, composite samples can be taken from the area between grid points, a practice that is referred to as grid cell sampling. Grid cell sampling schemes have not found extensive use in SSFM. This may be related to earlier reports that grid point sampling more accurately described soil properties than did grid cell sampling, such as that reported by Wollenhaupt et al. (1994) for two fields in central Wisconsin. The purpose of this study was to evaluate how grid sampling and interpolation schemes affect map accuracy based on measures of bias, precision, and prediction efficiency, or what Gotway et al. (1996) termed goodness. For this study, a field was subjected to several soil sampling strategies including grid point sampling at several grid spacings, grid cell sampling, and directed sampling based on soil type. Soil fertility and fertilizer recommendation data were interpolated using kriging and a range of IDW coefficients for the various sampling schemes. Each resulting prediction map was tested against a random validation set to evaluate map accuracy. MATERIALS AND METHODS Site description This study was conducted within a 20.4-ha field (42° 57' 54" N, 84° 43' 38 W) in Clinton County, Michigan, 6-km south of Fowler. The field has been in a corn (Zea mays L.)-soybean (Glycine max L. (Mcrr.)) rotation for 22 years and the southeastern portion of the field has been sub-irrigated since 1988. The field was selected because it contained multiple soil map units and exhibits a range of terrain features, both of which would be conducive to SSFM. Pregitzer (1978) described and mapped the soils in Clinton County, MI at a scale of 1 :15,840. The great group taxonomic classifications of the soils in this field were either Ochraqualfs (Capac and Metamora) or Hapludalfs (Morley and Wasepi; Table 1.1). The soils were somewhat poorly drained except for the Morley, which was moderately well drained. All of the soils formed in glacial till except for the Wasepi, which formed in loamy glaciofluvial deposits. Table 1.1 Map symbols, soil series or complex name, NRCS soil taxonomic description (Pregitzer, 1978), and area occupied in the field Symbol Name and slope Taxonomic Family and Subgroup Area (ha) MoB Morley loam Fine, illitic, mesic 1.3 ‘ (2 to 6 % slope) Typic Hapludalfs CaA Capac loam Fine-loamy, mixed, mesic 7.2 (0 to 4 % slope) Aeric Ochraqualfs MeA Metamora-Capac sandy F ine-loamy, mixed, mesic 8.9 loams Udollic and Aerie Ochraqualfs (0 to 4 % slope) WbA Wasepi sandy loam Coarse-loamy, mixed, nonacidic, 3.0 (0 to 3 % slope) mesic Aquollic Hapludalfs 10 70 600 500 Northing (m) A O O 300 If Easting (m) Figure 1.2. A digital orthophotograph overlain by the 30 m (black square), the 100 m (blue cross), the random validation data set (green circle) and NRCS soil types (CaA = Capac loam with 0 to 4 % slope; MeA = Metamora-Capac sandy loam with 0 to 4% slope; MoB = Morley Loam with 2 to 6% slope; and WbA = Wasepi sandy loam with 2 to 6 % slope).i T red solid and dashed lines indicate boundaries of directed sampling zones (DS 1- 11) and dashed yellow lines indicate the boundaries of the CELL based sampling. 11 Soil Sampling Design and Laboratory Analysis Soil samples were obtained from the field using four sampling designs: a 30.5-m regular grid, a 100-m grid cell, a 200-m unaligned grid, and random sample design. These samples were used as data sets or to calculate data sets as described in Table 1.2. The sample point locations were flagged using a DGPS system with a base station for on- the-go differential correction. At each grid and randomly distributed point, 5 sub- samples (1 at the grid point and 4 within a 1.5-m radius) were obtained to a depth of 20- cm using a 2.5-cm diameter core and composited. The grid cell samples were taken by compositing 9 individual cores of the same diameter and depth taken at regular intervals within each 100-m grid cell (Figure 1.2). Soils were dried under forced air at 35° C for 3 days and ground to pass a 2-mm sieve. Standard soil analyses were conducted by the Michigan State University Soil and Plant Nutrient Laboratory using the recommended chemical soil test procedures for the North Central Region (Brown, 1998). Analyses included pH (1 :1 soil water mixture), BpH (SMP buffer), P (Bray P-l extractable) K, Ca, and Mg (lM/L NILOAc extractable). Cation exchange capacity (CEC) was calculated by summation and lime (an), P (Pm), and K (Km) fertilizer recommendations were calculated using the tri-state fertilizer recommendations (V itosh et al., 1995) for corn with a uniform yield goal of 11.3 Mg ha'l (180 bu acre"). 12 Table 1.2. Data set descriptions Name N Description Prediction Data Sets G30 215 Soil samples taken on a 30-m regular grid. G100 24 Soil samples taken on a 100-m regular grid. CELL 15 Soil samples taken in IOO-m grid cells Gcomb 239 Created by combining G30 and G100 grid data sets. Gel-a 54 Created by separating the 30-m grid data set into four separate Gal-b 54 6l-m grids. Each grid was analyzed separately, but the results are G6i-c 54 given as the average results from the four grids. Gm-d 53 DIR 11 Created by overlaying the soil map units onto the 30-m grid and determining the mean fertility value for each management unit. A30 1 Field average value for the entire field calculated from the means of the 30-m grid. A100 1 Field average value for the entire field calculated from the means of the IOO-m grid. Validation Data Set VAL 62 Independent validation created by combining a 200-m unaligned grid with additional random samples. Data set for geostatistical analyses FULL 301 Created by combining G30, G100, and VAL. 13 Data Analysis Data analysis was performed in two phases. Phase I involved a quantitative analysis of the FULL data set, which consisted of the combination of the G30, the G100, and the random (VAL) data sets (Table 1.2). The purpose of Phase I was to assess the extent to which the spatial variability of soil fertility in this field lends itself to SSF M. The steps in phase I included an assessment of normality, calculation of descriptive statistics, and analysis of spatial variability. Phase II was an assessment of the accuracy of sampling designs and interpolation procedures using quantitative measures of map quality. Phase I: FULL data set The FULL data set consisted of the 301 sample locations corresponding to an average sampling intensity of 14.8 samples ha'l (Table 1.2). Normal probability (Q-Q) with 95 % confidence intervals (Friendly, 1991) were used to assess normality of the FULL data set. When the Q-Q normal probability plots indicated large deviations from normality, the natural log of these variables was calculated and the resultant tested for normality. If the log transformations were also non-normal, then the original variables were power transformed and again tested for normality. The power transformations were not successful in inducing normality and will not be discussed further. Finally, normal score transformations (Deutsch and Joumel, 1998; Goovaerts, 1997) were used to normalize the remaining variables that could not be normalized with log or power transformations. Contour maps of variogram surfaces were created for each original variable to determine the direction of the anisotropic axes if anisotropy existed 14 (Goovaerts, 1997; Isaaks and Srivastava, 1989). For directional (anisotropic) semivariograms, an angular tolerance of i 40° was used because it allowed the variograrns to be well defined while still preserving their essential features (Isaaks and Srivastava, 1989). Nested semivariogram models (combinations of spherical, exponential, and/or Gaussian models) were chosen based on their fit to the empirical variograms, if warranted. Using the modeled semivariograms, GSLIB (Deutsch and Joumel, 1998) was used to create kriged 4 by 4-m grids at search radii equal to the distances to which the semivariograms were modeled. Contour maps for each kriged grid were created with Surfer® (Golden Software, Golden, CO). Phase 11: Accuracy of SSFM sampling and interpolation Phase II was concerned with how different sampling schemes and estimation procedures affected map accuracy. For the FULL, G30, 661.4, (3100, and Goon“, grid data sets, empirical semivariograms were calculated using Variowin (Pannatier, 1996) and an omnidirectional exponential model was fit to the empirical semivariograms. Surfer was used to interpolate 4 x 4 m grids by kriging each data set with the modeled semivariogram and using IDW for distance exponents from 0.1 to 5.0 in 0.1 increments. For comparison purposes, omnidirectional semivariograms based on an exponential model were also developed for the FULL data set. To quantify the error of prediction for each attribute, sampling scheme, and interpolation method, the difference between the predicted surface and the validation points were estimated. Since the VAL points did not always coincide with the 4 x 4 m predicted grid locations, bilinear interpolation in Surfer was used to estimate the predicted value at each VAL grid location. Because only one value is assigned to each 15 soil management unit, each cell, and to the entire field, residuals for the DIR, CELL, A30, and A100 predictions were calculated as the distance between the VAL points and the measured value for the area containing that point. Prediction errors and efficiencies were calculated as described in Chapter 1. RESULTS AND DISCUSSION: Phase 1. FULL Data Set. While the average soil tests for this field indicated that soil fertility was adequate, there was considerable range in each parameter (Table 1.3), indicating that some portion of the field may respond to SSFM. For SSFM to be applicable to this field, the variation of soil fertility must be spatially structured, of sufficient magnitude, and within the manageable range. Semivariance analysis was conducted to quantify the extent to which these conditions were met. Table 1.3. Summary statistics for the FULL data set. Variate mean Median MinT MaxT CV (%) pH 6.1 6.0 5.2 8.0 9 P (mg kg") 26 25 8 86 44 K (mg kg“) 167 165 95 291 23 Ca (mg kg") 1482 1450 650 3053 22 Mg (mg kg") 274 267 80 474 23 CEC (cmole kg“) 13.0 13.0 7.7 20.5 17 L...c (Mg ha") 4 5 0 13 76 Free (kg ha") 63 75 0 114 42 Kmc (kg ha") 25 o 0 100 136 1 Min = minimum, Max = maximum. 16 While normality is not a requirement for developing a semivariogram or kriging, the classical linear kriging predictor does not retain optimal properties if the underlying spatial process is not Gaussian (Cressie, 1993). Only Mg and CEC were normally distributed, while P, K and Ca were log-normally distributed (Figure 1.3, Table 1.3). Soil pH, Lm, Pm, and Krec could not be transformed to normality with log or power transformations but could be through a normal score transformations (Figure 1.3). The reason Lm, Pm, and Km deviated so drastically from normality is because the tri-state lime and fertilizer recommendations (V itosh et al., 1995) combine stair stepping, nested functions. Because of the nature of these calculations, however, the back transformations for the normal score transformations of LM, Pm, and Km were problematic and therefore should not be used for SSFM of this field. It may be that native Mg and CEC levels were distributed normally and remained so because lime applications have been primarily calcitic and CEC has minimally been affected by management practices. In another geo- spatial study in Michigan, soil properties in an uncultivated landscape appeared to be distributed more normally than those in an adjacent cultivated field (Robertson et al., 1993). Anisotropy occurs when the semivariogram depends not only on separation distance, but also on the angular relationships between data points. The presence and direction of anisotropy must be known to create anisotropic models and can easily be detected and measured using contour maps of semivariogram surfaces (Goovaerts, 1997). Only CBC and Lrec were considered to be isotropic (Figure 1.4). Directional or 17 I 7 I 'J//4 In a 6 . 2 0 q 5 .1 -3 .4) ‘J . at; 80 _‘ fl. 4 ...,: 4 d /- g 40 - “/4 a‘ 2 3 4 0 - g n T: 300 - /J a: 6 “ ,x x E 150 I Q ‘2 /. , 3;“ a V 0 ‘ i? - “A 1 5':- 4 g / 3 '3. 3000 - “.3; 7:. /-/e g 1500 4/ E g 8.0 1 V v a ‘ o .1 1 E E. L I 1 Je- 500 4 3:6 2 g 250 d/ V o q u o; 21 - _4 8 0 14 4 i 7 .1 V 1 v-A 15 .1 2?: 3 -4 I“ i § ‘2 0 d/ :8; 0 / " 2" “2 d v -15 _ -3 _’ .- e" 80 - _r' g 3 r g8 '2 40 4/ If. 0 ‘ / (L g ..- g, V 0 4,...__.' -3 g, w: 200 .. 47/18 3 -1 . ~ 1 g f, 0 F/ x 0 '1 / x 2 ‘“ -200 . -3 .. " -3-2-1012 3 -3-2-1012 3 Normal Quantiles Normal Quantiles Figure 1.3. Normal probability (Q-Q) plots for variables and log transformations with 95% confidence intervals. 1 ns indicates a normal score transformation; pH already log transformed 18 omnidirectional semivariograms were calculated and modeled for each variable and their normal transformations (Figure 1.5). The semivariogram models accounted for geometric (range changes with direction, sill does not) and zonal (sill changes with direction; range does not) anisotropy and mixtures of both (Isaaks, and Srivastava, 1989). The directional and non directional variograms were described with one, two, or three nested spherical, exponential, or Gaussian transitional structures in addition to a nugget structure (Figure 1.5 and Table 1.4). Overall, anisotropy was not severe and, removable with transformations (pH and Ca). It would not be cost effective for anisotropy to be modeled for this field on a commercial basis because the anisotropy is not strong, its modeling time and resource expensive, and the coarseness of commercially accepted SSFM would not allow the short range anisotropy to be resolved. The kriging system of equations requires that a second order stationary model be fit to the empirical semivariograms. Second order stationarity can be inferred the semivariogram reaches a plateau as occurred for most variables (Figure 1.5). When semivariograms do not reach a plateau, intrinsic stationarity is ofien assumed (Pannatier, 1996). But, by limiting the search radius, a stationary model can still be applied. The semivariograms for pH in the N 54° E direction and and Ca in the N 49° E direction exhibited intrinsic stationarity in these directions. This behavior was attributable to a "hot spot" in the southwestern comer of the field where lime had been stored (Figure 1.6) rather than a gradual trend in the data. Normalizing the pH and Ca data removed the directional, intrinsic behavior. Second order and intrinsic stationarity was assumed for the FULL data set and issues of non-stationarity are irrelevant for SSFM management of this field. 19 pH N$°\N North-south direction (m) -200' Mg cec N91°E -200 0 200 -200 0 200 East-west direction (111) Figure 1.4. Directional variogram contour maps for the FULL data set. I T white and black represent the minimum and maximum semivariogram, respectively of the FULL data set (not transformed). 20 "5(pH) . 0.7 - / Omnidirectional 0.0 - ‘ ln(P) 0.15 "‘ . N1go‘wo 0 00 y . N 71 ° E . ln(K) 0.04 -fiflz‘:: N 00 W i . N 90° E 0 . , 0.00 1 Ca ln(Ca) . o 0.- 80000 ~ ‘ N . E 0.03 g 8 N w Omnidirectional 5 0.00 . 'c (U .2 E <13 0 CEO 0 1 3.5 ‘f—‘fifiu—f‘ Omnidirectional 0.0 m Lroc ns(Lm) . V 7 - 0.7 4 ’ ' Omnidirectional / Omnidirectional 0 . 0.0 1 PM , . ns(Pm) mufflw. 1‘ . 0°w - N 71 ° E - N 90° E 0 L 0 1 Km 4 a Ans(KnJ n ‘ ‘ ‘ 8501M ‘ 1“: ::.o.o O M ‘ N0°W no ". N09‘w.. ' N 90° E . o o I 0 IN 90 E 0 175 350 0 175 350 L39 (m) Lag (m) Figure 1.5. Directional and omnidirectional semivariograms for the original and transformed FULL data sets. I 1' ns indicates a normal score transform. 21 How—o Tr 980302: Ea 03393283— moamfiiomama 580— $385688 men 9a mCrr 38 m9 man—8:6 _ mag—2:3 m 5533 2.me Zeno—4 a: 9832.4 25mm _wm / . ° / . QC) 10 _, / outlrer '§ 0 f) / Mg c / 3% g/ = -34.9 + 1.06 x ‘6 o . , - 0.00534 x2 93 ' r2=0.74 o- 1 1 l (3100 20 2 o o o 0 ‘ o l -20 ‘ . o 40 ~ 0 '60 f I T 40 80 120 Range of spatial correlation Figure 1.12. The relationship between prediction efficiency for the kriged G30 data set and the range of spatial correlation for the FULL data set for variables with RSV values greater than 70%. 34 30 ‘ /"i /’o’ / 0 - // to / to to I// yfi 30 .1 .E’ :z F” / i2 0 - ./ o r. / 0 £2 0 E / .‘2 / O E 30 ‘ e). 9’ ,/d' C / 23 O<- / o .2 /° 8 ' ° 5 L l / /o¢ 30 ~ .9” ,/,o / 0 — / O or / /’e' 30 ~ I,” / .15 0 ‘ oi/g‘ ° ,/ -15 - / 15 - j. 0 . °/ 0 / O ’/ -15 -i / / -15 0 15 6100 J 0” o 4 ///" l/ -30 ~ / .° . 3 2‘ 0 q / 0 4 O/ -30 - / . . O .2: 0 .i / 0 4 q 0/ -30 . / . . o ’9 o _ A. 4b or -30 - / 1 1 O ‘V o - 6. / / O / o e '30 q / I r . -3o 0 Prediction efficiency for IDW Optimum IDVV Exponent IDW Exponent = 1.0 IDW Exponent =1.5 IDW Exponent = 2.0 IDW Exponent cakuuahxi flonr regnmunons in Hgme111 Figure 1.13. Prediction efficiency for kriging versus the prediction efficiency for IDW.'* 1' Prediction efficiencies are for the G30 data sets. 35 negative for K and ln(K), indicating that the field average approach was superior. Normalizing the data had an insignificant or negative effect on the accuracy of the data. The poor performance of kriging may be explained by the large variability of the semivariograms for individual pairs (Figure 1.14). A fitted semivariogram model to the average semivariograms would be accurate for a small fraction of the pairs. So for a kriged estimate, two sample points the same distance from the prediction point will be weighted the same even though the variances between the prediction points and the point to be estimated (if it were known) would vary widely. It may be that the scatter of the semivariogram cloud may be one of the best indicators of the spatial predictability of a given variable. CELL and directed sampling Alternatives to grid sampling include cell sampling and directed sampling. The CELL sampling approach had the highest RMSE for most variables (Figure 1.8) and the lowest prediction efficiencies, confirming the results of Wollenhaupt et al. (1994). The directed sampling approach had RMSEs that were similar to the A30 field approach (prediction efficiencies near 0), indicating no advantage of directed sampling over the use of a mean value for this field (Figure 1.8). 36 Semivariogram 14000 12000 ~ 10000 a 8000 q 6000 - 4000 2 2000 e O o o o o o O o o oo o o o o o o o o o o o o o oo o o o o o o o oo 0 0 oo o o o o o °8 o o o o o o 8 o o 0 0° 0 8 g o o o o o o o o o a 03 <2. °°§ o o°8 °§° o 808 g °o 8 8 o a °8§ 8 o o 88 0° 0 3 8 8 0 0 g 5 38° 0 oo 0 3°88) fig 3 Separation distance Figure 1.14. Average semivariograms (large solid circles) and semivariograms for individual pairs (hallow circles) for P using the Goomb data set. 37 CONCLUSIONS Analysis of the FULL data set indicated that most soil fertility variables were spatially structured. Nevertheless, the presence of spatial structure alone did not prove sufficient for producing accurate yield maps, as evidenced by the plots of measured vs. predicted in Figures 1.10. Sampling at lower intensities increasingly diminished the delectability of spatial structure and generally increased the error of prediction as measured by RMSE. Where spatial structure was poor, particularly for K and Km, accurately sampling the field average was sufficient for nutrient management because SSFM for these variables was not appropriate. These data suggest that grid sampling at coarse grids and directed sampling were not adequate to produce accurate nutrient condition maps for this field. Cell sampling at least at the course 100-m grid intensity was also inadequate. These data suggest that grid point sampling at the industry standard 100-m intensity was inadequate. Sampling at greater intensities only modestly improved prediction accuracy, likely not enough to justify the geometric increase in sampling costs. In the second chapter of this dissertation, I will examine methodology for incorporating secondary landscape information into spatial estimates of a soil property at several scales. In this study, the accuracy of IDW interpolation with a distance exponent of 1.5 generally equaled or exceeded the accuracy of kriging at each scale of measurement. If the data had been strongly anisotropic or was not second order stationary, kriging may have been superior to IDW interpolation because the semivariogram model could account 38 for these peculiarities. The poor performance of kriging may be explained by the large variability of the semivariograms for individual pairs (Figure 1.14). Some measure of the scatter of the semivariogram cloud may be an indicator of the predictability in space. 39 CHAPTER 2 COMPARISON OF TECHNIQUES TO OPTIMIZE SPATIAL ESTIMATES OF SOIL PROPERTIES USING TERRAIN ATTRIBUTES INTRODUCTION Traditional survey methods and the more recent use of grid sampling and interpolation methods have not produced maps of soil properties with the accuracy needed for soil surveys (Bell et al., 1995; Moore et al., 1993) and precision farming (Pierce and Nowak, 1999; Robert, 1993). New analytical approaches are being used to utilize geometric properties of a landscape (slope, aspect, and curvature), collectively referred to as terrain attributes, to improve spatial estimates of soil properties. Terrain attributes are predictive of soil properties because topography is a soil forming factor. A high resolution, digital elevation model (DEM) is needed to calculate terrain attributes. Only recently has the technology become generally available to map elevation at the needed resolution, achieved through advances in high resolution global positioning system (GPS) now universally available. Several methods exist to use terrain attributes in spatial estimates of soil properties, ranging in complexity from simple regression to geostatistical methods. However, there is little consensus regarding which terrain attributes are most useful or which analytical method is most appropriate for a given soil property (Table 2.1). 40 Table 2.1. Applications of regression, kriging with an external drift, and cokriging for enhancing spatial estimates of soil properties using auxiliary information. Primary Secondary Measure of Reported results 1' accuracy and or goodness 1 Regression Moore et al., A horizon depth 4 terrain attributes Visual Regression 1993 (R2 = 0.51) comparison and approach was soil P (R2 = 0.48) the regression R2 considered good soil pH (R2 = 0.41) values because terrain particle size (R2 = 0.64) attributes could explain substantial 43 points hal 43 points ha'l variability Bell et al., A horizon depth 3 terrain attributes VDS; plots of A- horizon and 1995 (R2=0.51) predicted versus depths to carbonate depth to free carbonates measured predicted within 20 (18:04:) cm for 70% of validation samples 7.25 points ha’1 75 points he" Thomson and Total carbon 2 terrain attributes None Unclear Robert, 1995 (R2 = 0.66 to 0.69) and photographic tone 2.7 points hal 30 points ha’l Gessler et al., A horizon depth 2 terrain attributes Prediction error Regression reduced 1995 Solum depth (not specific) deviance by 63 and 68% scale not given scale not given Bourennane et al., 1996 Gotway and Harford, 1996 Goovaerts, 1997 Thickness of silty-clay- one terrain attribute loam pedological horizon 0.62 samples ha'l 4.8 samples ha'I residual soil NO;l corn grain yield 11 samples ha" 66 samples ha" Soil Cd, Cu, Pb, and Zn blocked Co estimates??? (259 samples per field (359 samples per area) ' field area) Kriging with an external drift (KED) VDS; ME and RMSE improvement over OK and KT Cross validation, COK increased the compared MSE MSE by 7% over for OK and KT OK with the MSE for KED VDS; rank Correlations correlations for between predicted predicted and and measured for measured and % COK explain 16 to misclassifrcation 35% more of the for both SK, OK variability than for and KED OK 41 Table 2.1, continued. Primary Secondary Measure of Reported Results accuracy and or goodness J‘vokriging (COK) Zhang et al., 1992 N03" and Ca electrical Cross validation, cokriging reduced (COK with pseudo- conductivity compared MSE the MSE by 78% crossvariogram) for OK with MSE 0.8 samples ha" 1.3 samples ha'l for COK Vaughan et al., Water content and surface electrical Visual inspection improvement over 1995 soil salinity conductivity ordinary kriging 0.12 points ha'l 0.15 samples ha'l Rosenbaum and Soil Cd Soil Zn Independent Correlation SOderstrOm, 1996 validation data set; between predicted (standardized 0.00034 points ha‘l 0.0010 samples ha“n correlation of and measured was ordinary cokriging) predicted and greater for COK measured (p = 0.85) than OK (p = 0.68) Gotway and residual soil N03 corn grain yield Cross validation, COK reduced the Harford, 1996 compared MSE MSE by 2% 10.8 samples ha'1 66 samples ha’l for ordinary kriging with MSE for cokriging Zhang et al., 1997 soil solute soil solute Cross validation, COK reduced the concentrations concentrations compared MSE MSE between 30 measured at for OK with MSE and 60 % shallower depth for COK 1.3 - 1.8 points ha'l 1.3 - 1.8 samples ha'l Goovaerts, 1997 Soil Cd, Cu, Pb, Four combinations of VDS; correlations Correlation (isotropic and and Co (259 Ni, Zn, Pb, and or between predicted greater and errors anisotropic samples per field Cu. and measured and lower for COK standardized area) ??? ME for both OK compared with ordinary cokriging) and COK OK Juang and Lee, Soil Cd and Zn Soil Cd and Zn at Compared OK and COK improved 1998 same or lower depth COK predictions the r2 values by 6 (scale = 2 points to 60% over 2 points ha" 5.5 points ha'I ha") with OK COK. predictions (scale = 5.5 points ha") 1' VDS = Validation data set; MSE = Mean squared error; RMSE = Root mean squared error; ME = mean absolute error; SK = Simple Kriging; OK = Ordinary kriging; KT = Kriging with a trend model; COK = cokriging. 42 The simplest approach has been the use of simple or multiple regression in which a soil property is regressed on a single or multiple terrain attributes and the regression equation used to predict the soil property at unsampled locations within the field where terrain attributes are mapped. Success has been measured by the magnitude of the regression coefficient of determination (R2), which ranged from 0.41 to 0.69 for the studies reported in Table 2.1. In some cases, the accuracy of regression prediction was evaluated using a validation data set (Bell et al., 1995), which is a more robust accuracy measure . The regression approach relies solely upon the relationship between the soil property of interest and the selected terrain attributes. Geostatistical prediction approaches utilize a statistical model of the spatial variability either using distance alone (ordinary kriging) or in conjunction with other measured variables (e.g., co-kriging). For ordinary kriging, a search radius is defined for each point that is to be assigned an estimate. A mean attribute value is calculated from sample data points within this radius and subtracted from each sample data point value. A weighted average of the residuals is calculated. The weights are based on an empirical, statistical model of the relationship between the separation distance and sample variance (semivariogram model). The neighborhood mean is added to the weighted average of the residuals to calculate an ordinary kriging point estimate. Estimation may be improved by incorporating secondary information into kriged estimates by substituting the mean term with a smoothly changing, rescaled variable (e.g. terrain attribute) that is linearly related to the variable being predicted, a procedure known as kriging with an external drift (Goovaerts, 1997; Deutch and Joumel, 1998). Multiple secondary variables can be incorporated in this 43 fashion with a procedure known as random field analysis. However, this procedure has traditionally been used to remove spatial correlation from an analysis of variance (Stroup et a1, 1994). Standardized ordinary co-kriging does not use the mean to incorporate secondary information. Rather, the prediction is the sum of the weighted averages of the primary variable and each of the secondary variables. The performance of the various kriging approaches is mixed (Table 2.1). Goovaerts (1997) reports that, while correlated, kriging with an external drift and cokriging performed better than simple or ordinary kriging. In addition, cokriging performed better than kriging with an external drifi for three of the four variables and anisotropic cokriging performed better than isotropic cokriging. However, Gotway and Hartford (1996) found that cokriging and kriging were respectively worse or only slightly better than ordinary kriging. The fact that correlations between residual nitrate and yield were not significant (p = -0.09) may explain the poor performance of these techniques. From the studies in Table 2.1, there appears to be no consistency in which soil properties were analyzed, which terrain attributes were selected for prediction, the resolution of sampling of either soil properties or elevation, the scale of analysis, or the measures of accuracy of prediction, if used at all. Furthermore, the analytical techniques used in the various studies varied in complexity and in the effort required for analysis (regression < kriging with external drifi << random field analysis << cokriging. Increased complexity is only warranted if it leads to significantly improved spatial prediction. The objective of this study was to evaluate how different analytical techniques using terrain attributes affect the spatial prediction of soil properties and whether relative performance of these techniques is affected by the scale of soil sampling. 44 Four analytical techniques were used to generate spatial predictions of soil carbon obtained on 30 and 100 m regular grids using elevation, slope, and curvature as predictors and regression, residuals, and prediction efficiency as measures of performance. MATERIALS AND METHODS Site description This study was conducted in a 12.5 ha field (47° 47' 30" N, 83° 52' 30 W) located 6-km south of Durand, Michigan in the Shiawassee River watershed (Figure 2.1). The field had been in a corn (Zea mays L.)-soybean (Glycine max L. (Merr.)) rotation for more than lO-yr. The soil color differences in the aerial photograph (Figure 2.1a) were related primarily to differences in soil organic matter content and drainage but did not match well with the second soil order survey map unit boundaries (Figure 2.1b). Because moisture conditions were not optimal in other years when USDA-AF S aerial photographs were taken (e.g. 1979, 1983, and 1992), the striking visual differences were not captured as they were in this photo (Figure 2.1a). The soil scientists who created this survey relied on aerial photography taken prior to 195 8, which may have been of a lower quality or, taken with a full crop canopy. With better aerial photos, there may have been a better match between the color differences in Figure 2.1a and the soil survey map unit boundaries. 45 Northing (m) 200 300 400 500 600 Easting (m) Figure 2.1. Study location: (a) areal photograph and (b) soil and hepatic GPS measurement locations overlain by soil type boundaries. 1‘ The scanned and enlarged aerial photograph (original scale = 1:7,920; not georectified) taken 6/23/88 was purchased from the USDA-AF S Aerial Photography Field Office in Salt Lake City, UT. The locations of sample points for the three soil sampling strategies (G30 = V; Gioo = E]; VAL = +) and the kinematic survey (0) are overlain by NRCS soil map unit boundaries (Bt = Breckenridge sandy loam; CtA = Conover loam with 0 to 2 % slope; MaA = Macomb loam with 0 to 2 % slope; MsA = Metamora sandy loam with 0 to 2 % slopes; MsB = Metarnora sandy loam with 2 to 6 % slopes; WeA = Wasepi sandy loam with 0 to 2 % slopes; Threlkeld and Feenstra, 1974). 46 Threlkeld and Feenstra (1974) classified and described the soils in Shiawassee County, MI at a scale of 1:20,000. The soils were mapped (Figure 2. lb) as somewhat poorly drained Alfisols with the exception of the Breckenridge (Bt; Coarse-loamy, mixed, nonacid, frigid Mollie Haplaquepts), a poorly drained Inceptisol. The Metamora sandy loam and Macomb loam (F inc-loamy, mixed, mesic Udollic Ochraqualfs) were very similar but the Metamora was coarser in texture which means that its surface drained somewhat faster but they both have slow subsurface drainage. While permeability is moderately rapid for the Wasepi (Coarse-loamy, mixed, mesic Aquollic Hapludalfs) series, it has low available water holding capacity. Most of the field had slopes of less than 2% except for the Metamora map unit with a B slope (MsB), ranging from 2 to 6%. Soil Sampling Design and Laboratory Analysis Soil samples were obtained from the field (Figure 2.1b) in May of 1997 using a 30.5-m (G30; 11 = 134; 10.7 samples ha‘l) regular grid, a 100-m(Gloo; n = 12; 1 sample ha' 1) regular grid, and a set of validation points (VAL; n=26; 2.1 samples ha"). The VAL points were collected using a 200 m unaligned grid with additional random points. The sample point locations were flagged using a DGPS system with a base station for on-the- go differential correction. At the each sample locations, 5 sub-samples (l at the grid point and 4 within a 1.5-m radius) were obtained to a depth of 20-cm using a 2.5-cm diameter soil core and composited. Soils were dried under forced air at 35° C for 3 days and pulverized to pass a 2-mm sieve. Sieved soil was finely ground with a roller mill and then analyzed for total carbon using a Carlo Erba NA 1500 Series 2 N/C/S analyzer (CE Instruments Milan, Italy). A 61-m grid (G51; 11 = 38; 2.7 samples ha") was extracted from the G30 data set to be used as a third prediction data set. 47 DEM Creation and Terrain Analysis A kinematic GPS survey was conducted in January of 1996 using two Z-12 Ashtech GPS sensors. The mobile GPS unit was mounted on an all terrain vehicle (ATV) traveling at about 17 km hr'l logging GPS location and elevation. Every second, a data point was logged so that the approximate distance between measurements was 4.7 meters. The ATV traversed the field in the east-west direction making swaths every 4.6 meters so the field was sampled at an approximate scale of 463 samples ha". Data that had high position dilution of precision values (PDOP) and large vertical jumps between sequentially logged data points were removed. Several swaths were removed from the northern region of the field because of a systematic error in the GPS data (Figure 2.1b). Topogrid (Arclnfo ver. 7.1.1, ESRI 1997) was used to create a l x l-m grid without drainage enforcement. Slope, aspect, and curvature (plan, profile, and tangential) were calculated with Surfer® (Golden Software, Golden, CO). Data Analysis For the G30 data set, normal probability (Q-Q) plots with 95 % confidence intervals (Friendly, 1991) were used to assess normality. Contour maps of semivariogram surfaces were created for total carbon and the terrain attributes to determine the direction of the anisotropic axes if anisotropy existed (Goovaerts, 1997; Isaaks and Srivastava, 1989). Directional (anisotropic) semivariograms were calculated for soil properties and terrain attributes using angular tolerances of i 22.5° for the soil variables (Goovaerts, 1997) and i 15° for terrain attributes (a smaller angle was used because terrain model was more densely sampled). For total carbon, nested semivariograms models 48 (combinations of spherical, exponential, and/or Gaussian models) if warranted, were chosen based on their fit to the empirical variograms. All variogram modeling was performed with Variowin (Pannatier, 1996). Correlations (or = 0.15) and multiple regression (or = 0.15) were calculated using SAS (SAS, 1990) for each grid sampling interval. The G30 and G61 data sets were interpolated with regression, ordinary kriging, kriging with a trend model, kriging with an external drift, random field analysis, and standardized ordinary cokriging. Because there were so few points in the G100 data set, only regression analysis was performed. All geostatistical interpolation methods were conducted with GSLib (Deutsch and Joumel, 1998) except for random field analysis which was performed using SAS (SAS, 1990). Some theoretical understanding is required to fully appreciate these procedures. A geostatistical prediction at an unobserved location 3 of an attribute Z is the weighted average of the observed values of the attribute at spatial locations 5;. The weights wSm may be restricted to non-zero values in some neighborhood N“) of the target location 3. If n(s) is the number of observed values in the neighborhood, the predicted value is calculated as n(s ZKoo—m(u) = warms-mm] Eqn. 1 i=1 Goovaerts (1997, 1999) distinguishes between kriging interpolators by the treatment of the mean term m(u) or m(ui). The mean is constant throughout the study area for simple kriging (SK), and within each search neighborhood, but varies through the study area for OK, and varies gradually within each neighborhood N“) for KT, KED, 49 and RFA. The mean component is modeled as a linear combination of the coordinates for KT, and as a linear function of a smoothly varying secondary variable (e. g. terrain attribute) for KED, and a linear or nonlinear combination of secondary variables for RFA. The kriging weights are calculated by solving the kriging system presented in Chapter 1 of this dissertation. Measures of map error included MSE, RMSE (root mean squared error), and bias. Let V, denote the differences between predicted value and observed value at location 5;, i=1, ..., nv = 62 of the validation data set. Map errors and prediction efficiencies were calculated as described in the Introduction to the dissertation. RESULTS AND DISCUSSION The discussion here focuses on two issues. The first is whether variability in soil and terrain attributes within the field have the magnitude and structure needed for spatial prediction. Then interpolation methods that utilize terrain attributes will be evaluate using measures of prediction error and efficiency. Nature of the Data Due to its glacial origin, the elevation and derived terrain attributes within this field varied considerably (Figure 2.2). While total relief in the field is only 4 m, there is considerable micro-variability within the field as evidenced by rapid changes in slope and aspect over short distances. Therefore, micro-variability in terrain attributes may exert significant influence in the soil and hydrologic properties of this field. Elevation, slope and aspect (Figure 2.3) were normally distributed while plan curvature, profile curvature, and tangential curvature were not. Kriging or cokriging with non-Gaussian data is 50 20 x vertical exaggeration Benefits“ \m\ N N S\Q§% \Sio\ __-; to izrvrtitrrvtssxxxzfi" <9 i 261 g i Wfi‘ifiifiji‘) «1" ,1"?! LI. a °°°i "tr-«J» W 1E ‘ 2 iii-3'3“, 3;? o . 4d: 2 i 3 7007 Easting (m) Figure 2.2. Surface maps (a) of elevation and (b) slope and a contour map (c) of elevation overlain with arrows indicating aspect. 51 TC (mg kg") Elevation Aspect Profile Plan Slope Curvature (angle) (degrees) Curvature Tangential (m) Curvature 404 20- 04 268 4 264 a 260 _‘ . o... 400 l 200 - 0-... 43 2-4 0+0. 0.3 ~ 0.0 .4 -0.3 - 0.003 . 0.000 - -0.003 - i" -. 0.005 - 0.000 - -0.005 4 ' -3 -2 2 3 Normal Quantiles -1 Figure 2.3. Normal probability (Q-Q) plots for variables and log transformation with 95% confidence intervals. 52 Separation distance (meters North Separation distance (meters East) Figure 2.4. Semivariogram surfaces for total carbon (G30) and elevation (n = 1000). permissible but the predictions are not guaranteed to be best linear unbiased estimates (Cressie, 1993). Elevation was severely anisotropic (Figure 2.4) with the axis of maximum spatial continuity 62° East from due North. Intrinsic stationarity was assumed in the orthogonal direction because the semivariogram did not reach a plateau. The anisotropic axes for slope were similar to the anisotropic axes for elevation except they were rotated 90° and were less severe (not shown). Plan curvature, profile curvature, and tangential curvature were mildly anisotropic (not shown). The terrain attributes had large RSV values. Elevation had a range of spatial correlation of 275 m. Slope and aspect had ranges of about 70 m and the curvature parameters had ranges between 10 and 20 m. Elevation, slope and aspect were suitable to be used in a geostatistical analysis. The use of curvature in the geostatistical study is questionable because they were spatially correlated over such a short range and because they were not normally distributed. 53 Total carbon was normally distributed (Figure 1.3). The average value for the field was 13 g kg'l, typical for a Michigan landscape. Despite just moderate changes in relief and slope (Figure 2.2), total carbon content ranged substantially (2 to 29 g kg'l). The anisotropic axes occurred in the North-South and East—West directions but at distances of 200-m and greater the axes appear to shift to the same anisotropic axes system as for elevation (Figure 2.4). The RSV values were not as high as might be expected for total carbon which tends to be well structured (Table 2.2) indicating a large nugget effect. The nugget can not be accurately estimated when distances between sample points are great (e.g. 30.5 and 61-m). Therefore, the RSV could not be interpreted as a measure of spatial dependence for total carbon. The range parameters for the anisotropic model also was not interpreted because a technique had been employed to account for zonal anisotropic accomplished by manipulating the range parameters. The range of spatial correlation was quite large for the two isotropic models (244 and 249 m) indicating that the data were spatially well structured. Based on the large range of spatial correlation, total carbon is suitable for geostatistical analyses. Ordinary kriging using the geostatistical parameters in Table 2.2 reveals that carbon values were generally lower in the on hill tops and in the northern region of the field and greater in the depressions in the southern and southeastern regions of the field (Figure 2.5). Unfortunately, however, much of the detail apparent in the aerial photo (Figure 2. l a) was not represented in this interpolation. In short, ordinary kriging did an adequate job of predicting soil carbon across the landscape but there is room for improvement. To apply other interpolation techniques, however, additional data requirements must be satisfied. 54 Table 2.2. Directional (G30) and omnidirectional (G30 and G51) SCmivariogram model parameters for carbon to be used for ordinary kriging —————-Structure 1—— ———Strueture 2 nugget model sill direction range RSV model sill direction range (m) (%) EL Isotropic (Gm) 6.2 S 31 O 249 75 - - N90° E 146 N80° E 20000 An'sg‘mp'c 10.2 G 15 60 G (30) N 0°w 118 N10°W 950 Isotropic (G61) 2.6 S 50 O 244 95 i” S = spherical; G = Gaussian; O = omnidirectional 600 Easting (m) 15 1 7 20 25 30 Figure 2.5. Total carbon (g kg") contour map created with ordinary kriging using the G30 data set overlain with elevation contours (lines). For kriging with an external drift, the relationship between primary and secondary variables must be linear (Goovaerts, 1997). For cokriging variables must be both correlated and have structured cross semivariograms. Ahmed and De Marsily (1987) state that for cokriging to be of greater predictive value than ordinary kriging, the absolute value of the correlation coefficients between predicted and measured must 55 exceed 0.70 (Table 2.3). While most of the variables were significantly correlated with total carbon, only elevation could explain a substantial portion of its variability. In this study, elevation was the only variable suitable for cokriging and kriging with an external drift. Table 2.3. Correlations between total carbon and terrain attributes and explained variability (G30 data set) I. Correlations Variability in total with carbon explained Total by terrain Carbon attributes (%) Elevation -0.72 * 51 Aspect 0.17 3 Slope -0.40 * 16 Plan curvature 0.19 * 4 Profile curvature 0.20 * 4 Tangential curvature 0.21 * 4 l‘n = 134; * indicates significance at or = 0.05 An important requirement for cokriging is that a linear model of coregionalization be developed that has covariance matrices that are positive semi-definite (Goovaerts, 1997). The parameters for the model at the G30 scale are listed in Table 2.4. Another important requirement for kriging with a trend and kriging with an external drift is that a stationary trend exists. The models for the directional trend are presented in Table 2.5. In summary, total carbon and elevation had sufficiently large correlations and structured semivariograms and cross semivariograms so ordinary kriging, and cokriging were appropriate methods. Because total carbon also had a directional trend, kriging with 56 Table 2.4. Linear model of coregionalization (G30) used for cokriging. . Structure 1 Structure 2 Structure 3 (Gaussian) (Gaussian) (Gaussian) svl or NugT Sill Direction Range Sill Direction Range Sill Direction range Cross-8V (m) (m) (m) f N 90° E 145 N 90° E 2000 N 62 E 6000 TC 9 15 50 6.3 N 0°W 117 N 0°W 440 N28W 420 TC x N 90° E 145 N 90° E 2000 N 62 E 6000 . -0.23 -1.71 -1.56 -3.61 E'evatlon N 0° w 117 N 0° w 440 N 28 w 420 N 90° E 145 0 09 N 90° E 2000 N 62 E 6000 Elevation 0.02 0.225 ' 2.1 N 0°w 117 3 N 0°w 440 N28W 420 1 SV = semivariogram; Nug = nugget; TC = total carbon Table 2.5. Semivariogram model parameters fo the direction of maximum spatial continuity (N 90 E) for kriging with a trend model and kriging with an external drift. scale nugget modelI sill direction range (m) G30 9.4 G 16 N 90° E 190 G61 4.4 S 50 N 90° E 213 1' S = spherical; G = Gaussian a (quadratic) trend and kriging with an external drifi (elevation) were also appropriate methods. As will be presented in the next section, there were significant regression relationships between total carbon and the terrain attributes. Because of this, multiple regression and random field analysis are appropriate prediction methods. 57 Evaluation of Interpolation Methods Stepwise regression (or = 0.15) was used to predict total carbon each scale. The regressor variables were Easting (m), Northing (m), elevation (m), slope (%), plan curvature (m'l), profile curvature (ml), and tangential curvature (m") as independent variables; however, at each scale only various combinations of these regressors were selected to be in the model by the stepwise procedure. Total Carbon (G30) = 1451 - 5.45 x elevation - 0.0145 x Northing - 2.95 x slope + 3.24 x plan curvature + 619 x profile curvature Total carbon (G61) = 1567 - 5.88 x elevation - 0.0203 x Northing - 4.66 x slope + 18.6 x plan curvature - 993 x profile curvature Total Carbon (G100) = 59.48 - 0.0589 x Northing More than half of the variability in total carbon was predicted at the G30 (R2 = 0.66), G61 (R2 = 0.77), and G100 (R2 = 0.74) scales. At the G100 scale, only Northing was retained in the model and the R2 for this relation was large. Visually, the gradient in TC is from N to S (Figure 2.1) and the few data points in the G100 grid could only identify this major trend. Therefore the regression from the G100 data set (n=12) represents spurious results because the measure of goodness were low. This is evident by a large RMSE, low prediction efficiency (Figure 2.6), and low 1'2 between predicted and measured (Figure 2.7) for regression procedure at the 0100 scale. 58 c5 60-+'fi'—'—__fl flm 9: ’7 13.9.23 0: - o ,. , , 4- — tub _— (D'c: j 2.: “EZWIV—ll—‘Iflflfl fl HHl—lfl o , X D< (9 x 0 <0 (9 I)! x 00 58 §& E 05 S'i‘fific‘ 35 Figure 2.6. Prediction efficiencies and RMSE's. 'l' FA30 = mean value of the G30 data set; OK; = isotropic ordinary kriging; OKa = anisotropic ordinary kriging; KT = kriging with a trend; COK = cokriging; KED = kriging with an external drift; RFA = random field analysis. 59 Total carbon predicted (g kg") Total carbon measured (9 kg") Figure 2.7. Predicted versus measured values for several prediction methods at two scales of measurement. 7 ‘l' OK= isotropic ordinary kriging; KT = kriging with a trend model; COK = cokriging; KED = kriging with an external drift; RF A = random field analysis; 60 At the G30 intensity, there was little difference between any of the prediction methods as assessed with measures of prediction efficiency and RMSE (Figure 2.6) or by deviations from a 1:1 line of predicted versus measured (Figure 2.7). At this scale, ordinary kriging and kriging with a trend model, methods which rely solely on a statistical model of the spatial variability, were of similar predictive value as methods that relied only on the relationship between total carbon and terrain attributes (multiple regression). Methods that utilized both the spatial variability of carbon and its relationship with other variables (e. g. kriging with an external drift) only performed slightly better than these methods. At the G61 grid intensity, the regression approach had substantially lower RMSEs and higher prediction efficiencies than any other method at this scale. In fact, the RMSE for multiple regression was nearly in the same range as the RMSEs for the G30 interpolations (Figure 2.6). The plots of predicted versus measured show that at the 61-m grid scale, all of the methods that incorporated terrain information out performed those that used only geostatistical information, despite the large range of spatial correlation. A great deal of information about total carbon exists in the terrain attributes. The implication is that by using this information, it is possible to reduce the number of samples needed to achieve a certain level of accuracy. The cost of soil sampling is inversely proportional to the square of the grid increment. For regression, prediction efficiency and grid sampling interval were inversely related (Figure 2.8) which allow the relationship between prediction efficiency and cost to be modeled. It should be noted that by considering the prediction efficiency, a squared loss function is assumed (since prediction efficiency is based on the MSE). 61 This is likely an incorrect assumption but unfortunately is the basis for most statistical measures of map error. In other words, an accurate understanding of the cost benefit relationship for grid sampling will require a better understanding of the relationship between map accuracy and the relative benefit to the farmer. Recall that in Table 2.1, researchers who used the regression approach used the coefficients of determination (R2) as an indicator of goodness. Figure 2.8 shows that the prediction efficiency decreases going from the G30 to G61 scales however the stepwise regression R2 values increased from 0.66 and 0.77 for the G30 and G61 grids. This suggests that the R2 may not be a very robust indicator of goodness. 70 65- 60- r2=1.00 55- 502 45.. 40- 35 1 —T 1 1 20 40 60 80 100 120 Grid sampling interval (meters) Prediction efficiency (%) Figure 2.8. The relationship between prediction efficiency and sample point density for multiple regression interpolation. Standard measures of prediction accuracy and efficiency are important. Some researchers use the correlations between predicted and measured while others compare MSE values (Table 2.1). Unforttmately, these measures are not the same and can give conflicting results. In Figure 2.7, the r2 for ordinary kriging at G61 is 0.47 and the r2 for 62 regression interpolation at G100 is 0.41 but the RMSE values are lower for ordinary kriging than for the regression interpolation (Figure 2.6). Clearly some standards are needed because these approaches can give conflicting results. Although the predictions at the G30 intensity were not substantially different, kriging with an external drift performed the best based on higher correlation coefficients (Figure 2.7), greater prediction efficiencies, and lower RMSEs (Figure 2.6). However, the best interpolator at the G61 scale was multiple regression. The kriging with an external drift and multiple regression (G30 and G61) interpolations have been overlain with elevation contours in Figure 2.9. Of these interpolations, by visual inspection, there is more correspondence between the kriging with an external drift interpolation and the aerial photograph (Figure 2.1a). In the southern and northeastern regions of the field, the kriging with an external drift interpolation matches the darker tones in the aerial photo better than the two multiple regression interpolations. But regression analysis does a better job of assigning high carbon values to low areas like the veiny feature in the west half of the field and low carbon values to ridges the bright red area in the north central region of the field. The regression approach is successful for these areas because the ridges and valleys have ether extremely positive or negative curvature values. 63 Northing (m) 600 Easting (m) o 54’10121517202530 Figure 2.9. Predicted total carbon (g kg") using kriging with an external drift (G30) and multiple regression (G30 and G61). 1' KED = Kriging with an external drift; REG = Multiple regression; For display purposes regressed interpolations were matrix smoothed (8 surrounding cells, central cell weight was 2.5; other 8 cells weights were I) CONCLUSIONS Terrain attributes improved spatial predictions of total carbon particularly with the coarser sampling intensities. The comparable performance of multiple regression interpolation procedure suggests that it may not be necessary to use the more sophisticated geostatistical techniques. At high sampling densities (G30), interpolation method had little impact on overall map accuracy. There was an interaction between the scale of measurement and the most appropriate interpolation procedure. At lower sampling intensity (G61), methods that utilized terrain attributes were more precise than methods that did not. At this scale, multiple regression analysis yielded the best predictions, which were nearly as accurate as the methods sampled at the G30 scale. By using techniques that incorporate terrain attributes, sampling intensity can be substantially reduced, while maintaining high levels of prediction accuracy and precision. Since the cost of grid sampling is inversely related to the square of the grid sampling increment, enhancing spatial estimates with terrain attributes is economically appealing. SUMMARY AND RECOMENDATIONS In the first chapter of this dissertation, grid sampling was found to be inadequate for accurate spatial predictions of soil chemical properties for a field in Clinton County, MI. The second chapter provides an example of how auxiliary terrain information (elevation, slope, aspect, and curvature) can successfully be used to enhance spatial estimates of total carbon. More work is needed to determine if terrain attributes can be used to enhance the predictions of soil fertility variables (e.g. pH, P, K). Additionally, other terrain attributes (e. g. wetness index, specific catchment area), high-resolution 65 multi-spectral images, electromagnetic conductivity, ground-penetrating radar could be used as secondary information. Soil property sensors are greatly needed for precision agriculture, unfortunately there are few commercially available sensors that directly measure soil properties of agronomic importance. Development of on-the-go sensors for soil nutrients has been limited primarily to nitrogen and has not been very successful. Sensors for soil organic matter, water content, structure, compaction are still in development but have had greater SUCCESS. Until better sensors are developed, a two step approach is recommended. The first step is directed sampling based on field history, landscape features, remote sensed imagery, and yield map variability. Directed sampling may not provide accurate soil property maps. This was the case in the first chapter of this dissertation; however, management zones were based on NRCS soil types from a 1215,840 survey. However, if a producer has records of within field manure applications or records of the locations of old field boundaries, this approach may prove to be quite accurate. The second step is the composite sampling of areas that have depressed yields or stressed plants as indicated in yield maps and remote sensed imagery. Plant nitrogen deficiencies can cause chlorosis, which changes the reflective properties of the plants. Chlorosis can be identified by aerial images. 66 BIBLIOGRAPHY BIBLIOGRAPHY Ahmed, S. and De Marsily, G. 1987. Comparison of geostatistical methods for estimating transmissivity using data on transmissivity and specific capacity. Water Resources Research, 23: 1717-1737. Bell, J .C ., C.A. Butler, and J .A. Thomson. 1995. Soil-terrain modeling for site-specific agriculture management. p 208-227. In P.C. Robert et al. (ed.) Site —specific management for agriculture systems. ASA Misc. Publ. ASA, CSSA, and SSSA, Madison, WI. Bourennane, H. D. King, P. Chery, and A. Bruand. 1996. Improving the kriging of soil variables using slope gradient as external drift. European journal of soil science. 47: 473-483. Brown, J .R. 1998.(ed.) Recommended Chemical Soil Test Procedures for the North Central Region. Missouri Ag Exp. Sta. Bul 1001. Cressie, N. 1993. Statistics for spatial data, revised edition. John Wiley & Sons, New York, 1991. Deutsch, CV. and AG. JourneI. 1998. GSLIB: geostatistical software library and users guide, second edition. Oxford University Press, New York. Franzen, D.W. and TR. Peck. 1995. Field soil sampling density for variable rate fertilization. J. Prod. Agric. 8(4): 568-574. Friendly, M. 1991, SAS system for statistical graphics, lst edition, SAS Institute, Inc.', Cary NC. Gessler, P.E., I.D. Moore, N.J. McKenzie, and P.J. Ryan. 1995. Soil-landscape modelling and spatial prediction of soil attributes. Int. J. Geographical information systems. 9(4):421-432. Goovaerts. P. 1997. Geostatistics for natural resource evaluation. Oxford University Press, New-York. 483 pp. 68 Goovaerts. P. 1999. Geostatistics in soil science: state—of-the-art prospective. In press. Geoderrna. Gotway, C.A. R.B. Ferguson, G.W. Hergert, and TA. Peterson. 1996. Comparison of kriging and inverse-distance methods for mapping soil parameters. Soil Sci. Soc. Am. J. 60:1237-1247. Gotway, CA. and AH. Hartford. 1996. Geostatistical methods for incorporating auxiliary information in the prediction of spatial variables. Journal of Agricultural, biological, and Environmental Statistics. 1(1): 17-39. Isaaks, E. and R. Srivastava. 1989. An introduction to applied geostatistics. Oxford University Press, New York, NY. Jaynes, DB. 1996 Improved soil mapping using electromagnetic induction surveys. p. 169-179. In. P.C. Robert et al. (ed.) Proc. 3rd international conference on precision agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. Juang, K.W. and D.Y. Lee. 1998. A comparison of three kriging methods using auxiliary variables in heavy-metal contaminated soils. J. Environ. Qual. 27:355-363. McCann, B.L., D.J. Pennock., C. van Kessel, F.L. Walley. 1996. The development of management units for site-specific farming. In. P.C. Robert et al. (ed.) Proc. 3rd international conference on precision agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. Mohamed, S.B., E.J. Evans, R.S. Shiel. 1996. Mapping techniques and intensity of soil sampling for precision farming. p. 217 - 226. In. P.C. Robert et al. (ed.) Proc. 3rd international conference on precision agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. Moore, I.D., P.E. Gessler, G.A. Nielsen, and GA. Peterson. 1993. Soil attribute prediction using terrain analysis. Soil Sci.Soc. Am. J. 57:443-452. Pannatier, Y. 1996. Variowin: Software for spatial data analysis in 2D. Springer, New York, NY. Pierce, P.J. and Nowak. 1999. Aspects of precision farming. In. Advances in Agronomy. In Press. 69 Pocknee, S. B.C. Boydell, H.M. Green, D]. Waters, C.K. Kvien. 1996. Directed Soil Sampling. In. P.C. Robert et al. (ed.) Proc. 3rd international conference on precision agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. Pregitzer, K.E. 1978. Soil survey of Clinton county, Michigan. Soil Conservation Service. Robertson, G.P., J .R. Crum, B.G. Ellis. 1993. The spatial variability of soil resources following long-terrn disturbances. Oecologia. 96:451-456. Robertson, G.P., K.M. Klingensmith, M.J. Klug, E.A. Paul, J.R. Crum, and B.G. Ellis. 1997. Soil Resources, microbial activity, and primary production across an agricultural ecosystems. Ecological Applications, 7(1): 158-170. Rosenbaum, MS. and M. Stiderstrtim. 1996. Cokriging of heavy metals as an aid to biogeochemical mapping. Acta Agric. Scand. Sect. B, Soil and plant Sci. 46: 1-8. Sadler, E.J., W.J. Busscher, P.J. Bauer, and D.L. Karlen. 1998. Spatial scale requirements for precision farming: a case study in the southeastern USA. Agron. J. 90: 191- 197. SAS Institute 1990. SAS/ STAT user's guide. Version 6. SAS Inst., Cary, NC. Sayer, J.E. 1994. Concepts of variable rate technology with considerations for fertilizer application. J. Prod. Agric., 7(2): 195-201. Snyder, C., T. Schroeder, J. Havlin, and G. Kluitenberg. 1996. An economic anaysis of variable rate nitrogen management. p. 989-998. In P.C. Robert et al. (ed.) Proc. 3rd international conference on precision agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. Stroup, W.W., P.S. Baenziger, and D.K. Multize. 1994. Removing spatial variation from wheat yield trials: a comparison of methods. Crop Science. 34: p. 62-66. Sudduth, K.A., J .W. Hummel, S.J. Birrell. 1997. Sensors for site-specific management. p. 183-210. In F.J. Pierce and E.J. Sadler (ed.) The state of site-specific management for agriculture. ASA Misc. Publ., ASA, CSSA, and SSSA, Madison, WI. 70 Thompson, W.H. and RC. Robert. 1995. Evaluation of mapping strategies for variable rate applications p303-323. In P.C. Robert et al. ed.) Site —specific management for agriculture systems. ASA Misc. Publ. ASA, CSSA, and SSSA, Madison, WI. Threlkeld, G.W. and J .E. F eenstra. 1974. Soil survey of Shiawassee county, Michigan. Soil Conservation service. Vaughan, P.J., S.M. Lesch, D.L. Corwin, and D.G. Cone. 1995. Water content effect on soil salinity prediction: a geostatistical study using cokriging. Soil Sci. Soc. Am. J. 59:1146-1156. Vitosh, M.L, J .W. Johnson, and DB. Mengel: 1995. Tri-state fertilizer recommendations for corn, soybeans, wheat and alfalfa. Michigan State University Ext. Bull. E- 2567. Wibawa, W.D., D.L. Dludlu, L.J. Swenson, D.G. Hopkins, and WC. Dahnke. 1993. Variable fertilizer application based on yield goal, soil fertility, and soil map unit. J. Prod. Agric., 6(2): 255-262. Wollenhaupt, NC, and DD. Buchholz. 1993. Profitability of farming by soils. p. 199- 211. In P.C. Roberts et al. (ed). Soil specific crop management. ASA Misc. Publ. ASA, CSSA, and SSSA, Madison, WI. Wollenhaupt, N.C., R.P. Wolkowski, and MK. Clayton. 1994. Mapping Soil Test phosphorus and potassium for variable-rate fertilizer application. J. Prod. Agric. 7(4): 441 -448. Zhang, R., D.E. Myers, and A.W. Warrick. 1992. Estimation of the spatial distribution of soil chemicals using pseudo-cross-variograrns. Soil Sci. Soc. Am. J. 5621444- 1452. Zhang, R., P. Shouse, and S. Yates. 1997. Use of pseudo-crossvariograms and cokriging to improve estimates of soil solute concentrations. Soil Sci Soc. Am. J. 61 :1342- 1347. 71 MICHIGAN srnre UNIV. LreRnRrEs lllWilliillWillIllllillillllllllllllllllllllllll 31293017665062