169
Caractérisation par spectroscopie et analyse compositionnelle des formes du phosphore dans des sols agricoles canadiens Thèse Dalel Abdi Doctorat en sols et environnement Philosophiae doctor (Ph. D.) Québec, Canada © Dalel Abdi, 2014

Caractérisation par spectroscopie et analyse compositionnelle des

Embed Size (px)

Citation preview

Page 1: Caractérisation par spectroscopie et analyse compositionnelle des

I

Caractérisation par spectroscopie et analyse compositionnelle des formes du phosphore dans des

sols agricoles canadiens

Thèse

Dalel Abdi

Doctorat en sols et environnement

Philosophiae doctor (Ph. D.)

Québec, Canada

© Dalel Abdi, 2014

Page 2: Caractérisation par spectroscopie et analyse compositionnelle des

II

Page 3: Caractérisation par spectroscopie et analyse compositionnelle des

III

RÉSUMÉ

Une meilleure compréhension de la dynamique et des changements des formes du

phosphore (P) dans les sols agricoles est indispensable pour le maintien de leur productivité

et de la qualité des eaux de surface. L’objectif de cette thèse a été de développer et

d’utiliser des méthodes innovatrices pour mesurer les différentes formes de P et caractériser

leurs changements dans des sols soumis à différentes pratiques culturales. Nous avons

évalué le potentiel de la spectroscopie dans le proche infrarouge (SPIR) à prédire le P total

(PT), le P chimiquement extrait à la solution Mehlich-3 (PM3) et à l’eau (Cp) et le P

organique (Po) dans deux sites de teneurs variables en P situés au Québec et en

Saskatchewan. Les résultats obtenus ont démontré que la prédiction du PT et du PM3 dans

le site du Québec est modérément utile et non acceptable, respectivement. Cependant, des

résultats inverses ont été trouvés dans le site du Saskatchewan. La prédiction du Po est de

modérément utile à modérément réussie dans le site du Saskatchewan. Le potentiel de la

prédiction de ces formes du P par la SPIR dépend de la texture du sol, de leur variation

dans le sol et de leur lien à la matière organique. En outre, nous avons démontré que les

résultats de l’analyse de variance et de la corrélation des espèces moléculaires de P, brutes

ou ordinairement transformées, varient en fonction de leur unité de mesure. L’utilisation de

l’analyse compositionnelle avec les transformations du log ratio centré ou du log ratio

isométrique permet d’éviter ce biais et d’avoir des interprétations cohérentes des résultats.

Finalement, l’analyse par résonance magnétique nucléaire des sols sous rotation de maïs et

de soya de longue durée au Québec a démontré que l’accumulation de PT (1326 mg kg-1)

dans la couche superficielle du sol (0-5 cm) soumis à la fertilisation phosphatée et au semis

direct était principalement due aux ions orthophosphates (49,7% du PT). Cependant, les

formes organiques s’accumulaient en profondeur sous forme d’inositols monoesters et de

nucléotides qui sont donc susceptibles d’atteindre les cours d’eaux adjacents par drainage.

Ce projet de recherche nous permet de mieux caractériser et gérer les formes de P dans les

écosystèmes cultivés en adoptant les pratiques culturales adéquates.

Page 4: Caractérisation par spectroscopie et analyse compositionnelle des

IV

Page 5: Caractérisation par spectroscopie et analyse compositionnelle des

V

ABSTRACT

Understanding of phosphorus (P) forms dynamics and changes in agro-ecosystems

is essential for the development of best management practices to maintain soil productivity

and surface water quality. The objective of this thesis was to develop and use innovative

methods to characterize soil P forms and their changes under different management

practices. We examined the potential of near infrared spectroscopy (NIRS) to predict soil

total P (TP), soil P extracted by Mehlich-3 solution (M3P) and by water (Cp), and soil

organic P (OP) for soil samples taken from two sites with different P content located at

Quebec and Saskatchewan. The results showed that the prediction of TP and M3P in the

site of Quebec were moderately useful and not acceptable, respectively. However, the

opposite was found in the site of Saskatchewan. The prediction of OP was moderately

useful to moderately successful in experimental site of Saskatchewan. The potential of

NIRS to predict P depends to the soil texture, to P soil content variation and to the relation

of P to organic matter. Furthermore, contradictory results of variance and correlations

analyses were found for the raw and ordinary log transformed molecular P species

expressed as proportions or concentrations, indicating spurious correlations. Using

compositional analysis with centred log ratio or isometric log ratio transformations avoid

the methodological biases and allow coherent interpretation. Finally, phosphorus-31

nuclear magnetic resonance spectroscopy was used to characterize P species for soil

samples collected from a long-term corn-soybean rotation experiment in Quebec. Results

showed an accumulation of TP (1326 mg kg-1) on the top 5 cm of P fertilized soil under no-

till primarily due to orthophosphate ions accumulation (49.7% of TP). However, the

organic P forms of inositol monoesters and nucleotides had accumulated in the deep layer;

indicating a potential loss through different hydrological pathways. Overall, these studies

allow us to improve our understanding of P forms and to better monitor them under

different agro-ecosystems using the best management practices.

Page 6: Caractérisation par spectroscopie et analyse compositionnelle des

VI

Page 7: Caractérisation par spectroscopie et analyse compositionnelle des

VII

TABLE DES MATIÈRES

RÉSUMÉ .............................................................................................................................. III

ABSTRACT ........................................................................................................................... V

TABLE DES MATIÈRES .................................................................................................. VII

LISTE DES TABLEAUX ................................................................................................... XI

LISTE DES FIGURES ...................................................................................................... XIII

ABRÉVIATIONS ET DÉFINITIONS ............................................................................... XV

DÉDICACE ..................................................................................................................... XVII

REMERCIEMENTS .......................................................................................................... XIX

AVANT-PROPOS ............................................................................................................. XXI

CHAPITRE I: INTRODUCTION .......................................................................................... 1

CHAPITRE II: REVUE DE LA LITTÉRATURE ................................................................. 5

2.1 Cycle biogéochimique du phosphore dans le sol .......................................................... 5

2.2 Les formes du phosphore dans les sols ......................................................................... 6

2.2.1 Phosphore inorganique .......................................................................................... 6

2.2.2 Phosphore organique .............................................................................................. 7

2.3 Mesures du phosphore du sol ...................................................................................... 10

2.3.1 Méthodes conventionnelles .................................................................................. 10

2.3.1.1 Phosphore total ............................................................................................. 10

2.3.1.2 Phosphore disponible aux plantes ................................................................. 10

2.3.1.3 Phosphore organique ..................................................................................... 11

2.3.1.4 Pools du P ..................................................................................................... 12

2.3.2 Méthodes spectroscopiques ................................................................................. 14

2.3.2.1 Spectroscopie dans le proche infrarouge ...................................................... 14

2.3.2.2 Spectroscopie de résonance magnétique nucléaire du 31P ............................ 15

2.4 Changements des formes du phosphore selon les pratiques culturales....................... 16

2.4.1 Changements des pools du P ............................................................................... 16

2.4.2 Changements des espèces de P ............................................................................ 17

2.5 Concept d’analyse des données compositionnelles .................................................... 18

2.6 Hypothèses .................................................................................................................. 21

2.7 Objectifs ...................................................................................................................... 22

2.8 BIBLIOGRAPHIE ...................................................................................................... 23

CHAPITRE III: PREDICTING SOIL PHOSPHORUS-RELATED PROPERTIES USING

NEAR-INRARED REFLECTANCE SPECTROSCOPY .................................................... 31

3.1 RÉSUMÉ .................................................................................................................... 32

3.2 ABSTRACT ................................................................................................................ 33

3.3 INTRODUCTION ...................................................................................................... 34

3.4 MATERIALS AND METHODS ................................................................................ 35

3.4.1 Experimental site description ............................................................................... 35

3.4.2 Soil and plant analyses ......................................................................................... 36

3.4.3 Near-infrared reflectance spectroscopy spectrum acquisition ............................. 37

3.4.4 Pretreatment, calibration, and cross-validation .................................................... 38

3.4.5 Validation ............................................................................................................. 39

3.5 RESULTS AND DISCUSSION ................................................................................. 39

Page 8: Caractérisation par spectroscopie et analyse compositionnelle des

VIII

3.5.1 Reference data ..................................................................................................... 39

3.5.2 Spectral pretreatments ......................................................................................... 40

3.5.3 Near-infrared reflectance spectroscopy prediction of soil and crop P properties 40

3.5.4 Near-infrared reflectance spectroscopy prediction of other soil properties ........ 42

3.6 CONCLUSIONS ........................................................................................................ 43

3.7 ACKNOWLEDGEMENTS ....................................................................................... 43

3.8 REFERENCES ........................................................................................................... 44

CHAPITRE IV: PREDICTING SOIL ORGANIC PHOSPHORUS USING NEAR-

INFRARED REFLECTANCE SPECTROSCOPY ............................................................. 55

4.1 RÉSUMÉ .................................................................................................................... 56

4.2 ABSTRACT ............................................................................................................... 57

4.3 INTRODUCTION ...................................................................................................... 58

4.4 MATERIALS AND METHODS ............................................................................... 59

4.4.1 Experimental site description .............................................................................. 59

4.4.2 Soil sampling and analysis .................................................................................. 59

4.4.3 Near-infrared reflectance spectroscopy spectrum acquisition ............................ 60

4.4.4 Spectral pre-treatment ......................................................................................... 60

4.4.5 Calibration, cross-validation and validation ........................................................ 60

4.5 RESULTS AND DISCUSSION ................................................................................ 61

4.5.1 Soil reference data ............................................................................................... 61

4.5.3 Spectral pre-treatment, calibration, and prediction of soil organic P .................. 62

4.5.4 Spectral pre-treatment, calibration, and prediction of soil total and Mehlich-3 P

...................................................................................................................................... 62

4.5.5 Spectral pre-treatment, calibration, and prediction of soil organic matter and

Mehlich-3 nutrients ...................................................................................................... 63

4.6 CONCLUSION .......................................................................................................... 64

4.7 REFERENCES ........................................................................................................... 65

CHAPITRE V: UNBIASED STATISTICAL ANALYSIS OF SOIL 31P-NMR ................. 73

5.1 RÉSUMÉ .................................................................................................................... 74

5.2 ABSTRACT ............................................................................................................... 75

5.3 INTRODUCTION ...................................................................................................... 76

5.4 MATERIALS AND METHODS ............................................................................... 77

5.4.1 Datasets ............................................................................................................... 77

5.4.1.1 Compositional data transformations ................................................................. 78

5.4.1.1.1 Centred log-ratio transformation ............................................................... 78

5.4.1.1.2 Isometric log-ratio transformation ............................................................ 78

5.4.1.1.3 Choice of SBP ........................................................................................... 79

5.4.1.1.4 Ordinary logarithmic transformation ........................................................ 80

5.4.1.2 Statistical analysis ............................................................................................ 81

5.5 RESULTS AND DISCUSSION ................................................................................ 81

5.5.1 Biased analysis of variance ................................................................................. 81

5.5.2 Spurious correlations ........................................................................................... 83

5.6 CONCLUSIONS ........................................................................................................ 84

5.7 REFERENCES ........................................................................................................... 85

CHAPITRE VI: LONG-TERM IMPACT OF TILLAGE PRACTICES AND P

FERTILIZATION ON SOIL P FORMS AS DETERMINED BY 31P NUCLEAR

MAGNETIC RESONANCE SPECTROSCOPY ................................................................ 99

Page 9: Caractérisation par spectroscopie et analyse compositionnelle des

IX

6.1 RÉSUMÉ .................................................................................................................. 100

6.2 ABSTRACT .............................................................................................................. 101

6.3 INTRODUCTION .................................................................................................... 102

6.4 MATERIELS AND METHODS .............................................................................. 104

6.4.1 Experimental site ............................................................................................... 104

6.4.2 Soil sampling and chemical analysis ................................................................. 105

6.4.3 Solution 31P-NMR spectroscopy ........................................................................ 105

6.4.4 Compositional data analysis .............................................................................. 106

6.4.5 Statistical analysis .............................................................................................. 107

6.5 RESULTS ................................................................................................................. 107

6.5.1 Chemical soil properties .................................................................................... 107

6.5.2 Identification of phosphorus forms by 31P nuclear magnetic resonance

spectroscopy ................................................................................................................ 108

6.5.3 Distribution of 31P nuclear magnetic resonance phosphorus forms ................... 109

6.6 DISCUSSION ........................................................................................................... 110

6.7 CONCLUSIONS ...................................................................................................... 113

6.8 ACKNOWLEDGMENTS ........................................................................................ 113

6.9 REFERENCES ......................................................................................................... 114

CHAPITRE VII: CONCLUSIONS ET RECOMMANDATIONS .................................... 127

CHAPITRE VIII: ANNEXE .............................................................................................. 131

COMPOSITIONAL ANALYSIS OF POOLS IN CANADIAN MOLLISOLS ................ 131

8.1 RÉSUMÉ .................................................................................................................. 132

8.2 ABSTRACT .............................................................................................................. 133

8.3 INTRODUCTION .................................................................................................... 134

8.4 MATERIALS AND METHODS .............................................................................. 135

8.4.1 Isometric log ratio transformation and the Aitchison distance .......................... 136

8.4.2 The Mackenzie et al. (1992) dataset .................................................................. 137

8.4.3 The Tiessen et al. (1983) dataset ....................................................................... 138

8.5 CONCLUSION ......................................................................................................... 138

8.6 REFERENCE ............................................................................................................ 140

Page 10: Caractérisation par spectroscopie et analyse compositionnelle des

X

Page 11: Caractérisation par spectroscopie et analyse compositionnelle des

XI

LISTE DES TABLEAUX

Tableau 2-1 Composés organiques phosphatés communs du sol (Turner et al., 2005). ........ 9

Tableau 3-1 Descriptive statistics† for the soil P-related and other properties analyzed

using reference methods. .............................................................................................. 48

Tableau 3-2 NIRS spectral pre-treatments and statistics† of calibration, cross-validation,

and validation for the P-related soil properties. ............................................................ 49

Tableau 3-3 NIRS spectral pre-treatment and statistics† of calibration, cross-validation,

and validation for the other soil properties. .................................................................. 50

Tableau 4-1 Descriptive statistics for the soil organic (OP) and total (TP) P analysed for

long- and short-term no-till (NT) treatments. ............................................................... 67

Tableau 4-2 Descriptive statistics for the soil Mehlich-3 extracted nutrients and organic

matter for long- and short-term no-till (NT) treatments. .............................................. 67

Tableau 4-3 Statistics of near-infrared reflectance spectroscopy calibration, cross-

validation, and validation for soil (OP) and total (TP) P analysed for long- and short-

term no-till (NT) treatments. ......................................................................................... 68

Tableau 4-4 Statistics of near-infrared reflectance spectroscopy calibration, cross-

validation, and validation for soil Mehlich-3 extracted nutrients and organic matter for

long- and short-term no-till treatments. ....................................................................... 69

Table 5-1 Sequential binary partition of soil 31P-NMR P species analyzed by Abdi et al.

(2014). ........................................................................................................................... 88

Table 5-2 Sequential binary partition of soil 31P-NMR P species analyzed by Cade-Menun

et al. (2010). .................................................................................................................. 89

Table 5-3 ANOVA of the effect of tillage (T), P fertilization (P) and soil depth (D) on log-

and clr-transformed P compositions (Abdi et al., 2014). P species defined in Table 1.

...................................................................................................................................... 90

Table 5-4 ANOVA of the effect of tillage and soil depth on log- and clr-transformed P

compositions (Cade-Menun et al., 2010). P species defined in Table 2. ...................... 92

Table 6-1 Analysis of variance for the effects of tillage, P fertilization and depth on clr

transformed concentrations of soil total P (TP), Mehlich-3 extractable P (PM3),

aluminium (Al), iron (Fe), calcium (Ca), magnesium (Mg), total carbon (TC) and total

nitrogen (TN), and pH. ............................................................................................... 118

Table 6-2 Chemical shift of P forms detected in the 31P-NMR spectrum of the soil as

affected by tillage and P fertilization management and depth. ................................... 119

Table 6-3 Analysis of variance for the effects of tillage, P fertilization, and depth on

centered log ratio–transformed soil P forms determined by 31P nuclear magnetic

resonance spectroscopy. .............................................................................................. 120

Table 6-4 Back-centered log ratio–transformed soil P forms determined by 31P nuclear

magnetic resonance spectroscopy as affected by tillage, P fertilization, and depth. .. 121

Page 12: Caractérisation par spectroscopie et analyse compositionnelle des

XII

Table 8-1 Sequential binary partitions of soil P fractions (r is number of P fractions with

plus sign and s is number of P fractions with minus sign). ........................................ 141

Table 8-2 Mollisol P fractions following crop sequence and NP fertilization (data from

McKenzie et al., 1992). .............................................................................................. 142

Table 8-3 Ilr coordinates of P pools following crop sequence and fertilization (data from

McKenzie et al., 1992). .............................................................................................. 143

Table 8-4 Ilr differences in P pools between treatments and uncultivated check (data from

McKenzie et al., 1992). .............................................................................................. 144

Table 8-5 Effect of time on P pools in a Mollisol (data from Tiessen et al., 1984). ......... 145

Page 13: Caractérisation par spectroscopie et analyse compositionnelle des

XIII

LISTE DES FIGURES

Figure 1-1. Schéma général du projet de recherche ............................................................... 4

Figure 2-1 Cycle du P dans le sol (adapté du Pierzynski et al., 2005). .................................. 5

Figure 2-2 Effet du pH sur la forme des ions orthophosphates de la solution du sol (Holtan

et al., 1988). .................................................................................................................... 7

Figure 2-3 Méthode d’extraction séquentielle du P du sol selon la méthode de Hedley et

al., 1982 (tirée de Cross et Schlesinger, 1995). ............................................................ 13

Figure 2-4 Spectre obtenu par la spectroscopie magnétique nucléaire du 31P montrant les

composés phosphatés détectés dans la couche superficielle (5-10 cm) d’un sol cultivé

non labouré (Cade-Menun et al., 2010). ....................................................................... 16

Figure 3-1 NIRS predicted values against measured values of (a) soil P content extracted

using the Mehlich 3 method and analysed by colorimetry (M3P_Col); (b) soil P

content extracted using the Mehlich 3 method and analysed by ICP (M3P_ICP); (c)

soil P content extracted with water and analysed by colorimetry (Cp); (d) total soil P;

(e) annual timothy crop P-uptake, and; (f) annual P-budget. Based on validation

statistics reported here and in Table 3.2, NIRS predictions were considered moderately

useful (MU) when 0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25, and less reliable (LR)

when Rv2 < 0.70 and RPD < 1.75 (Malley et al., 2004). ............................................... 53

Figure 4-1 Near-infrared reflectance spectroscopy (NIRS) predicted vs. measured values

of soil organic P analysed for (a) long- and short term no-till, (b) long-term no-till, and

(c) short-term no-till treatments. ................................................................................... 70

Figure 5-1 Relationship between Mahalanobis distance from ilr with (a) ordinary log

transformed 31P NMR-P species concentration, and (b) raw of 31P NMR-P species

concentrations (data from Abdi et al., 2014). ............................................................... 97

Figure 6-1 Distribution of total P (TP), Mehlich-3 extractable P (PM3) and orthophosphate

concentrations at various soil depths under (a, c, e) mouldboard plow (MP) and (b, d,

f) no-till (NT) treatments. P0 and P35 represent additions of 0 and 35 kg P ha−1,

respectively. Values are means of three replications. For each treatment, different

letters indicate significantly different means among soil depth according to LSD

(0.05). † For each treatment, different letters indicate significantly different means

among depth according to LSD (0.1). ......................................................................... 122

Figure 6-2 Distribution of (a) total carbon (TC) and (b) total nitrogen (TN) content, and (c)

Al Mehlich-3 and (d) Mg Mehlich-3 at various soil depths under mouldboard plow

(MP) and no-till (NT) treatments. Values are means of three replicates. For each

Page 14: Caractérisation par spectroscopie et analyse compositionnelle des

XIV

treatment, different letters indicate significantly different means among soil depth

according to LSD (0.05). ............................................................................................ 123

Figure 6-3 Phosphorus-31 nuclear magnetic resonance spectroscopy spectrum showing the

range of P compounds detected at the 0 to 5 cm depth of the mouldboard plow

fertilized treatment (Oth.D1, other diester 1; Oth.D2, other diester 2). ..................... 124

Figure 6-4 Phosphorus-31 nuclear magnetic resonance spectroscopy spectrum showing the

P compounds detected in the monoester region at the 0 to 5 cm depth of mouldboard

plow fertilized treatment. (A) neo-IP6; (B) orthophosphate; (C) myo-IP6; (D) glucose-

6P; (E) unknown; (F) α-glycerophosphate; (G) β-glycerophosphate; (H) nucleotides;

(I) choline-P; (J) scyllo-IP6; (M1) monoester 1; (M2) monoester 2. .......................... 125

Figure 6-5 Distributions of (a) pyrophosphate, (b) scyllo-IP6, (c) DNA and (d) nucleotides

concentrations at various soil depths under mouldboard plow (MP) and no-till (NT)

treatments. Values are means of three replicates. For each treatment, different letters

indicate significantly different means among depths according to LSD (0.05). ........ 126

Figure 8-1 Conceptual relational model between P pools in Mollisols (modified from

Tiessen et al., 1984). ................................................................................................... 146

Figure 8-2 Time change in P balance distances from initial conditions in a Blaine lake soil

(data from Tiessen et al., 1983)……………………………………………….……147

Page 15: Caractérisation par spectroscopie et analyse compositionnelle des

XV

ABRÉVIATIONS ET DÉFINITIONS

clr: Log ratio centré (centred log ratio)

Cp: Concentration des ions phosphates dans la solution du sol estimée par une

extraction à l’eau

CV: Coefficient de variation (coefficient of variation)

ADN : Acide désoxyribonucléique (DNA: Deoxyribonucleic acid)

Espèce de P: Composé ionique ou moléculaire de P

Forme de P: Désigne la nature chimique du P (Pi, Po)

ilr: Log ratio isométrique (isometric log ratio)

IP6: Inositol Hexakisphoaphate

MP: Labour conventionnel (Mouldboard plow)

NT: Semis direct (No-till)

P: Phosphore

Pi: Phosphore inorganique

PM3: Phosphore disponible aux plantes estimé avec la méthode Mehlich-3

Po: Phosphore organique

Pools de P: Fractions du P total déterminées par une extraction séquentielle

PT: Phosphore total

R2: Coefficient de détermination (Coefficient of determination)

RMN-31P: Spectroscopie de resonance magnétique nucléaire du 31P (31P-NMR :

Phosphorus-31 nuclear magnetic resonance)

RPD: Rapport de déviation de la performance de prédiction (Ratio of standard error of

prediction to standard deviation)

SPIR: Spectroscopie dans le proche infrarouge (NIRS: Near-infrared reflectance

spectroscopy)

Page 16: Caractérisation par spectroscopie et analyse compositionnelle des

XVI

Page 17: Caractérisation par spectroscopie et analyse compositionnelle des

XVII

DÉDICACE

À ma famille et à ma Tunisie !

Page 18: Caractérisation par spectroscopie et analyse compositionnelle des

XVIII

Page 19: Caractérisation par spectroscopie et analyse compositionnelle des

XIX

REMERCIEMENTS

Je tiens à remercier vivement ma co-directrice docteur Noura Ziadi, chercheure au

Centre de recherche et de développement sur les sols et les grandes cultures d’Agriculture

et Agroalimentaire Canada, pour la confiance qu’elle m’a accordée en acceptant de diriger

mes travaux de recherches, pour son support scientifique, sa disponibilité, ses conseils, sa

patience, et son encouragement continu sans lesquels cette thèse n’aurait pas été possible.

Je tiens à l’assurer de ma profonde gratitude pour les facilités et les opportunités qu’elle

m’a permis de bénéficier.

Mes remerciements les plus sincères vont également à mon directeur de thèse docteur

Léon-Étienne Parent pour m’avoir fait bénéficier de sa grande compétence et de sa rigeur

scientifique tout au long de mon programme de doctorat. Les discussions que nous avons

eues ensemble et son esprit critique et innovateur m’ont toujours motivée et m’ont permis

de progresser avec beaucoup de succès.

Je suis très honorée à remercier docteur Alfred Jaouich pour avoir accepté de faire la

prélecture de ma thèse. Ses commentaires et suggestions m’ont permis d’améliorer

beaucoup la qualité de cette thèse.

J’adresse mes remerciements aussi aux Dr. Judith Nyiraneza et Dr. Christian Morel

d’avoir accpeté de consacrer du temps pour l’évaluation de cette thèse. Leurs remarques et

suggestions ont permis l’amélioration de la version finale.

Je suis très reconnaisante au docteur Barbara J. Cade-Menun, chercheure au Centre de

recherche sur l’agriculture des Prairies semi-arides d’Agriculture et Agroalimentaire

Canada, pour son acceuil bienveillant durant mon stage et pour m’avoir fait bénéficier de sa

compétence distinguée dans l’étude du phosphore organique, et pour m’avoir formée au

traitement des spectres de la résonance magnétique nucléaire.

Je voudrais aussi adresser ma gratitude aux docteurs Gaëtan F. Tremblay et Gilles

Bélanger, chercheurs au Centre de recherche et de développement sur les sols et les grandes

cultures d’Agriculture et Agroalimentaire Canada, pour leurs remarques et orientations

Page 20: Caractérisation par spectroscopie et analyse compositionnelle des

XX

dans les études faites par la spectroscopie dans le proche infrarouge. Nos nombreuses

discussions étaient à la fois enrichissantes et intéressantes.

Je souhaiterais remercier tout particulièrement Sylvie Michaud, Sylvie Côté, Mario

Laterrière, Bernard Gagnon et Claude Levesque des laboratoires du Centre de recherche et

de développement sur les sols et les grandes cultures d’Agriculture et Agroalimentaire

Canada, pour leur support technique et leur gentillesse.

Je tiens à remercier également Dr. Serge-Étienne Parent, professionnel de recherche en

agrostatistiques à l’Université Laval, pour sa précieuse aide aux analyses mathématiques et

statistiques.

Je remercie le Ministère de l’Enseignement Supérieur et de la Recherche Scientifique

de Tunisie pour le soutient financier de cette thèse. Je tiens à remercier aussi les organismes

canadiens qui m’ont accordée des bourses pour présenter mes résultats de recherches dans

des congrès scientifiques, notamment le centre SÈVE, l’Association Québécoise des

Spécialiste en Sciences du sol et le Bureau des bourses et de l’aide financière de

l’Université Laval.

Je remercie chaleureusement mes amis Aimé Messiga, Yichao Shi, et Mervin St. Luce

pour les discussions qu’on a eues ensemble et pour l’ambiance du travail très agréable.

Enfin, j’adresse mes vifs remerciements à mes parents et à tous les membres de ma

famille pour leur amour, leur confiance et leur encouragement. Un grand merci pour

m’avoir fait de moi ce que je suis aujourd’hui !

Page 21: Caractérisation par spectroscopie et analyse compositionnelle des

XXI

AVANT-PROPOS

Cette thèse est le fruit d’un travail de recherche doctoral qui s’inscrit dans le cadre

d’une coopération entre l’Université Laval et le Centre de recherche et de développement

sur les sols et les grandes cultures d’Agriculture et Agroalimentaire à Québec, Canada.

Dans cette thèse, nous présentons huit chapitres dont le premier et le deuxième

consistent à une introduction générale et à une revue de littérature sur l’ensemble des

thèmes de recherches à l’étude. Les quatre chapitres qui suivent ont fait l’objet de : deux

articles scientifiques publiés dans « Soil Science Society of America Journal » (chapitre 3)

et « Journal of Environmental Quality » (chapitre 6), et un article soumis à Geoderma

(chapitre 5). Le chapitre 4 sera aussi soumis à Geoderma. Le chapitre 7 présente la

conclusion générale pour l’ensemble des études et les recommandations, et le dernier

chapitre présente l’annexe déjà publié suite à un congrès qui a lieu en Espagne.

Tous ces articles ont été rédigés en anglais et sont insérés dans cette thèse comme

publiés ou soumis. Un court résumé en français (envrion 150 mots selon les exigences de

rédaction de thèse de l’Université Laval) précède chacun d’eux. Je suis l’auteure principale

de chaque article et la responsable des analyses de laboratoire, de traitement de données et

d’interprétation des résultats. Les co-auteurs sont le docteur Léon-Étienne Parent de

l’Université Laval et d’autres chercheurs des centres de recherche d’Agriculture et

Agroalimentaire Canada situés à Québec, Qc, et à Swift Current, SK.

Les résultats de ces études ont été présentés dans des congrès nationaux (Québec) et

internationaux en Espagne, Italie, Suède, Panama, France et en Corée du Sud pour un total

de sept communications orales et neuf affiches. Un résumé du chapitre 6 a été publié

également dans « Crops Soils Agronomy News » de la revue américaine « American

Society of Agronomy » en Juillet, 2014.

Page 22: Caractérisation par spectroscopie et analyse compositionnelle des
Page 23: Caractérisation par spectroscopie et analyse compositionnelle des

1

CHAPITRE I: INTRODUCTION

Le phosphore (P) est un élément naturel qui se trouve dans tout organisme vivant. Il

est impliqué dans plusieurs fonctions vitales telles que les métabolismes énergétiques

(AMP, ADP, ATP) et la constitution des acides nucléiques (ADN, ARN) et des membranes

cellulaires (phospholipides). Chez les plantes, le P joue un rôle important dans la

photosynthèse, la floraison et la production des fruits (Morel, 1989).

Dans la nature, le P est peu abondant; il constitue environ 0,12% des éléments de

l’écorce terrestre (Cathcart, 1980). Il se trouve dans le sol, les sédiments et les eaux de

surface en provenance de la désagrégation des roches minérales phosphatées,

majoritairement de l’apatite (Stevenson, 1986). Les principales réserves naturelles en

phosphate se trouvent au Maroc-Sahara Occidental (74,6%), en Chine (5,5%), en Algérie

(3,2%), en Syrie (2,7%), en Afrique du sud (2,2%), en Jordanie (1,9%), en Russie (1,9%) et

aux États-Unis (1,6%) (U.S. Geological Survey, 2014). Bien qu’elles soient abondantes, ces

réserves en P sont en voie d’épuisement face à l’accroissement de l’industrie des engrais

phosphatés (Cordell et al., 2009).

La forme naturelle du P (apatite, variscite, strengite ou vivianite,..) est peu

disponible aux plantes étant donné qu’elle est peu soluble à l’eau, d’où la nécessité de la

fertilisation phosphatée. Cependant, seulement une faible proportion du P ajouté est

assimilée par la plante et le reste est rapidement fixé par les argiles ou les sesquioxydes de

fer et d’aluminium (Khiari et Parent, 2005). D’autre part, l’intensification des apports des

fertilisants phosphatés en agriculture est à l’origine des pertes en P qui atteignent les cours

d’eau et perturbent leur équilibre écologique (Liu et Chen, 2008). Le maintien de la

durabilité des écosystèmes agricoles et l’optimisation de leur productivité nécessitent une

meilleure compréhension du fonctionnement du cycle biogéochimique du P.

Dans les sols cultivés, le P se trouve sous des formes inorganiques et organiques. Le

P inorganique existe sous forme d’ions phosphates libres en solution, adsorbés ou précipités

dans les minéraux apatites. Les plantes prélèvent le P essentiellement dans la solution du

sol. En absence de tout apport de P sous forme d’engrais, le réapprovisionnement de la

solution du sol se fait par dissolution des minéraux apatites et/ou par minéralisation du P

Page 24: Caractérisation par spectroscopie et analyse compositionnelle des

2

organique. Beaucoup d’études ont révélé l’importance de la contribution du P organique

dans la nutrition des plantes (Firsching et Claassen, 1996; Oehl et al., 2001; Chen et al.,

2002). Le changement de P de la forme non disponible à court terme à la forme disponible

dépend de la nature chimique des formes du P dans le sol et de leur concentration.

La dynamique des formes de P et leur distribution dans le sol sont contrôlées par la

pédogenèse (Tiessen et al., 1984) et les pratiques agricoles (Ross et al., 1999, Negassa et

Leinweber, 2009). Par exemple, le P s’accumule sous la forme d’ions phosphates dans la

couche superficielle des sols sous semis direct et augmente le risque de transport de P par

ruissellement vers les eaux de surface (Sharpley et Smith, 1994). Dans les sols labourés, le

P associé aux particules du sol peut être facilement transporté par érosion vers les cours

d’eau. Selon la texture du sol et les conditions climatiques, le P peut aussi être transporté de

la surface vers les couches de profondeur par écoulement préférentiel (Simard et al., 2000).

L’étude de la dynamique des formes de P et leur distribution dans le sol peut aussi

nous permettre d’améliorer notre compréhension de la contribution du P du sol à la

nutrition minérale des plantes. En effet, il a été démontré dans des études faites au Québec

que l’apport de P sous forme d’engrais ne permettait pas d’augmenter le rendement en grain

du soja (Glycine max L.) ou du maïs (Zea mays L.) dans des sols sous soumis direct ou

labour conventionnel (Messiga et al., 2010) et le rendement de la fléole des prés (Phleum

pratense L.; Bélanger et Ziadi, 2008). D’autres études ont rapporté que les apports continus

des engrais phosphatés ou des amendements organiques augmentaient la fraction du P

organique disponible à long terme (Shi et al., 2013; Zhang et MacKenzie, 1997).

Cependant, la mise en culture des sols sans apports de P peut diminuer leurs teneurs en P

organique (Tiessen et al., 1982).

La dynamique des formes de P et leur distribution dans les sols ont jusqu’ici été

étudiées par la méthode d’extraction séquentielle de Hedley et al. (1982). L’une des

contraintes de cette méthode est la caractérisation opérationnelle des fractions de P suivant

leur solubilité à des extractifs chimiques. Une autre est son analyse statistique étant donné

la contrainte à 100% des fractions de P obtenues. La majorité des études utilisant cette

méthode présentait des quantités brutes des pools de P sans ne leur attribuer aucun sens

biologique, géochimique ou environnemental. L’utilisation du « pathways analysis » a

Page 25: Caractérisation par spectroscopie et analyse compositionnelle des

3

permis d’établir des relations entre les différents pools de P et notamment l’effet des

pratiques agricoles sur celles-ci (Tiessen et al., 1984; Zheng et al., 2002). Cependant, le

« pathways analysis » repose essentiellement sur des coefficients de corrélations de Pearson

qui peuvent varier en fonction du temps dans un système dynamique. De plus, de fausses

corrélations se produisent en raison de la redondance d’information dans un système clos à

100% et de la dépendance d’échelle de mesure (pourcentage par rapport au P total ou à la

matière sèche du sol) (Aitchison, 1986). Il devient alors difficile d’interpréter sans biais les

transformations du P dans les sols à diverses échelles de temps sans tenir compte de la

contrainte à 100% des données compositionnelles.

L’application de nouvelles méthodes d’analyses telle que l’analyse des données

compositionnelles utilisant les coordonnées du log ratio isométrique (ilr) calculées pour des

répartitions binaires des formes de P (Egozcue et al., 2003) peut complémenter les

connaissances obtenues sur les sols à l’aide du « pathways analysis » en décrivant les

interactions entre les différentes fractions du P. L’analyse des données compositionnelles

est un domaine récent des mathématiques appliquées consacré à l’analyse des données

strictement positives comprises entre zéro et une quelconque unité ou échelle de mesure.

Dans un essai de longue durée sur un loam sableux dans la province canadienne de

l’Ile du Prince-Édwouard, il a été démontré à l’aide de la résonance magnétique nucléaire

au 31P (RMN-31P) que la plupart des formes organiques de P ne contribuaient pas à

l’accumulation de P dans la couche superficielle sous le semis direct parce qu’elles sont

plus facilement emportées vers les couches de profondeur par lixiviation (Cade-Menun et

al., 2010). La RMN-31P a l’avantage d’identifier aussi bien les formes inorganiques

qu’organiques. L’utilisation de la RMN couplée aux analyses chimiques (Mehlich, 1984;

Hedley et al., 1982) peut permettre d’améliorer notre compréhension du cycle

biogéochimique de P dans les sols du Québec.

La quantification du P du sol et l’estimation de sa disponibilité aux plantes est un

important aspect de la réussite de systèmes de cultures à la fois productifs et sains pour

l’environnement. La majorité des méthodes d’analyse du P (digestion acide, ignition,

Olsen, Bray 1, Bray 2 et Mehlich-3) sont relativement lentes et utilisent des extractifs

chimiques. Le développement de nouvelles techniques ou outils d’analyse du sol visant à

Page 26: Caractérisation par spectroscopie et analyse compositionnelle des

4

maximiser la gestion des éléments n’a cessé de progresser, et ce, pour des raisons

économiques et environnementales. L’utilisation de la spectroscopie dans le proche

infrarouge (SPIR) pour prédire la composition chimique des sols est de plus en plus

considérée. Les récentes études effectuées au Québec sur la SPIR (Nduwamungu et al.,

2009a) ont démontré l’efficacité de cette technique pour, notamment, prédire la texture,

l’azote potentiellement minéralisable et le carbone du sol. Nduwamungu et al. (2009b) ont

aussi réussi à prédire les concentrations en Ca et Mg extraits au Mehlich-3 en utilisant la

SPIR.

Ce projet de recherche a eu pour objectif général de mesurer les formes de P et de

caractériser leurs changements dans les sols sous l’effet de différentes pratiques culturales

(rotation, fertilisation minérale phosphatée, travail du sol), en utilisant de nouvelles

méthodes spectroscopiques (SPIR, RMN-31P) et mathématique (analyse compositionnelle),

tel que présenté dans la Figure 1.1. Une brève revue de littérature sur les connaissances

actuelles sur les méthodes de mesure et les transformations de différentes formes du P dans

les sols agricoles sera présentée en premier lieu, suivie par les hypothèses et les objectifs de

cette étude.

Figure 1-1. Schéma général du projet de recherche

Page 27: Caractérisation par spectroscopie et analyse compositionnelle des

5

CHAPITRE II: REVUE DE LA LITTÉRATURE

2.1 Cycle biogéochimique du phosphore dans le sol

La dynamique du phosphore (P) dans les écosystèmes cultivés est représentée par

un cycle biogéochimique (Fig. 2.1) qui intègre les réserves et les mécanismes de

transformation du P dans le sol, et les différents flux impliqués dans le transfert du P vers

les compartiments de l’écosystème.

Le P peut être ajouté au sol sous forme d’engrais minéraux ou de ferme (fumier,

lisiers), d’amendements organiques (composts, boues d’épuration, biosolides papetiers), ou

de résidus de cultures. Dans le sol, le P existe sous différentes formes qui interagissent via

différents mécanismes physico-chimiques, biologiques et biochimiques impliquant des

réactions d’adsorption et de désorption, de précipitation et de dissolution, de minéralisation

et d’immobilisation (Fig. 2.1). Le cycle biogéochimique du P inclut aussi des flux de P sous

forme de prélèvements par les plantes, et des pertes par érosion, ruissellement de surface et

de profondeur, lessivage et drainage (Kleinman et al., 2009; Haygarth et al, 1998).

Figure 2-1 Cycle du P dans le sol (adapté du Pierzynski et al., 2005).

P adsorbé Argile, Oxydes d’Al, Fe Carbonates de Ca

Minéraux secondaires Phosphates de Fe, Al et Ca

Minéraux primaires Apatites

P dans la solution du sol (H 2 PO 4

- , HPO 4

2 - )

P organique Biomasse microbienne Matière organique P organique dissous

Résidus végétaux

Lessivage

Eaux de surface

Érosion, ruissellement

Dissolution

Immobilisation

Minéralisation

Adsorption

Désorption

Précipitation

Dissolution

Prélèvements par la plante

Engrais organiques Engrais minéraux

Page 28: Caractérisation par spectroscopie et analyse compositionnelle des

6

2.2 Les formes du phosphore dans les sols

Dans la couche superficielle des sols (0-15 cm), le P total (PT) varie entre 50 et 3000 mg

kg-1 dépendamment de la roche mère, du type de sol, de la nature de la végétation et de

l’aménagement du sol (Sims et Pierzynski, 2005). Dans les sols minéraux, 35 à 70% du PT

se trouve sous forme inorganique et le reste est sous forme organique. Le P organique peut

présenter jusqu’à 90% du PT dans les sols organiques (Harrison, 1987).

2.2.1 Phosphore inorganique

Le phosphore inorganique (Pi) existe en plus fortes proportions dans les sols cultivés que

dans les sols de prairies ou forestiers. À titre d’illustration, il constitue dans les sols de

grandes cultures environ 75% du P total (Morel, 2002). Dans la solution du sol, il est

présent sous forme d’ions orthophosphates libres dont la nature et la proportion dépendent

du pH du sol (Fig. 2.2). Pour un pH entre 4.0 et 9.0 pour la majorité des sols, les formes

dominantes sont les orthophosphates H2PO4- et HPO4

2- (Pierzynski et al., 2005).

Dans les sols agricoles, la concentration des ions du P en solution est de l’ordre de 0.01 à

3.0 mg P L-1 (Frossard et al., 2000). Les transferts de ces ions à l’interface sol-solution sont

régulés par différents mécanismes chimiques, physiques et biologiques dont l’adsorption-

désorption, la précipitation-dissolution, la minéralisation-immobilisation, la diffusion intra-

particulaire et plusieurs autres réactions (Holtan et al., 1988; Frossard et al., 2000;

McGechan et Lewis, 2002; Oberson et Joner, 2005). Ces divers processus qui conditionnent

la mobilité du P, sont responsables du maintien de l’équilibre de la concentration des ions

phosphates dans la solution du sol, et par la suite, de leur disponibilité aux plantes et aux

microorganismes (Hinsinger, 2001; Hinsinger et al., 2007). En effet, suite au prélèvement

des ions P par les plantes, la concentration du P soluble diminue dans la solution du sol et le

P adsorbé diffuse vers la solution. La dissolution du P de la phase solide et la minéralisation

du Po par les microorganismes constituent aussi des sources importantes de

réapprovisionnement de la solution du sol en ions P (Hinsinger, 2001). Cependant, les ions

orthophosphates peuvent être aussi immobilisés par la biomasse microbienne du sol qui

représente alors un compétiteur pour les racines de la plante.

Page 29: Caractérisation par spectroscopie et analyse compositionnelle des

7

À l’interface solide-solution du sol, les ions phosphates se trouvent adsorbés à la surface

des minéraux argileux, du fer et d’aluminium, des cations échangeables et du carbonate de

calcium (Holford, 1997). Dans les sols acides, les constituants les plus responsables de la

fixation du P sont les oxydes et les hydroxydes de fer et d’aluminium. Tandis que dans les

sols neutres à alcalin, les facteurs les plus impliqués sont le carbonate de calcium et le

magnésium de façon moins importante (Holford, 1997; Hinsinger, 2001). A l’état solide, le

Pi précipite avec les métaux du sol tels que l’Al, le Fe et le Ca en formant des minéraux

phosphatés plus ou moins cristallisés (Hinsinger, 2001). Il est à noter que seulement 10 à

20% du P ajouté sous forme d’engrais est prélevé par la plante (Richardson, 2001;

McLaughlin et al., 1988) puisqu’une grande partie de la dose appliquée évolue rapidement

vers des formes fixées (Khiari et Parent, 2005).

Figure 2-2 Effet du pH sur la forme des ions orthophosphates de la solution du sol (Holtan

et al., 1988).

2.2.2 Phosphore organique

Le phosphore organique (Po) est défini comme étant le P trouvé dans des composés

organiques en liaison avec le carbone. Dans les sols sous prairie et forêts, le P organique

représente 41 à 88% du P total (Bowman et Cole, 1978; Ross et al., 1999; Motavalli et

Miles, 2002; Chen et al., 2003). Dans les sols de grandes cultures, cultivés pendant

plusieurs décennies, il ne représente que 20% du PT en moyenne (Morel, 2002).

Page 30: Caractérisation par spectroscopie et analyse compositionnelle des

8

Le P organique se trouve dans les végétaux, les microorganismes et la matière

organique du sol (Harrison, 1982; Stewart et Tiessen, 1987). Il est synthétisé par les plantes

et les microorganismes via des processus biochimiques à partir des ions orthophosphates

absorbés de la solution du sol (Condron et al., 2005). Outre les résidus de cultures et les

amendements organiques de ferme (fumier, compost) et des industries (biosolides), le Po

peut être introduit aux sols cultivés sous forme synthétisée d’insecticides, fongicides ou

herbicides tel que le phosphonate glyphosate (Condron et al., 2005). Les stocks en Po dans

le sol constituent une source importante du P disponible aux plantes et aux

microorganismes via des processus biotiques (Quiquampoix et Mousain, 2005) et

abiotiques (Baldwin et al., 2005). Le taux de libération des ions phosphates à travers ces

mécanismes de dégradation dépend de la nature chimique des composés organiques

(Quiquampoix et Mousain, 2005).

Selon la nature des liaisons, les composés organiques phosphatés sont classés en

esters phosphate (monoester et diester), en phosphonates (composés avec liaison directe

carbone-phosphore), et en polyphosphates organiques (Mckelvie, 2005; Condron et al.,

2005; Turner et al., 2005). Quelques exemples de composés sont présentés dans le tableau

2.1. Les orthophosphates monoesters sont caractérisés par une seule liaison avec un radical

organique et représentent la forme la plus abondante du Po dans le sol (>90%, Turner et al.,

2003). Cependant, les orthophosphates diesters sont liés à deux radicaux et sont beaucoup

moins présents que les monoesters dans les sols agricoles (<10%, Condron et al., 2005).

Les principales formes de phosphate d’esters rencontrés dans les sols sont les inositols

phosphates (60%), les acides nucléiques (5-10%), et les phospholipides (1%) (Halstead et

McKercher, 1975). Malgré l’abondance de ces formes organiques dans les sols, leur

dynamique et leur devenir dans l’écosystème sont beaucoup moins étudiés par rapport au P

inorganique.

L’adsorption des composés organiques phosphatés aux minéraux du sol dépend de

leur nature et des propriétés édaphiques (Condron et al., 2005). Les inositols phosphates et

les phosphonates sont fortement retenus dans le sol et constituent, par conséquence, la

forme stable du Po. Par contre, les orthophosphates monoesters simples (avec un seul

groupe de phosphate), tels que les glucides phosphate et les mononucléotides, les

Page 31: Caractérisation par spectroscopie et analyse compositionnelle des

9

orthophosphates diesters et les polyphosphates organiques sont moins retenus et peuvent

être dégradés rapidement par les enzymes de type phosphatase (Celi et al., 1999).

Cependant, l’acide désoxyribonucléique (ADN) peut persister à la biodégradation lorsqu’il

est lié à l’argile, au sable ou aux acides humiques (Khanna et al., 1998). La rétention de

l’ADN est inversement proportionnelle au pH du sol; elle augmente particulièrement à un

pH inférieur à 5, qui correspond à son point isoélectrique (Khanna et al., 1998; Condron et

al., 2005).

Tableau 2-1 Composés organiques phosphatés communs du sol (Turner et al., 2005).

Page 32: Caractérisation par spectroscopie et analyse compositionnelle des

10

2.3 Mesures du phosphore du sol

2.3.1 Méthodes conventionnelles

2.3.1.1 Phosphore total

La caractérisation du phosphore total nécessite la solubilisation de toutes les formes

inorganiques et organiques du P dans le sol. Deux anciennes méthodes ont été largement

appliquées; soient, la fusion avec le carbonate de sodium (NaCO3), et la digestion avec

l’acide perchlorique (HClO4) (Olsen et Sommers, 1982). Cependant, ces méthodes ne sont

plus utilisées étant donné que la première était fastidieuse et inappropriée pour un grand

nombre d’échantillons de sol, et la deuxième présentait un risque potentiel d’explosion dû à

la réaction de l’acide perchlorique avec les composés organiques (O’Halloran et Cade-

Menun, 2007). Actuellement, trois autres méthodes sont adoptées par de nombreux

laboratoires. La première consiste à l’oxydation de l’échantillon du sol avec l’hypobromite

de sodium (NaOBr) et l’hydroxyde de sodium (NaOH) (Dick et Tabatabai, 1977). Les deux

autres méthodes consistent à une digestion acide humide avec une solution d’acide

sulfurique (H2SO4), de peroxyde d’hydrogène (H2O2), et (1) d’acide fluoridrique (HF)

(Bowman, 1988); ou (2) de sulfate de lithium (Li2SO4) et de sélénium (Se) (Parkinson et

Allen, 1975). La détermination du PT est alors jusqu’à date relativement lente et implique

l’utilisation de plusieurs produits chimiques.

2.3.1.2 Phosphore disponible aux plantes

La forme du P inorganique disponible aux plantes peut être mesurée par de

nombreuses méthodes dont les plus utilisées pour les analyses routinières sont les méthodes

d’extractions chimiques et à l’eau (Tableau 2.2). Le type de l’extractif utilisé dépend

fortement de la nature du sol. Pour les sols acides, les méthodes recommandées sont celles

qui reposent sur des solutions acides telles que les méthodes de Bray 1 et Bray 2 qui

permettent d’extraire principalement le P lié à l’aluminium et au fer (Bray et Kurtz, 1945).

Pour les sols calcaires, les solutions alcalines sont les mieux appropriées comme la

méthode Olsen utilisant le bicarbonate de sodium (Olsen et al., 1954). La méthode de

Mehlich 3 (Mehlich, 1984), est bien adaptée pour une large gamme de sols acides, neutres

et légèrement calcaires. C’est la méthode de référence au Québec. L’extraction à l’eau est

Page 33: Caractérisation par spectroscopie et analyse compositionnelle des

11

aussi utilisée comme une méthode commune aux sols minéraux et organiques avec des

rapports sol/eau allant de 1/5 à 1/10 et 1/60 (Morel, 2002; Koopmans et al., 2001; Fardeau,

1996; Sissingh, 1971). Le P extrait par ces différentes méthodes est dosé par la suite soit

par colorimétrie, par spectrophotométrie d’émission au plasma ou par spectrométrie au

plasma à couplage inductif (Kuo, 1996).

Tableau 2-2 Méthodes d’extraction du P (tiré et adapté du Ziadi et al., 2013).

Méthode Extractif

pH de la

solution

extractive

Volume :

masse de

sol

pH du sol Durée

d’extraction Référence

Bray 1 0.5N HCl +

1N NH4F 3.0 7:1

<6.0; 6.0

à 7.2 1 mn

Bray et

Kurtz, 1945

Mehlich 3

0.015N NH4F

+ 0.025N

NH4NO3 +

0.2N

CH3COOH +

0.013N HNO3

+ 0.001N

EDTA

2.3 10:1 <6.0; 6.0

à 7.2 5 mn

Mehlich,

1984

Olsen 0.5M NaHCO3 8.5 20:1

<6.0; 6.0

à 7.2;

>7.2

30 mn Olsen et al.,

1954

Eau 7

10 :1

60 :1

<6.0; 6.0

à 7.2 24 h

Morel, 2002;

Sissingh,

1971

2.3.1.3 Phosphore organique

La détermination du Po total du sol est indirecte. Elle consiste à mesurer

l’augmentation du Pi dans l’extrait d’un échantillon de sol soumis à une combustion ou une

digestion par rapport à un échantillon témoin. La méthode d’ignition se fait par une

oxydation du Po en Pi à des faibles températures (250°C, Legg et Black, 1955) ou à des

températures élevées (550°C, Saunders et Williams, 1955), suivie d’une extraction acide du

PT. La digestion suit une série d’extractions acides et/ou basiques (Bowman et Moir, 1993;

1980; Stewart and Oades, 1972; Mehta et al., 1954) du Po, et permet de déterminer le PT.

Le Pi est mesuré dans l’extrait du sol en utilisant l’acide sulfurique (H2SO4; Anderson,

1960). La concentration du Po peut être surestimée par une solubilisation du Pi suite à la

Page 34: Caractérisation par spectroscopie et analyse compositionnelle des

12

combustion de l’échantillon du sol (Williams et al., 1970), ou sous-estimée suite à une

oxydation ou une extraction incomplète (O’Halloran et Cade-Menun, 2007). Par

conséquent, la méthode de la combustion est adoptée pour des traitements avec le même

type de sol, alors que les techniques d’extraction sont recommandées pour comparer les

niveaux du Po dans différents types de sol (Bowman, 1989).

2.3.1.4 Pools du P

Dans le sol, le P est réparti en différents pools selon leur disponibilité aux plantes

(Cross et Schlesinger, 1995). Ces pools peuvent être déterminés par une méthode

d’extraction séquentielle utilisant des solutions chimiques de plus en plus fortes (Pierzynski

et al., 2005). La méthode de Chang et Jakson (1957) fut la première développée dans ce

contexte selon la séquence suivante : NH4Cl, NH4F, NaOH, H2SO4 et NH4F ou NaOH,

pour extraire respectivement le Pi disponible, le Pi associé à l’aluminium, au fer, au

calcium et le P fortement retenu. La fraction du P résiduel est obtenue par différence entre

le P total déterminé par digestion et la somme de ces cinq fractions (Chang et Jakson, 1957;

Pierzynski et al., 2005). Des modifications ont été rapportées à cette méthode (Peterson et

Corey, 1966; Smith, 1965; Williams et al., 1967) dans le but de corriger son inefficacité

d’extraction du P lié au fer et de l’adapter aux sols calcaires ainsi qu’aux sédiments.

Afin de faire ressortir la contribution du phosphore organique à la dynamique du P

dans le sol, une méthode de fractionnement développée par Hedley et al. (1982; Fig. 2.3) a

été largement utilisée (Cross et Schlesinger, 1995; Zheng et al., 2003; Agbenin et Tiessen,

1994; Frossard et al., 1989). Le pool inorganique labile, correspondant au Pi adsorbé aux

surfaces des composés phosphatés plus cristallins : sesquioxydes ou carbonates (Mattingly,

1975), est extrait par la résine échangeuse d’anions et par le NaHCO3. Cependant, la

fraction du Pi la moins biodisponible associée aux oxyhydroxydes de fer et d’aluminium

amorphes et cristallins, est extraite avec le NaOH. Le Pi extractible au HCl est lié au

calcium (Tiessen et Moir, 2007). Le Po facilement minéralisable est extrait au NaHCO3 et

constitue le pool labile du Po, tandis que le Po extrait au NaOH est plus stable (Bowman et

Cole, 1978). Le P total résiduel de cette extraction séquentielle, déterminé par digestion à

l’aide de H2SO4 et H2O2, regroupe le Pi occlus aux minéraux de sol et le Po non extractible

(Tiessen et al., 1984). Toutefois, la fraction du Po active dans le cycle de transformation du

Page 35: Caractérisation par spectroscopie et analyse compositionnelle des

13

P dans le sol à court terme est sous-estimée par cette méthode, d’où la nécessité de

l’utilisation de l’extractif HCl concentré dans la nouvelle méthode de fractionnement de

Tiessen et Moir (2007) pour surmonter ce problème. Néanmoins, ces méthodes de

fractionnement sont opérationnelles et ne déterminent pas les formes spécifiques des pools

inorganique et organique du P (Condron et Newman, 2011).

Figure 2-3 Méthode d’extraction séquentielle du P du sol selon la méthode de Hedley

et al., 1982 (tirée de Cross et Schlesinger, 1995).

Page 36: Caractérisation par spectroscopie et analyse compositionnelle des

14

2.3.2 Méthodes spectroscopiques

2.3.2.1 Spectroscopie dans le proche infrarouge

La spectroscopie dans le proche infrarouge (SPIR) est une technique analytique

indirecte qui permet d’estimer la composition de la matière à partir de ses propriétés

d’absorption de la lumière (Stenberg et al., 2010). Cette technique consiste à soumettre

l’échantillon aux rayonnements dans la gamme du proche infrarouge allant de 800 à 2500

nm et à mesurer la reflectance (R) de la lumière par des détecteurs à chaque longueur

d’onde et la convertir par la suite en absorbance (log (1/R)).

Les rayonnements infrarouges sont absorbés par les liaisons chimiques entre les atomes

de la matière (C-H, O-H, N-H, C-O, S-H, CH2, et C-C), causant des mouvements de

torsion, de flexion ou d’étirement (Ludwig and Khanna, 2000). Les mesures de

l’absorbance sont utilisées pour calibrer la SPIR avec des mesures de référence obtenues au

laboratoire via différentes méthodes de régression telle que la régression linéaire multiple et

la régression de moindre carrée partielle. Le modèle de calibration obtenu est utilisé par la

suite pour prédire la composition de l’échantillon à partir de son spectre d’absorbance

(Martens et Naes, 2001). La performance de la prédiction de la SPIR est évaluée en

utilisant des paramètres statistiques tels que le coefficient de détermination et le rapport de

déviation de la performance (Nduwamungu et al., 2009a).

La SPIR est une technique rapide, peu coûteuse, non destructive et permet d’analyser

plusieurs propriétés de l’échantillon à partir du même spectre et de réduire les coûts

d’analyse de l’ordre d’au moins 50%. Elle trouve des applications dans plusieurs

domaines : la biologie, la chimie, la médecine et les industries pharmaceutique et

agroalimentaire. Son utilisation dans les sciences du sol est de plus en plus considérée pour

une évaluation plus rapide et plus précise de la qualité du sol (Guerrero et al., 2010). De

récentes études ont démontré que la SPIR est efficace pour, notamment, prédire la matière

organique, la texture, le carbone, l’azote, la capacité d’échange cationiqueet le pH du sol

(Cozzolino et Morón, 2006; Brunet et al., 2007; Nduwamungu et al., 2009a), de même que

les concentrations en Ca et Mg extraits au Mehlich-3 (Chang et al., 2001; Nduwamungu et

al., 2009b).

L’évaluation de l’efficacité du potentiel de la SPIR dans la prédiction de la teneur en P

des sols a fait aussi l’objet de plusieurs études dont les résultats trouvés divergent. À titre

Page 37: Caractérisation par spectroscopie et analyse compositionnelle des

15

d’illustration, Chang et al. (2001), Nduwamungu et al. (2009b), McCarty et Reeves (2006)

et Ludwig et al. (2002) ont montré que la SPIR ne pouvait prédire le P-Mehlich 3, le P-

Mehlich 1 et le P-Bray 2, respectivement. Cependant, Maleki et al. (2006) et van Groenigen

et al. (2003) ont trouvé des résultats de prédiction acceptables pour le P-Olsen. Cette

différence de résultats peut être expliquée par la nature des extractifs chimiques utilisés

(Guerrero et al., 2010). D’après Chang et al. (2001), le P disponible peut être prédit par la

SPIR s’il est relié aux propriétés primaires du sol telles que la matière organique et la

texture. La performance de la SPIR peut être aussi affectée par l’hétérogenéité de

l’échantillon du sol, de la texture du sol et par la variation dans les méthodes de préparation

de l’échantillon (Nduwamungu et al., 2009a).

2.3.2.2 Spectroscopie de résonance magnétique nucléaire du 31P

La spectroscopie de résonance magnétique nucléaire du 31P (RMN-31P) est une

technique qui utilise la résonance magnétique du noyau du P dans l’échantillon pour

identifier et quantifier sa forme chimique (Cade-Menun et al., 2005). En effet, le noyau du

P émet de l’énergie après avoir été soumis à des impulsions de radiofréquences dans un

champ magnétique; les quelles sont détectées sous forme de signal dans un spectre et

transformée par la suite en pics relatifs à chaque espèce ionique ou moléculaire du P (Cade-

Menun et al., 2005). L’avantage de cette technique est qu’elle permet de caractériser

simultanément les espèces du P, extraites au préalable du sol par une une solution alcaline

(NaOH-EDTA; He et al., 2008).

Newman et Tate (1980) furent les premiers à utiliser la RMN pour caractériser le P dans

les couches superficielles des sols de prairies en Nouvelle-Zélande. Beaucoup de travaux de

recherche l’ont par la suite utilisée pour caractériser le Po dans les sols agricoles (Hawkes

et al. 1984, Leinweber et al. 1997; Smernik et Dougherty, 2007; Redel al., 2011 et Cade-

Menun et Liu, 2014), les écosystèmes aquatiques (Carman et al., 2000; Paytan et al., 2003)

ou les amendements organiques (Jing et al., 1992; Crousse et al., 2002). L’ensemble de ces

études ont permis d’améliorer les connaissances sur l’origine, la distribution, la

biodisponibilité et la dynamique des formes chimiques du P dans l’environnement, ce qui

permet de mieux gérer le P dans les écosystèmes cultivés. A titre d’illustration, la figure 2.4

Page 38: Caractérisation par spectroscopie et analyse compositionnelle des

16

montre des espèces inorganiques et organiques du P caractérisées par la RMN-31P dans la

couche superficielle d’un sol cultivé non labouré à l’Île du Prince Edward, Canada (Cade-

Menun et al., 2010).

Figure 2-4 Spectre obtenu par la spectroscopie magnétique nucléaire du 31P montrant

les composés phosphatés détectés dans la couche superficielle (5-10 cm) d’un sol

cultivé non labouré (Cade-Menun et al., 2010).

2.4 Changements des formes du phosphore selon les pratiques culturales

2.4.1 Changements des pools du P

L’effet des pratiques culturales sur les changements dans les pools de P a fait l’objet de

plusieurs travaux de recherche dans le but d’améliorer la gestion du P dans les systèmes

agricoles. A titre d’exemple, il a été démontré que l’apport des fertilisants phosphatés au

sol durant une longue durée engendrait une augmentation dans les fractions labiles du Pi

(Pi-résine, Pi-NaHCO3, Pi-NaOH) significativement plus importante en système de culture

continue qu’en rotation (McKenzie et al,. 1992a; McKenzie et al., 1992b). Ce P accumulé

risque d’atteindre les cours d’eau adjacents constituant ainsi un problème environnemental

Page 39: Caractérisation par spectroscopie et analyse compositionnelle des

17

majeur (Leinweber et al., 1999). Des résultats similaires conformant la forte corrélation

entre l’apport du P et l’augmentation des teneurs des pools de P labiles ont été observés par

Tran et N’Dayegamie (1995), Dobermann et al. (2002) et Redel et al. (2007) dans des

Inceptisols, des Oxisols et des Ultisols, respectivement. Zheng et al. (2002) ont montré que

l’apport d’une source organique de P en sol cultivé durant huit ans favorisait aussi

l’accumulation de la fraction disponible du Pi provenant de la minéralisation du Po ajouté.

En revanche, la non fertilisation phosphatée menait à l’épuisement des réservoirs en P

labile des sols (Agbenin et Goladi, 1998; McKenzie et al., 1992a). D’autre part,

l’application des sous-produits industriels tels que les biosolides papetiers en absence de la

fertilisation phosphatée a favorisé la mobilité et/ou la minéralisation du Po-NaOH et la

transformation du pool récalcitrant au cours du temps en forme labile de P (Fan et al.,

2010).

Les rotations culturales affectent aussi les teneurs et la distribution des pools de P à

différents degrés selon les cultures. En effet, Redel et al. (2007) ont observé une

accumulation du P total et de P non labile dans le sol après une culture de blé,

contrairement à la culture d’avoine qui a réduit la fraction du P non labile et a augmenté le

P relativement labile. De leur côté, Zheng et al. (2001) ont montré qu’une monoculture

d’orge combinée aux fertilisants minéraux réduisait les formes labiles du Po en faveur des

formes labiles du Pi. D’autre part, ils ont démontré que la rotation orge-fourrages en un

gleysol labouré et recevant du lisier conduisait à une augmentation en P labile plus

importante qu’avec les fertilisants minéraux. Magid (1993) a observé une augmentation des

fractions inorganiques labiles du P (Pi-résine et Pi-NaHCO3) sous une végétation d’hêtres,

alors que les formes organiques labiles étaient les plus favorisées sous un système cultural

sous prairie.

2.4.2 Changements des espèces de P

Le statut du Po dans le sol est fortement influencé par la matière organique. En effet,

Vincent et al. (2010) ont montré qu’un apport de litière à un sol minéral d’une forêt tropical

durant trois ans à une dose de 6 kg de P/ha/année a augmenté la teneur en Po de 16% dont

31% est sous forme d’ADN. Cependant, l’enlèvement de la litière durant 3 ans a réduit de

4,2 ± 1,6 kg ha-1 la concentration du Po à la surface du sol (0- 2 cm) incluant 0,84 ± 0,32 kg

Page 40: Caractérisation par spectroscopie et analyse compositionnelle des

18

ha-1 des phosphates monoesters et 1,26 ± 0.48 kg ha-1 d’ADN. À plus long terme, He et al.

(2008) ont trouvé que l’apport de litière de volaille à un sol argilo-loameux n’avait pas

changé les teneurs des fractions hydrolysables du Po (NaOH, HCl) alors qu’il avait

augmenté celles des pools labiles et stables du Pi; ce qui indique que les formes du P

organiques provenant de la litière se transformaient en Pi dans le sol. De leur côté, Cade-

Menun et al. (2010) ont observé que le labour du sol modifie la distribution des espèces du

P par rapport à un sol non labouré en augmentant les concentrations en ions

orthophosphates et en phytate dans la couche de 5 à 10 cm. D’autre part, Condron et al.

(1985) rapportent que la fertilisation phosphatée d’un sol sablo-loameux favorise à long

terme l’accumulation des orthophosphates monoesters par rapport aux autres espèces

jusqu’au 99% du Po.

2.5 Concept d’analyse des données compositionnelles

Les données compositionnelles sont définies comme étant des données strictement

positives d’un ensemble clos, de somme constante égale à 1 ou à 100%, ou à une unité de

mesure (e.g., mg kg-1, kg m-3, etc.), qui transmettent des informations relatives (Aitchison,

1986). Cet ensemble nommé simplexe, est défini comme suit avec un degré D représentant

le nombre de ses composantes (X) :

SD = [(X1,…, XD), X1 > 0,…, XD > 0, X1 + … + XD =1]

Les données compositionnelles sont multivariées et dépendantes les unes des autres; ce

qui signifie que toute augmentation d’une composante est accompagnée d’une réduction

d’au moins une autre composante indiquant une corrélation fausse et négative. Par

conséquent, chaque composante ne peut pas être interprétée indépendamment des autres

(Tolosana-Delgado et van den Boogart, 2011). Les données compositionnelles sont

caractérisées aussi par une distribution non-normale, une redondance de l’information au

sein du simplexe, et une dépendance de leurs résultats d’analyse à l’échelle de mesure

(Aitchison, 1986). Ces biais intrinsèques peuvent générer des résultats erronés menant à des

interprétations contradictoires suite à l’application des analyses statistiques classiques

(Pawlowsky-Glahn et Egozcue, 2006; Filzmoser et al., 2009). Pour corriger ces problèmes,

Aitchison (1986) a proposé d’utiliser le log ratio additif (alr, Éq. 1) et le log ratio centré

Page 41: Caractérisation par spectroscopie et analyse compositionnelle des

19

(clr, Éq. 2) pour projeter les données compositionnelles de leur espace fermé à l’espace

réel, et de créer de nouvelles variables indépendantes de l’échelle de mesure pour que

l’interprétation soit cohérente :

Éq. 1

Éq. 2

Où est une composante quelconque, est une composante sélectionnée comme

dénominateur commun et ) est la moyenne géométrique de toutes les composantes.

Egozcue et al. (2003), ont développé plus tard une transformation du log ratio

isometrique (ilr, Éq. 3) qui permet de corriger la redandance en générant D-1 variables

orthogonales (90°) à partir de D composantes sans perdre de l’information. Cette

transformation permet d’analyser les composantes du simplexe en termes de D-1 balances

entre deux composantes ou deux groupes de composantes avec des signes + et – comme

suit :

Éq. 3

Où r est le nombre de composantes dans le groupe +, s est le nombre de composantes

dans le groupe –, est la moyenne géométrique des composantes dans le groupe et

est la moyenne géométrique des composantes dans le groupe .Les balances sont

conçues selon une base théorique spécifique à chaque système à l’étude (Egozcue et al.,

2003; Egozcue et Pawlowsky-Glahn, 2005). La variation du système d’un état initial x à un

état final y, peut être mesurée par le concept de la distance d’Aitchison (A, Éq. 4; Aitchison

et Egozcue 2005) suivant:

et Éq. 4

L’analyse des données compositionnelles a été largement utilisée en géochimie

(Buccianti et Pawlowsky-Glahn, 2005; Borgheresi et al., 2013; Engle et Rowan, 2013; Zuo

et al., 2013; Buccianti et Grunsky, 2014) et en sciences des aliments (Korhoňová et al.,

2009; Hron et al., 2012; Veverka et al., 2012). En agronomie, des études récentes ont été

faites sur le diagnostic nutritif des tissus végétaux (Parent et al., 2012a; Parent et al., 2013a;

Page 42: Caractérisation par spectroscopie et analyse compositionnelle des

20

Parent et al., 2013b), la mesure de l’agrégation dans les sols (Parent et al., 2012b), la

décomposition des résidus organiques (Parent et al., 2011), et la dynamique des pesticides

dans le sol (Aslam et al., 2013).

L’analyse compositionnelle est tout à fait adaptée aux formes du P (fractions ou espèces

chimiques) puisque ce sont des données fermées à 100% du P total. Dans une étude récente,

Parent et al. (2014) ont démontré que l’analyse classique des fractions de P est biaisée par

la dépendance de l’échelle de mesure, et ont élaboré une hiérarchie de balances entre ces

fractions pour évaluer le risque de perte du P dans les écosystèmes agricoles.

Page 43: Caractérisation par spectroscopie et analyse compositionnelle des

21

2.6 Hypothèses

De cette revue de littérature, il découle que de nouvelles méthodes pourraient être

utilisées pour mesurer les formes de P et mieux caractériser leurs changements dans les sols

agricoles. D’où, les hypothèses suivantes ont été formulées :

1. La spectroscopie dans le proche infrarouge peut déterminer la concentration du P

total et du P disponible extrait à la solution Mehlich-3 et à l’eau dans un sol sablo-

loameux.

2. La spectroscopie dans le proche infrarouge est une technique efficace pour prédire le

P organique du sol.

3. L’analyse compositionnelle des espèces de P évite le biais lié à la dépendance de

l’échelle de mesure, contrairement aux analyses statistiques classiques.

4. Le système du travail du sol et la fertilisation phosphatée changent les concentrations

et la distribution des espèces chimiques du P dans un gleysol sous rotation maïs-soya.

Page 44: Caractérisation par spectroscopie et analyse compositionnelle des

22

2.7 Objectifs

L’objectif général de cette thèse a été de mesurer et de caractériser les changements des

formes opérationnelles et fonctionnelles du P dans les sols selon différentes pratiques

culturales en utilisant des nouvelles techniques spectroscopiques et mathématiques. Les

objectifs spécifiques sont les suivants:

1. Evaluer le potentiel de la spectroscopie dans le proche infrarouge à prédire le P total

et le P disponible extrait à la solution Mehlich-3 et à l’eau dans un sol sableux

loameux.

2. Evaluer le potentiel de la spectroscopie dans le proche infrarouge à prédire le P

organique dans des sols loameux et argileux-loameux.

3. Démontrer que les analyses statistiques conventionnelles des formes de P sont

biaisées par la dépendance de l’échelle et que l’utilisation de l’analyse

compositionnelle permet d’éviter ce biais et d’avoir des résultats cohérents et fiables.

4. Etudier l’effet du système du travail du sol et de la fertilisation phosphatée sur les

teneurs et la distribution des espèces de P à l’aide de la résonance magnétique nucléaire

du 31P et l’analyse compositionnelle.

Page 45: Caractérisation par spectroscopie et analyse compositionnelle des

23

2.8 BIBLIOGRAPHIE

Aitchison, J. 1986. The statistical analysis of compositional data. Cox, D.R., Hinkley, D.V.,

Rubin D. and Silverman, B.W. Eds. 416 p.

Agbenin, J.O., and Goladi, J.T. 1998. Dynamics of phosphorus fractions in a savanna

Afisol under continuous cultivation. Soil Use Manag. 14: 59-64.

Agbenin, J.O., and Tiessen, H. 1994. Phosphorus transformations in a toposequence of

Lithosols and Cambisols from semi-arid northeastern Brazil. Geoderma, 62: 345-362.

Anderson, G. 1960. Factors affecting the estimation of phosphate esters in soil. J. Sci. Food

Agric. 11: 497–503.

Aslam, S., Garnier, P., Rumpel, C., Parent, S-É., and Benoit, P. 2013. Adsorption and

desorption behavior of selected pesticides as influenced by decomposition of maize

mulch. Chemosphere 91: 1447–1455.

Baldwin, D.S., Howitt, J.A., and Beattie, J.K. 2005. Abiotic degradation of organic

phosphorus compounds in the environment. In: B.L. Turner, E. Frossard, and D.S.

Baldwin, Eds, Organic phosphorus in the environment. CABI Publishing. Oxfordshire,

U.K. pp.75-88.

Bélanger, G. and Ziadi, N. 2008. Phosphorus and nitrogen relationships during spring

growth of an aging timothy sward. Agron. J. 100:1757-1762.

Borgheresi, M., Buccianti, A., Benedetto, F., and Vaughan, D., 2013. Application of

Compositional Techniques in the Field of Crystal Chemistry: A Case Study of

Luzonite, a Sn-Bearing Mineral. Math. Geosci. 45: 183-206.

Bowman, R.A. 1988. A rapid method to determine total phosphorus in soils. Soil Sci. Soc.

Am. J. 52: 1301–1304.

Bowman, R.A. 1989. A sequential extraction procedure with concentrated sulphuric acid

and dilute base for soil organic phosphorus. Soil Sci. Soc. Am. J. 53: 362–366.

Bowman, R.A. and Cole, C.V. 1978. An exploratory method for fractionation of organic

phosphorus from grassland soils. Soil Sci. 125: 95-101.

Bowman, R.A. and Moir, J.O. 1993. Basic EDTA as an extractant for soil organic

phosphorus. Soil Sci. Soc. Am. J. 57: 1513–1518.

Bray, R.H., and Kurtz, L.T. 1945. Determination of total, organic, and available forms of

phosphorus in soils. Soil Sci. 59: 39- 45.

Brunet, D., Barthès, B.G., Chotte, J.-L. and Feller, C. 2007. Determination of carbon and

nitrogen contents in Alfisols, Oxisols and Ultisols from Africa and Brazil using NIRS

analysis: Effects of sample grinding and set heterogeneity. Geoderma, 139:106–117.

Buccianti, A., and Grunsky, E. 2014. Compositional data analysis in geochemistry : Are we

sure to see what really occurs during natural processes ? J. Geochem. Expl. 141: 1-5.

Buccianti, A., and Pawlowsky-Glahn, V., 2005. New perspectives on water chemistry and

compositional data analysis. Math. Geol. 37: 703-727.

Cade-Menun, B.J. 2005. Using phosphorus-31 nuclear magnetic resonance spectroscopy to

characterize phosphorus in environmental samples. In: B.L. Turner et al., editors,

Organic phosphorus in the environment. CABI Publ., Wallingford, UK. p. 21–44.

Cade-Menun, B.J., Carter, M.R., James, D.C., and Liu, C.W. 2010. Phosphorus forms and

chemistry in the soil profile under long-term conservation tillage: A phosphorus-31

nuclear magnetic resonance study. J. Environ. Qual. 39: 1647-1656.

Page 46: Caractérisation par spectroscopie et analyse compositionnelle des

24

Cade-Menun, B.J., and Liu, C.W. 2014. Solution phosphorus-31 nuclear magnetic

resonance spectroscopy of soils from 2005 to 2013: A review of sample preparation

and experimental parameters. Soil Sci. Soc. Am. J. 78:19–37.

Carman, R., Edlund, G. and Damberg, C. 2000. Distribution of organic and inorganic

phosphorus compounds in marine and lacustrine sediments: a 31P NMR study. Chem.

Geol. 163: 101–114.

Cathcart, J.B. 1980. World phosphate reserves and resources. In Khasawneh et al. (ed.).

The rôle of phosphorus in agriculture. American Society of Agronomy, Madison,

Wisconsin, USA, pp. 1–18.

Chang, C.W., Laird, D.A., Mausbach, M.J., and Hurburgh, C.R.Jr. 2001. Near-infrared

reflectance spectroscopy – Principal components regression analyses of soil properties.

Soil Sci. Soc. Am. J. 65:480–490.

Chang, S.C. and Jakson, M.L. 1957. Fractionation of soil phosphorus. Soil Sci. 84: 133-

144.

Chen, C.R., Condron, L.M., Davis M.R., and Sherlock, R.R.. 2002. Phosphorus dynamics

in the rhizosphere of perennial ryegrass (Lolium perenne L.) and radiata pine (Pinus

radiata D. Don). Soil Biol. Biochem. 34:487–499.

Chen, C.R., Condron, L.M., Davis, M.R., and Sherlock, R.R. 2003. Seasonal changes in

soil phosphorus and associated microbial properties under adjacent grassland and forest

in New Zealand. F. Ecol. Manag. 177: 539-557.

Celi, L., Lamacchia, S., Marsan, T.A. and Barberis, E.1999. Interaction of inositol

phosphate on clays: Adsorption and charging phenomena. Soil Sci. 164:574–585.

Condron, L.M., Goh, L.M., and Newman, R.H. 1985. Nature and distribution of soil

phosphorus as revealed by a sequential extraction method followed by 31P nuclear

magnetic resonance analysis. J. Soil Sci. 36: 199-207.

Condron, L.M., and Newman, S. 2011. Revisiting the fundamentals of phosphorus

fractionation of sediments and soils. J. Soils Sed. 11:830–840.

Condron, L.M., Turner, B.L., and Cade-Menun, B.J. 2005. Chemistry and dynamics of soil

organic phosphorus. Phosphorus: agriculture and the environment. Sims, J.T. and

Sharpley, A.N. Eds. pp: 87-121.

Cordell, D., Drangert, J. O. and White. S. 2009. The story of phosphorus: Global food

security and food for thought. G. Environ. Change, 19: 292–305.

Cozzolino, D., and Morón, A. 2006. Potential of near-infrared reflectance spectroscopy and

chemometrics to predict soil organic carbon fractions. Soil Tillage Res. 85:78–85.

Cross, A.F, and Sclesinger, W.H. 1995. A literature review and evaluation of the Hedley

fractionation: Applications to the biogeochemical cycle of soil phosphorus in natural

ecosystems. Geoderma, 64: 197- 214.

Crouse, D.A., Sierzputowska-Gracz, H., Mikkelson, R.L. and Wollum, A.G. 2002

Monitoring phosphorus mineralization from poultry manure using phosphatase assays

and phosphorus-31 nuclear magnetic resonance spectroscopy. Comm. in Soil Sci. Plant

Anal. 33: 1205–1217.

Dick, W.A. and Tabatai, M.A. 1977. An alkaline oxidation method for determination of

total phosphorus in soils. Soil Sci. Soc. Am. J. 41: 511–514.

Dobermann, A., George, T., and Thevs, N. 2002. Phosphorus fertilizer effects on soil

phosphorus pools in acid upland soils. Soil Sci. Soc. Am. J., 66: 652-660.

Egozcue, J. J., and Pawlowsky-Glahn, V. 2005. Groups of Parts and Their Balances in

Compositional Data Analysis. Math. Geol. 37: 795-828.

Page 47: Caractérisation par spectroscopie et analyse compositionnelle des

25

Egozcue, J. J., Pawlowsky-Glahn, V., Mateu-Figueras, G., and Barcel´o-Vidal, G. 2003.

Isometric logratio transformations for compositional data analysis. Math. Geol. 35: 279-

300.

Engle, M.A., and Rowan, E.L., 2013. Interpretation of Na-Cl-Br Systematics in

Sedimentary Basin Brines: Comparison of Concentration, Element Ratio, and Isometric

Log-ratio Approaches. Math. Geosci. 45: 87-101.

Fan, J.L., Ziadi, N., Gagnon, B., and Hu, Z.Y. 2010. Soil phosphorus fractions following

annual paper mill biosolids and liming materials application. Can. J. Soil Sci. 90: 467-

479.

Fardeau, J. C. 1996. Dynamics of phosphate in soils. An isotopic outlook. Fert. Res. 45: 91-

100.

Filzmoser, P., Hron, K., and Reimann, C. 2009. Univariate statistical analysis of

environmental (compositional) data: Problems and possibilities. Sci. Total Environ.

407: 6100-6108.

Firsching, B.M., and N Claassen. 1996. Root phosphatase activity and soil organic

phosphorus utilization by Norway spruce (Picea abies (L) Karst). Soil Biol. Biochem.

28: 1417–1424.

Frossard, E., Condron, L.M., Oberson, A., Sinaj, S., and Fradeau, J.C. 2000. Processes

governing phosphorus availability in temperate soils. J. Environ. Qual. 29: 15-23.

Frossard, E., Stewart, J.W.B., and Arnaud, R.J.ST. 1989. Distribution and mobility of

phosphorus in grassland and forest soils of Saskatchewan. Can. J. Soil Sci. 69: 401-

416.

Guerrero, C., Rossel, R.A.V., and Mouazen, A.M. 2010. Special issue ‘Diffuse reflectance

spectroscopy in soil science and land resource assessment’. Geoderma, 158: 1-2.

Halstead, R.L., and McKercher, R.B. 1975. Biochemistry and cycling of phosphorus. Soil

Biol. Biochem. 4: 31- 63.

Harrison, A.F. 1982. Labile organic phosphorus mineralization in relationship to soil

properties. Soil Biol. Biochem. 14: 343-351.

Harrison, A.F. 1987. Soil organic phosphorus-A review of world literature. CAB Int.,

Wallingford, Oxon, UK. 257 p.

Haygarth, P.M., L. Hepworth, and S.C. Jarvis. 1998. Forms of phosphorus transfer in

hydrological pathways from soil under grazed grassland. Eur. J. Soil Sci. 49:65-72.

Hawkes, G.E., Powlson, D.S., Randall, E.W., and Tate, K.R. 1984. A 31P nuclear magnetic

resonance study of the phosphorus species in alkali extracts of soils from long-term

field experiments. J. Soil Sci. 35: 35-45.

He, Z., Honeycutt, C. W., Cade-Menun, B.J. Senwo, Z. N., and Trazisong, I. A. 2008.

Phosphorus in poultry and soil: enzymatic and nuclear magnetic resonance

characterization. Soil Sci. Soc. Am. J. 72: 1426-1433.

Hedley, M.J., Stewart, J.W.B., and Chauhan, B.S. 1982. Changes in inorganic and organic

soil phosphorus fractions induced by cultivation practices and by laboratory

incubations. Soil Sci. Soc. Am. J. 46: 970-976.

Hron, K., Jelínková, M., Filzmoser, P., Kreuziger, R., Bednář, P., and Barták, P., 2012.

Statistical analysis of wines using a robust compositional biplot. Talanta, 90: 46-50.

Hinsinger, P. 2001. Bioavailability of soil inorganic P in the rhizosphere as affected by

root-induced chemical changes: a review. Plant Soil. 237: 173-195.

Hinsinger, P., Jaillard, B., Le Cadre. É., and Plassard, C. 2007. Spéciation et biodisponiblité

du phosphore dans la rhizosphère. Océanis, 33: 37- 50.

Page 48: Caractérisation par spectroscopie et analyse compositionnelle des

26

Holford, I.C.R. 1997. Soil phosphorus: its measurement, and its uptake by plants. Austr. J.

Soil Res. 35: 227-239.

Holtan, H., Kamp-Nielson, L., and Stuanes, A.O. 1988. Phosphorus in soil, water and

sediment: an overview. Hydrobiologia, 170: 19-34.

Jing, S.R., Benefield, L.D. and Hill, W.E. 1992. Observations relating to enhanced

phosphorus removal in biological systems. Water Res. 26: 213–223.

Khiari, L., and Parent, L.É. 2005. Phosphorus transformations in acid light-textured soils

treated with dry swine manure. Can. J. Soil Sci. 85: 75- 87.

Khanna, M., Yoder, M., Calamai, L. and Stotzky, G. 1998. X-ray diffractometry and

electron microscopy of DNA from Bacillus subtilis bound on clay minerals. Sci. Soils.

Sedim. 3: 1–8.

Kleinman, P. J. A, Sharpley, A.N., Saporito, L.S., Buda, A.R., and Bryant, R.B. 2009.

Application of manure to no-till soils: phosphorus losses by sub-surface and surface

pathways. Nutr. Cycl. Agroeco. 84: 215–227.

Koopmans, G. F.; van Der Zeeuw, M. E., Chardon,W. J., and Dolfing, J. 2001. Selective

extraction of labile phosphorus using dialysis membrane tubes filled with hydrous

iron hydroxide. Soil Sci. 166: 475-483.

Korhoňová, M., Hron, K., Klimčíková, D., Müller, L., Bednář, P., and Barták, P., 2009.

Coffee aroma—Statistical analysis of compositional data. Talanta 80: 710-715.

Kuo, S. 1996. Phosphorus. In methods of soil analysis. Part 3-Chemical methods. Ed. D.L.

Sparks. 1390 p.

Legg, J.O. and Black, C.A. 1955. Determination of organic phosphorus in soils. II. Ignition

method. Soil Sci. Soc. Am. Proc. 19: 139–143.

Leinweber, P., Haumaier, L., and Zech, W. 1997. Sequential extractions and 31P-NMR

spectroscopy of phosphorus forms in animal manures, whole soils and particle-size

separatesfrom a densely populated livestock area in northwest Germany. Biol. Fert.

Soils. 25: 89-94.

Leinweber, P., Meissner, R., Eckhardt, K.U., and Seeger, J. 1999. Management effects on

forms of phosphorus in soil and leaching losses. Eur. J. Soil Sci. 50: 413-424.

Liu, Y. and Chen, J. 2008. Phosphorus cycle. Global Ecol. 2715- 2724.

Ludwig, B., Khanna, P.K., Bauhus, J., and Hopmans, P. 2002. Near infrared spectroscopy

of forest soils to determine chemical and biological properties related to soil

sustainability. Forest. Ecol. Manag. 171: 121- 132.

Magid, J. 1993. Vegetation effects on phosphorus fractions in set-aside soils. Plant Soil.

149: 111-119.

Maleki, M.R., Van Holm, L., Ramon, H., Merckx, R., De Baerdemaeker, J., and Mouazen,

A.M. 2006. Phosphorus Sensing for Fresh Soils using Visible and Near Infrared

Spectroscopy. Biosys. Engin. 95: 425- 436.

Mattingly, G.E.G. 1975. Labile phosphate in soils. Soil Sci. 119: 369- 375.

Martens, H., and T. Naes. 2001. Multivariate calibration by data compression. In: P.C.

Williams and K.H. Norris, editors, Near infrared technology in the agricultural and

food industries. 2nd ed. Am. Assoc. of Cereal Chemists, St. Paul, MN. pp. 59–100.

McCarty, G.W., and Reeves, J.B. 2006. Comparaison of Near infrared and mid infrared

diffuse reflectance spectroscopy for field scale measurement of soil fertility

parameters. Soil Sci. 171: 94-112.

McGechan, M.B., and Lewis, D.R. 2002. Sorption of phosphorus by soil, Part 1: Principles,

equations and models. Biosyst. Eng. 82: 1-24.

Page 49: Caractérisation par spectroscopie et analyse compositionnelle des

27

McKelvie, I.D. 2005. Separation, preconcentration and speciation of organic phosphorus in

environmental samples. Turner, B.L., Frossard, E., and Baldwin, D.S. Eds. 399 p.

McKenzie, R.H., Stewart, J.W.B., Dormaar, J.F., and Schaalje, G.B. 1992a. Long-term

crop rotation and fertilizer effects on phosphorus transformations: I. In a chernozemic

soil. Can. J. Soil Sci. 72: 569-579.

McKenzie, R.H., Stewart, J.W.B., Dormaar, J.F., and Schaalje, G.B. 1992b. Long-term

crop rotation and fertilizer effects on phosphorus transformations: II. In a luvisolic soil.

Can. J. Soil Sci. 72: 581-589.

McLaughlin, M.J., Alston, A.M., and Martin, J.K. 1988. Phosphorus cycling in water-

pasture rotations. II.*The role of the microbial biomass in phosphorus cycling. Aust. J.

Soil Res. 26: 333-342.

Mehlich, A. 1984. A modification of Mehlich 2 extractant. Commun. Soil Sci. Plant.Anal.

15: 1409–1416.

Mehta, N.C., Legg, J.O., Goring, C.A.I., and Black, C.A. 1954. Determination of organic

phosphorus in soils. I. Extraction method. Soil Sci. Soc. Am. Proc. 18: 443–449.

Messiga, A.J., Ziadi, Z., Morel, C., and Parent, L.É. 2010. Soil phosphorus availability in

no-till versus conventional tillage following freezing and thawing cycles. Can. J. Soil

Sci. 90: 419-428.

Morel, R. 1989. Les sols cultivés. Technique et Documentation-Lavoisier, Paris, France.

373 p.

Morel, C. 2002. Caractérisation de la phytodisponibilité du phosphore du sol par la

modélisation du transfert des ions phosphates entre le sol et la solution. Mémoire

préparé en vue de l’obtention du diplôme d’habilitation à diriger des recherches. Institut

National Polytechnique de Lorraine. INRA, Bordeaux, 80 p.

Motavalli, P.P., and Miles, R.J. 2002. Soil phosphorus fractions after 111 years of animal

manure and fertilizers applications. Bio. Fert. Soils. 36: 35-42.

Nduwamungu, C., Ziadi, N., Tremblay, G.F., and Parent, L.-É.. 2009a. Near-infrared

reflectance spectroscopy prediction of soil properties: Effects of sample cups and

preparation. Soil Sci. Soc. Am. J. 73: 1896–1903.

Nduwamungu, C., Ziadi, N., Parent, L.-É., and Tremblay, G.F.. 2009b. Mehlich 3

extractable nutrients as determined by near-infrared reflectance spectroscopy. Can. J.

Soil Sci. 89: 579–587.

Negassa, W., and Leinweber, P. 2009. How does the Hedley sequential phosphorus

fractionation reflect impacts of land use and management on soil phosphorus: A

review. J. Plant Nutr. Soil Sci. 172: 305–325.

Newman, R.H., and Tate, K.R. 1980. Soil phosphorus characterisation by 31P nuclear

magnetic resonance. Commun. Soil Sci. Plant Anal. 11: 835-842.

O’Halloran, I.P. and Cade-Menun, B.J. 2007. Total and organic phosphorus. In Soil

Sampling and Methods of analysis. Carter, M.R., and Gregorich, E.G., (2nd eds). pp.

265–291.

Oberson, A., and Joner, E.J. 2005. Microbial turnover of phosphorus in soil. Organic

phosphorus in the environment. Turner, B.L., Frossard, E., and Baldwin, D.S. Eds.

399 p.

Oehl, F., Oberson, A., Sinaj, S., and Frossard, E. 2001. Organic phosphorus mineralization

studies using isotopic dilution techniques. Soil Sci. Soc. Am. J. 65: 780–787.

Page 50: Caractérisation par spectroscopie et analyse compositionnelle des

28

Olsen, S.R., Cole, C.V., Watanabe, F.S. and Dean, L.A. 1954. Estimation of available

phosphorus by extraction with sodium bicarbonate. USDA Circ. 939, U.S. Govt.

Print. Office, Washington, D.C.

Olsen, S.R. and Sommers, L.E. 1982. Phosphorus. In A.L. Page, R.H., Miller and D.R.,

Keeney, Eds. Methods of Soil analysis. Part 2. (2nd ed.) Agronomy No. 9. American

Society of Agronomy, Madison, WI, USA. pp. 403–430.

Parent, L.É., Parent, S.É., Hébert-Gentile, V., Naess, K., and Lapointe, L. 2013a. Mineral

Balance Plasticity of Cloudberry (Rubus chamaemorus) in Quebec-Labrador Bogs.

Am. J. Plant Sci. 4: 1508-1520.

Parent, L.É., Parent, S.É., and Ziadi, N. 2014. Biogeochemistry of soil inorganic and

organic phosphorus: A compositional analysis with balances. J. Geochem. Expl. 141:

52–60.

Parent, L.É, Almeida, C.X., Hernandes, A., Egozcue, J.J., Gülser, C, Bolinder, M.A.,

Kätterer, T., Andrén, O., Parent, S.É., Anctil, F., Centurion, J.F., and Natale, W. 2012b.

Compositional analysis for an unbiased measure of soil aggregation. Geoderma 179-

180: 123–131.

Parent S.É., Karam A., and Parent L.E. 2011. Compositional modeling of C mineralization

of organic materials in soils. 2nd Int. Symp. Agricultural Waste Management, 15-17

March 2011 (Sigera Proceedings paper), Foz do Iguaçu, Brazil.

Parent, S.É., Parent, L.E., Rozanne, D.E., Hernandes, A. and Natale, W. 2012a. Nutrient

Balance as Paradigm of Soil and Plant Chemometrics. INTECH.

http://dx.doi.org/10.5772/53343.

Parent, S.É, Parent, L.É., Rozane, D.E. and Natale, W. 2013b. Plant ionome diagnosis

using sound balances: case study with mango (Mangifera Indica). Front. Plant Sci. 4: 1-

12.

Parkinson, J.A. and Allen, S.E. 1975. A wet oxidation procedure suitable for the

determination of nitrogen mineral nutrients in biological material. Comm. Soil Sci.

Plan. 6: 1–11.

Paytan, A., Cade-Menun, B.J., McLaughlin, K. and Faul, K.L. 2003. Selective phosphorus

regeneration of sinking marine particles: evidence from 31P NMR. Marine Chem. 82:

55–70.

Pawlowsky-Glahn, V., and Egozcue, J.J. 2006. Compositional data and their analysis: an

introduction. Compositional data analysis in the geosciences: from theory to pratice.

Buccianti, Mateu-Figueras, G., and Pawlowsky-Glahn, V. Eds. pp: 1-10.

Petersen, G.W. and Corey, R.B. 1966. A modified Chang and Jackson Procedure for

Routine Fractionation of Inorganic Soil Phosphates. Soil Sci. Soc. Am. J. 30: 563-565.

Pierzynski, G.M., McDowell, and Sims, J.T. 2005. Chemistry, cycling and potential

movement of inorganic phosphorus in soils. Phosphorus: agriculture and the

environment. Sims, J.T. and Sharpley, A.N. Eds. 53-86.

Quiquampoix, H. and Mousain, D. 2005. Enzymatic Hydrolysis of Organic Phosphorus. In

Phosphorus: agriculture and the environment. Sims, J.T. and Sharpley, A.N. Eds. pp:

89-112.

Redel, Y.D., Escudey, M., Alvear, M., Conrad, J. and Borie, F. 2011. Effects of tillage and

crop rotation on chemical phosphorus forms and some related biological activities in a

Chilean Ultisol. Soil Use Manag. 27: 221–228.

Page 51: Caractérisation par spectroscopie et analyse compositionnelle des

29

Redel, Y.D., Rubio, R., Rouanet, J.L., and Borie, F. 2007. Phosphorus bioavailability

affected by tillage and crop rotation on a chilean volcanic derived ultisol. Geoderma,

139: 388-396.

Richardson, A.E. 2001. Prospects for using soil microorganisms to improve the acquisition

of phosphorus by plants. Austr. J. Plant Physiol.. 28: 897- 906.

Ross, D.J., Tate, K.R., Scott, N.A., and Feltham, C.W. 1999. Land-use change: Effects on

soil carbon, nitrogen and phosphorus pools and fluxes in three adjacent ecosystems.

Soil Biol. Biochem. 31: 803-813.

Saunders, W.M.H. and Williams, E.G. 1955. Observations on the determination of total

organic phosphorus in soils. J. Soil Sci. 6: 254–267.

Sharpley, A.N., and Smith, S.J., 1994. Wheat tillage and water quality in the Southern

plains. Soil Tillage Res. 30: 33–38.

Shi, Y., Ziadi, N., Messiga, J.A., Lalande, R., and Hu, Z. 2013. Changes in soil phosphorus

fractions for a long-term corn-soybean rotation with tillage and phosphorus

fertilization. Soil Sci. Soc. Am. J. 77: 1402–1412.

Simard, R.R., Beauchemin, S., and Haygarth, P.M. 2000. Potentiel for preferential

pathways of phosphorus transport. J. Environ. Qual. 29: 97-105.

Sims, J.T., and Pierzynski, G.M. 2005. Chemistry of phosphorus in soil. In: Tabatabai A.M,

Sparks D.L. (eds) Chemical processes in soil, SSSA book series 8. SSSA, Madison, pp.

151-192.

Sissingh, H.A. 1971. Analytical technique of the Pw method used for the assessment of the

phosphate status of arable soils in the Netherlands. Plant Soil 34: 483-486.

Smernik, R.J., and Dougherty, W.J. 2007. Identification of phytate in phosphorus-31

nuclear magnetic resonance spectra: the need for spiking. Soil Sci. Soc. Am. J. 71:

1045- 1050.

Smith, A.N. 1965. Distinction between iron and aluminium phosphate in Chang and

Jackson’s procedure for fractionating inorganic soil phosphorus. Agrochimica, 9: 162-

168.

Stenberg, B., Rossel, R.A.V., Mouazen, A.M., and Wetterlind, J. 2010. Visible and near

infrared spectroscopy in soil science. Adv. Agron. 107: 163- 215.

Stevenson, F.J. 1986. Cycles of soil: carbon, nitrogen, phosphorus, sulfur, micronutrients.

380 p.

Steward, J.H. and Oades, J.M. 1972. The determination of organic phosphorus in soils. J.

Soil Sci. 23: 38–49.

Stewart, J.W.B., and Tiessen, H. 1987. Dynamics of soil organic phosphorus. Biogeochem.

4: 41-60.

Tiessen, H., and Moir, J.O. 2007. Characterization of available P by sequential extraction.

Soil sampling and methods of analysis. Carter, M.R., and Grecorich, E.G. Eds. pp. 75-

86.

Tiessen, H., Stewart, J.W.B., and Bettany, J.R. 1982. Cultivation effect on the amounts and

concentrations of carbon, nitrogen and phosphorus in grassland soils. Agron. J. 74:

831–835.

Tiessen, H., Stewart, J.W.B., and Cole, C.V. 1984. Pathways of phosphorus

transformations in soils of differing pedogenesis. Journal paper no. R361 of the

Saskatchewan Institute Of pedology, University of Saskatchewan. 8p.

Page 52: Caractérisation par spectroscopie et analyse compositionnelle des

30

Tolosana-Delgado, R. and K.-G. van den Boogart. 2011. Linear models with compositions

in R. pp. 356-371. In V. Pawlowsky-Glahn, and A. Buccianti, editors. Compositional

data analysis: Theory and Applications. John Wiley and Sons, New York.

Turner, B.L., Cade-Menun, B.J., Condron, L.M. and Newman, S. 2005. Extraction of soil

organic phosphorus. Talanta, 66: 294-306.

Turner, B.L., Mahieu, N. and Condron, L.M. 2003. Phosphorus composition of temperate

pasture soils determined by NaOH–EDTA extraction and solution 31P NMR

spectroscopy. Org. Geochem. 34: 1199–1210.

Tran, T.S., and N’Dayegamiye, A. 1995. Long term effects of fertilizers and manure

applications on the forms and availability of soil phosphorus. Can. J.Soil Sci. 75: 281-

285.

U.S. Geological Survey. 2014. Mineral commodities summaries. (Online). Available at

http://minerals.usgs.gov/minerals/pubs/commodity/phosphate_rock/ (verified 15

September 2014).

Van Groenigen, J.W., Mutters, C.S., Horwath, W.R. and van Kessel, C. 2003. NIR and

DRIFT-MIR spectrometry of soils for predicting soil and crop parameters in a flooded

field. Plant Soil. 250: 155- 165.

Veverka, L., Jelinkova, M., Hron, K., Balik, J., Stávek, J., and Barták, P., 2012. Chemical

markers in the aroma profiles of south Moravian red wine distillates. Czech J. Food Sci.

30: 369-376.

Vincent, A . G ., Turnerb, B. L., and Tannera, E. V. J. 2010. Soil organic phosphorus

dynamics following perturbation of litter cycling in a tropical moist forest. Eur. J. Soil

Sci. 61: 48–57.

Williams, J. D. H., Syers, J. K., and Walker, T. W. 1967. Fractionation of Soil Inorganic

Phosphate by a Modification of Chang and Jackson's Procedure. Soil Sci.Soc. Am. J.

31: 736-739.

Williams, J.D.H., Syers, J.K., Walker, T.W., and Rex, R.W. 1970. A comparison of

methods for the determination of soil organic phosphorus. Soil Sci. 110: 13–18.

Zhang, T.Q., and Mackenzie, A.F. 1997. Changes of soil phosphorous fractions under long-

term corn monoculture. Soil Sci. Soc. Am. J. 61: 485-493.

Zheng, Z., Simard, R.R., Lafond, J., and Parent, L.E. 2001. Changes in phosphorus

fractions of a humic gleysol as influenced by cropping systems and nutrient sources.

Can. J. Soil Sci. 81: 175-183.

Zheng, Z., Simard, R.R., Lafond, J., and Parent, L.E. 2002. Pathways of soil phosphorus

transformations after 8 years of cultivation under contrasting cropping practices. Soil

Sci. Soc. Am. J. 66: 999-1007.

Zheng, Z., Simard, R.R., and Parent, L-É. 2003. Anion exchange and Mehlich-III

phosphorus in Humaquepts varying in clay contents. Soil Sci. Soc. Am.J. 67: 1287-

1295.

Zuo, R., Xia, Q., and Wang, H., 2013. Compositional data analysis in the study of

integrated geochemical anomalies associated with mineralization. App. Geochem. 28:

202-211.

Ziadi, N., Whalen, J.K., Messiga, A.J., and Morel, C. 2013. Assessment and modeling of

soil svailable phosphorus in sustainable cropping systems. Adv. Agron. 122: 85-126.

Page 53: Caractérisation par spectroscopie et analyse compositionnelle des

31

CHAPITRE III: PREDICTING SOIL PHOSPHORUS-RELATED

PROPERTIES USING NEAR-INRARED REFLECTANCE

SPECTROSCOPY

Dalel Abdi1,2, Gaëtan F. Tremblay1, Noura Ziadi1, Gilles Bélanger1, and Léon-Étienne

Parent2

1Agriculture and Agri-Food Canada, Soils and Crops Research and Development Centre,

2560 Hochelaga Boulevard, Québec, QC, Canada G1V 2J3.

2Department of Soils and Agri-Food Engineering, Université Laval, Québec, QC, Canada

G1K 7P4.

Soil Science Society of America Journal, 2012. 76 (6): 2318–2326.

Page 54: Caractérisation par spectroscopie et analyse compositionnelle des

32

3.1 RÉSUMÉ

La spectroscopie dans le proche infra-rouge (SPIR) est une technique d’analyse

rapide, précise et peu coûteuse. Les objectives de cette étude étaient d’évaluer le potentiel

de la SPIR dans la prédiction (i) du P du sol extrait avec deux méthodes [Mehlich-3 (PM3)

et à l’eau (Cp)], P total (TP), P prelevé par les plantes, et bilan annuel de P, et (ii) d’autres

propriétés chimiques du sol [C total (TC), N total (TN), pH, K, Al, Fe, Ca, Mg, Cu et Zn

extraits au Mehlich-3]. Des échantillons du sol (n = 448) ont été prelevés d’un site

expérimental situé à Lévis, Québec. Les modèles de prédiction de la SPIR ont été

developpés en utilisant 80% des échantillons pour la calibration et 20% pour la validation.

Les résulats ont démontré que le PM3, Cp, P prelevé, bilan annuel de P, K et Cu n’ont pas

été prédictibles par la SPIR. Cependant, des prédictions fiables ont été trouvées pour TP,

TC, TN, Al, Fe, Zn, Mg, Ca, Mn, et pH.

Page 55: Caractérisation par spectroscopie et analyse compositionnelle des

33

3.2 ABSTRACT

Near infrared reflectance spectroscopy (NIRS) is a rapid, inexpensive, and accurate

analysis technique for a wide variety of materials and it is increasingly used in soil science.

The objectives of our study were to examine the potential of NIRS to predict: (i) soil P

extracted by two methods [Mehlich 3 (M3P) and water (Cp)], soil total P (TP), annual crop

P-uptake, and annual P-budget, and (ii) other soil chemical properties [total C (TC), total N

(TN), pH, and K, Al, Fe, Ca, Mg, Mn, Cu, and Zn extracted by Mehlich 3]. Soil samples (n

= 448) were taken over a 7-yr period from an experimental site in Lévis (Québec, Canada)

where timothy (Phleum pratense L.) was grown under four combinations of P and N

fertilizer. The NIRS equations were developed using 80% of the samples for calibration

and 20% for validation. The predictive ability of NIRS was evaluated using the coefficient

of determination of validation (Rv2) and the ratio of standard error of prediction to standard

deviation (RPD). Results show that M3P, Cp, crop annual P-uptake, and annual P-budget

were not accurately predicted by NIRS (Rv2 < 0.70 and RPD < 1.75). Similar results were

found for K and Cu. However, NIRS predictions were moderately useful for TP, TN, Fe,

and Zn (0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25), moderately successful for TC and Al

(0.80 ≤ Rv2 < 0.90 and 2.25 ≤ RPD < 3.00), successful for pH and Mg (0.90 ≤ Rv

2 ≤ 0.95

and 3.00 ≤ RPD ≤ 4.00), and excellent for Ca and Mn (Rv2 > 0.95 and RPD > 4.00). The

NIRS predictive ability of several soil properties appears to be related to their relationship

with soil organic C. Although NIRS can predict several soil properties, prediction of total P

was the only soil P-related property, correlated to soil C, that was moderately useful.

Abbreviations: b, slope of linear regression; Cp, P extracted in water; CV, coefficient of

variation; DM, dry matter; ICP, inductively coupled plasma; M3, Mehlich 3; M3P_Col, soil

P content extracted using the Mehlich 3 method and analysed by colorimetry; M3P_ICP,

soil P content extracted using the Mehlich 3 method and analysed by ICP; N, total number

of samples; NIRS, near infrared reflectance spectroscopy; PLSR, partial least squares

regression method; Rc2, coefficient of determination of calibration; Rv

2, coefficient of

determination of validation; Rep File, repeatability file; RPD, ratio of standard error of

prediction to standard deviation; SD, standard deviation; SEC, standard error of calibration;

SECV, standard error of cross-validation; SEP, standard error of prediction; SNVD,

standard normal variate and detrending; TC, total carbon; TN, total nitrogen; TP, total

phosphorus; 1-VR, coefficient of determination of cross-validation.

Page 56: Caractérisation par spectroscopie et analyse compositionnelle des

34

3.3 INTRODUCTION

Phosphorus is an essential nutrient and one of the most limiting for crop production.

Mineral and organic P fertilizers are often applied to agricultural soils to achieve optimal

crop yield but amounts exceeding crop requirements can have a negative environmental

impact.

Several methods and/or techniques of soil analysis, including chemical extraction

methods, have been developed to estimate the quantity of plant-available P in soils. Current

soil P extraction methods, such as Mehlich 3 (Mehlich, 1984), Olsen (Olsen et al., 1954),

Bray 1, Bray 2 (Bray and Kurtz, 1945), and water (Morel et al., 2000), are expensive,

destructive, and both time and space consuming. The recommended Mehlich 3 method for

a large range of soil types (Ziadi and Tran, 2007) requires five chemical reagents (acetic

acid, ammonium fluoride, ammonium nitrate, nitric acid, and ethylenediaminetetraacetic

acid).

Near infrared reflectance spectroscopy (NIRS) is a cost-effective, time-saving, non-

destructive, and environmentally-sound technique that can predict many constituents from

the single spectrum of a soil sample (Coûteaux et al., 2003; Viscarra Rossel et al., 2006),

including P (Chang et al., 2001; Ludwig et al., 2002; McCarty and Reeves, 2006).

Combined with minimal conventional reference methods, NIRS provides a good alternative

to routine soil analysis (Nduwamungu et al., 2009a) with at least an 80% reduction in

chemical use and laboratory costs (Foley et al., 1998). The NIRS measures the radiation

absorbed by various bonds of C-H, C-C, C-N, N-H, and O-H found in organic constituents

resulting in bending, twisting, stretching, or scissoring (Miller, 2001). Diffusely reflected

near infrared radiation is then correlated to measured material properties using various

multivariate calibration techniques (Martens and Naes, 2001; Mouazen et al., 2010).

Successful NIRS predictions have been reported for soil organic matter and texture

(Ben-Dor and Banin, 1995; Fidêncio et al., 2002; Coûteaux et al., 2003; Viscarra Rossel et

al., 2006; Stenberg, 2010) and for other soil properties including pH, CEC, N, P, K, Al, Fe,

Ca, and Zn (Reeves et al., 1999; Chang et al., 2001; Nduwamungu et al., 2009a). Although

some results appear promising, most studies use a limited number of samples (Malley et al.,

2004; Nduwamungu et al., 2009a, b, c). Nduwamungu et al. (2009b) report moderately

useful NIRS predictions for Mehlich 3 extractable Ca, Cu, and Mg, and less reliable

Page 57: Caractérisation par spectroscopie et analyse compositionnelle des

35

predictions for Al, Fe, K, Mn, P, and Zn. They conclude that further studies should

incorporate larger sample sizes and more diverse soils. To our knowledge, NIRS was used

to estimate water soluble P and total P (Bogrekci and Lee, 2005) but it has never been used

to estimate P extracted in water (Cp) according to the method of Morel et al. (2000).

Extracted soil P is often correlated with plant growth and P-uptake under controlled

conditions (Simard et al., 1991; Tran et al., 1992) and from field studies (Ziadi et al., 2001;

Messiga et al., 2010). Predicting crop P-uptake and P-budget from soil spectra would

eliminate the need to establish relationships between soil test P and crop response to P

fertilization. Börjesson et al. (1999), Terhoeven-Urselmans et al. (2008), and more recently

St. Luce et al. (2012) link NIRS soil spectra to winter cereal N-uptake and report good

predictions (Rv2 ≥ 0.70), but to our knowledge, the prediction of crop P-uptake and annual

P-budget by NIRS has not been documented.

The objective of this study was to evaluate the potential of NIRS to predict soil P-related

properties (total soil P, soil P extracted using a Mehlich 3 solution or water, annual crop P-

uptake, and annual P-budget) and other soil properties (pH, TC, TN, K, Al, Fe, Ca, Mg,

Mn, Cu, and Zn).

3.4 MATERIALS AND METHODS

3.4.1 Experimental site description

Detailed information on the experimental site is provided by Bélanger and Ziadi (2008)

and Bélanger et al. (2008). Briefly, the experiment was conducted between 1998 and 2007

on a gravely-sandy loam soil of the Saint-André series located at the Agriculture and Agri-

Food Canada research farm at Lévis, QC, Canada (46°47’ N, 71°07’ W, elevation 65 m).

The experimental design was a split-plot with four P treatments (0, 15, 30, and 45 kg P ha-

1) as main plots and four N treatments (0, 60, 120, and 180 kg N ha-1) as sub-plots. The

experiment had four replicates with a total of 64 sub-plots of equal size (1.5 m × 2.1 m).

Nitrogen (calcic ammonium nitrate) and P (triple superphosphate) fertilizers were applied

each year, during the first week of May from 1999 to 2006, prior to the start of timothy

growth. Potassium (KCl; 84 kg K ha-1) was applied with N and P to satisfy crop

requirements. The soil Mehlich-3 P content was 35.2 mg P kg­1 when the experiment was

initiated in 1998.

Page 58: Caractérisation par spectroscopie et analyse compositionnelle des

36

3.4.2 Soil and plant analyses

Soil from plots (n=64) was sampled to a depth of 15 cm in the spring before N and P

was applied, each year from 2001 to 2007. Each sample consisted of 3 to 4 soil cores (2.5-

cm diameter) taken randomly within the experimental plot. The composite samples were

carefully mixed on site, air-dried, and gently crumbled by hand to pass through a 2-mm

sieve.

Soil P available to plants was characterized in the whole sample set (n=448, set 1) by

using two methods: water extraction to determine the concentration (Cp, mg P L-1) of P

ions in solution (Morel et al., 2000; Messiga et al., 2010) and the Mehlich-3 extraction

(Mehlich, 1984). To determine Cp, 2 g of air-dried soil were mixed with 20 mL of distilled

water and 200 μL of toluene to inhibit microbial activity. The solution was gently shaken

for 16 h on a horizontal roller shaker (40 cycles min-1) before passing through a disposable

cellulose acetate filter with a 0.2-μm cut-off (Minisart, Sartorius Gottingen, Germany). For

the Mehlich-3 extraction, 2.5 g of air-dried soil was mixed with 25 mL of a Mehlich-3

solution (0.25 M NH4NO3 + 0.015 M NH4F + 0.001 M EDTA + 0.2 M CH3COOH + 0.013 M

HNO3 buffered at pH 2.3), shaken for 5 min, and then filtered through Whatman No. 42

paper. Total soil P concentration was determined in 192 samples (collected in 2001, 2003,

and 2006, set 2, Table 1) using a method adapted from Nelson (1987) and used by Messiga

et al. (2012). Briefly, 0.1 g of finely ground soil (0.2 mm) was mixed in a 50-mL boiling

flask with 0.5 g K2S2O8 and 10 mL 0.9 M H2SO4, and digested at 121ºC in an autoclave for

90 min. Following Mehlich 3, water, and total P extractions, P was quantified by the

colorimetric blue method (Murphy and Riley, 1962). Also, Mehlich-3 P (M3P) was

measured by the inductively coupled plasma (ICP) emission spectroscopy (M3P_ICP) in

192 samples collected in 2005, 2006, and 2007 (set 3, Table 1).

Soil pH was measured in distilled water with a 1:2 soil to solution ratio (Hendershot et

al., 1993). Total C and TN were quantified by dry combustion with a LECO CNS-1000

analyzer (LECO Corp., St. Joseph, MI). The concentrations of K, Al, Fe, Ca, Mg, Mn, Cu,

and Zn were measured by ICP emission spectroscopy after the Mehlich 3 extraction

(Mehlich, 1984). Potassium, Al, and Fe were analyzed from samples collected in 2005,

2006, and 2007 (n=192, set 3), whereas Ca, Mg, Mn, Cu, and Zn were determined from

Page 59: Caractérisation par spectroscopie et analyse compositionnelle des

37

samples collected in 2005 and 2007 (n=128, set 4, Table 1). One chemical determination of

each soil property was done on soil samples.

From 2001 to 2006, timothy was harvested twice a year (n=378, set 5); the first harvest

was in mid-June, at the late heading stage of development, and the second harvest in early

August. Dry matter (DM) yield of each plot was determined from strips (0.91 m wide × 2.1

m long) harvested at a 5-cm height using a self-propelled flail forage harvester (Carter

MGF Co. Inc., Brookston, IN). A forage sample of approximately 500 g was collected from

each plot, dried at 55°C in a forced-draft oven for 3 d, and ground with a Wiley mill

(Standard model 3, Arthur H. Thomas Co., Philadelphia, PA) fitted with a 1-mm screen.

Plant samples of 0.1 g were digested using a mixture of sulphuric and selenious acids, as

described by Isaac and Johnson (1976). Phosphorus concentration was measured with a

QuikChem 8000 Lachat autoanalyzer (Lachat Instruments) using the Lachat method 13-

115-01-2-A (Lachat Instruments, 2011). The P-uptake at each harvest was calculated as the

product of forage P concentration and DM yield. Annual DM yield and crop P-uptake were

the sum of their first and second harvest values. The annual P-budget was computed as the

difference between P applied as fertilizer and annual P-uptake, as reported in Messiga et al.

(2012).

3.4.3 Near-infrared reflectance spectroscopy spectrum acquisition

Each soil sample was mixed and scanned by measuring his absorbance [log (1/R), where

R is reflectance] in the visible and near-infrared regions between 400 and 2500 nm at 2-nm

intervals using a NIRSystems 6500 monochromator Instrument (Foss NIRSystems Inc.,

Silver Spring, MD) with a cup (quarter cup, rectangular ¼) containing approximately 25

mL of the soil sample. This NIRS instrument is equipped with a tungsten-halogen light

source, a silicon detector for wavelength between 400-1100, a Pbs (Lead (II) Sulfide)

detector for wavelength in the range of 1100-2500 nm, and two intern standards

(polystyrene and didymium) that are used during sample spectrum acquisition. Each

spectrum was the mean of 16 co-added scans. A check test was performed prior to scanning

the soil sample and a performance test was done daily. One randomly selected soil sample

was scanned 12 times to create a repeatability file that was used to account for possible

operator errors and to improve calibration equations by minimizing errors associated with

soil heterogeneity and compaction (Nie et al., 2009).

Page 60: Caractérisation par spectroscopie et analyse compositionnelle des

38

3.4.4 Pretreatment, calibration, and cross-validation

To improve the calibration models of each property, the following 40 spectral pre-

treatments (2 × 2 × 2 × 5 factorial arrangement) were tested using WinISI III (ver.1.61)

software (Infrasoft International, LLC, Silver Spring, MD): (critical T-outlier values of 2.0

and 2.5) × (with and without a repeatability file) × (400-2500 nm and 1100-2500 nm

wavelength section) × (1-4-4-1, 2-8-6-1, 2-10-10-1, 0-4-4-1, and 0-8-6-1). Low and high

limits of the critical T-statistic, for T-outlier detection, were set to 2.0 and 2.5, respectively.

The math treatments that were compared are identified with four numbers (i.e., 1-4-4-1);

the first number is the derivative order, the second is the size of the gap in nm, the third is

the number of smoothing points, and the last is the second smooth (Ludwig et al., 2002;

Coûteaux et al., 2003). For each property, two criteria were used to select the best of the 40

spectral pre-treatments: simultaneous low standard error (SE) and high coefficient of

determination in cross-validation (1-VR) [Nduwamungu et al., 2009a]. The spectral pre-

treatments selected for each soil property are listed in tables 2 and 3, and only the results

using the best pre-spectral treatments are presented. Scatter correction with standard normal

variate and detrending (SNVD) was used to remove, or reduce, particle size and noise

effects (Brunet et al., 2007). The modified partial least squares regression method (PLSR)

of the WinISI III software was used to develop calibration equations for the soil and crop

properties. To maximize the probability of developing a robust calibration equation for

each property, a maximal number of soil samples, corresponding to, 80% of each soil

sample set, were randomly selected by the software to be used for the calibration set, and

the remainder samples were used for the validation set (Ludwig et al., 2002; Brunet et al.,

2007; St. Luce et al., 2012). General calibration equations were selected based on Martens

and Naes (2001) as follows: Reference data = f (spectral data) + SEC, where f () means

“function of” and SEC is the standard error of calibration. The best NIRS calibration

equations were the ones that minimize the SEC. Cross-validation was performed by using

four sub-groups from the calibration set in order to choose the optimal number of terms and

to avoid over-fitting the calibration model (Shenk and Westerhaus, 1991).

Page 61: Caractérisation par spectroscopie et analyse compositionnelle des

39

3.4.5 Validation

Calibration equations were validated using WinISI III software by comparing predicted

against reference values. Predicted values were generated using the modified PLSR method

of the WinISI III software according to Martens and Naes (2001): Predicted values = f

(spectrum data) = Reference data + error.

The accuracy of NIRS predictions was assessed with the following statistics: the

coefficient of determination of validation (Rv2) and the ratio of standard error of prediction

to standard deviation (RPD), which is the standard deviation of samples in the validation

set (SD) divided by the standard error of prediction corrected for the bias (SEP(C)) [RPD =

SD/SEP(C)]. Calibration equations were considered to be excellent when Rv2 > 0.95 and

RPD > 4.00; successful when 0.90 ≤ Rv2 ≤ 0.95 and 3.00 ≤ RPD ≤ 4.00; moderately

successful when 0.80 ≤ Rv2 < 0.90 and 2.25 ≤ RPD < 3.00; moderately useful when

0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25; and less reliable when Rv

2 < 0.70 and

RPD < 1.75 (Malley et al., 2004). The coefficient of variation (CV) in the reference set was

defined as the SD divided by the mean of chemical values, whereas the coefficient of

variation in the calibration set was computed as the ratio of standard error of calibration

(SEC) to the mean of calibration data (Williams, 2001).

Predictive graphs, illustrating the relationships between predicted and reference values

for P-related properties, were created with SigmaPlot for Windows (SYSTAT, 2012,

version 12.1). Pearson correlations between soil total C and soil and plant properties

analyzed on samples collected in 2001, 2003, and 2006 [soil P content extracted using the

Mehlich 3 method and analysed by colorimetry (M3P_Col), M3P_ICP, Cp, TP, P-uptake,

P-budget, TN, and pH] were computed with the SPSS 19 software (SPSS, 2010).

3.5 RESULTS AND DISCUSSION

3.5.1 Reference data

The CV values for the measured properties were relatively high (> 50%, Table 1) for Cp,

annual P-budget, and Mn and intermediate (20–50%) for M3P_Col, M3P_ICP, annual P-

uptake, K, Mg, and Zn. The CV values for the measured TP, TC, TN, pH, Al, Fe, Ca, and

Cu were low (< 20%). This variability of the soil properties is mostly related to direct and

indirect effects of the different N and P fertilization treatments during the seven years of the

Page 62: Caractérisation par spectroscopie et analyse compositionnelle des

40

study. According to Dardenne et al. (2000), a wide range of values for a given property is

required to obtain high NIRS calibration accuracy and thus good predictive performance.

3.5.2 Spectral pretreatments

Spectral pre-treatments that gave the best calibration equation for each property are

provided in Tables 2 and 3. Based on the Rv2 and RPD statistics, selected calibration

equations performed slightly better for M3P_Col, Cp, TC, TN, pH, K, Fe, Mg, Cu, and Zn

when a repeatability file was used. Also, M3P_ICP, P-uptake, TP, TC, TN, Fe, and Ca were

better predicted with the wavelength region of 1100–2500 nm than with the 400–2500 nm

region, whereas the other properties were more accurately predicted with the whole visible-

near-infrared spectrum. McCarty and Reeves (2006) report that the use of mid-infrared

spectra may yield better calibrations, than the NIR spectral region, for K, Ca, and Mg

extracted with Mehlich 1 and analysed using atomic absorption spectroscopy.

3.5.3 Near-infrared reflectance spectroscopy prediction of soil and crop P properties

Statistics for calibration and cross-validation are listed in Table 2. The number of T-

outliers was 20 for Cp and less than 8 for the other P properties, indicating that the

development of calibration equations was based on more than 94% of the soil samples in

the calibration set. Based on the high standard error and CV (> 23%) and the low

coefficients of determination of calibration equations (Rc2 ≤ 0.70) for M3P_Col, M3P_ICP,

Cp, P-uptake, and P-budget, NIRS calibration performances were considered poor for these

properties. As a result, cross-validation of calibration equations showed high standard error

of cross-validation (SECV) and low coefficient of determination of cross-validation (1-VR)

values (≤ 0.55). However, the calibration for TP resulted in an acceptable coefficient of

determination (Rc2 = 0.78) while the cross-validation was acceptable with a 1-VR of 0.76.

The number of soil samples used in the validation set varied between 38 and 90

depending on the P-related property (Table 2). Slopes of the linear regression (b) between

reference and predicted values of M3P_Col, M3P_ICP, Cp, P-uptake, and P-budget from

the validation set were less than 0.60 while the Rv2 (≤ 0.49) and RPD values (≤ 1.37) were

low (Fig. 3.1). The relationship between reference and predicted values for these properties

was therefore poor (b < 0.80, Rv2 < 0.70, RPD < 1.75; Nduwamungu et al., 2009c),

indicating that these properties cannot be predicted by NIRS.

Page 63: Caractérisation par spectroscopie et analyse compositionnelle des

41

This result agrees with Nduwamungu et al. (2009b) which reports a poor calibration

performance for M3P when analysed by ICP emission spectroscopy on a limited number of

soil samples (n = 150). Chang et al. (2001) also reports a low accuracy for M3P_ICP

prediction using NIRS with a principal component regression technique (Rv2 = 0.40, RPD =

1.18). Similarly, McCarty and Reeves (2006) found that soil P content extracted using the

Mehlich 1 method and analysed by colorimetry cannot be predicted by NIRS (Rv2 = 0.21).

However, NIRS has been shown to be useful (Rv2 = 0.71; RPD = 1.81) to predict P when

measured by the Olsen method (van Groenigen et al., 2003). Thus, it appears that the

performance of NIRS calibration could be affected by the reference method used that might

produce different reference values. Morón and Cozzolino (2007) report that NIRS

predictions of soil P were slightly more reliable when based on a resin extracted P method

(Rv2 = 0.61, RPD = 2.2) rather than the Bray method (Rv

2 = 0.58, RPD = 1.72). Sørensen

and Dalsgaard (2005) suggest that NIRS could be useful to predict soil P if there is an

indirect relationship between soil P and organic components, which means that P relates to

NIRS by covariation (Stenberg, 2010). Indeed, Ludwig et al. (2002) reports useful

calibration for soil P measured by the Olsen method, which was highly correlated with soil

C content (r = 0.67). In our study, soil C content was not significantly correlated to

M3P_Col (r = -0.04, P = 0.56), M3P_ICP (r = 0.10, P = 0.45), P-uptake (r = 0.08,

P = 0.30), and P-budget (r = 0.02, P = 0.80), and significantly but weakly correlated to Cp

(r = 0.31, P < 0.001, data not shown). Moderately useful NIRS prediction (Rv2 = 0.78) was

previously reported for water soluble P in 150 fine sandy soil samples collected from three

sites in Florida (USA, Bogrekci and Lee, 2005), but to our knowledge, the potential of

NIRS to predict Cp (Morel et al., 2000), P-uptake, and P-budget has not previously been

studied.

Successful calibration was found for TP as indicated by high Rc2 and 1-VR values,

which resulted in a moderately useful prediction of TP from the validation set (Fig. 1) with

Rv2 = 0.75 and RPD = 1.98. This result can be explained, in part, by the fact that TP

contains a certain proportion of organic P that is related to organic matter (Turner et al.,

2005). Bogrekci and Lee (2005) report successful NIRS prediction of TP (Rv2 = 0.92) in

150 fine sandy soil samples. Future research is needed to verify whether the ability of NIRS

Page 64: Caractérisation par spectroscopie et analyse compositionnelle des

42

to predict soil TP is related to soil texture and to evaluate the potential for NIRS to predict

soil organic P, since it is highly correlated with the concentration of organic matter.

3.5.4 Near-infrared reflectance spectroscopy prediction of other soil properties

Statistics of calibration and validation for all other soil properties are provided in Table

3. The number of T-outliers excluded from the calibration set was < 14; at least 89% of the

calibration samples were used to generate the prediction equations (Table 3). The NIRS

predictions were moderately successful for TC (Rv2 = 0.87; RPD = 2.82) and moderately

useful for TN (Rv2 = 0.79; RPD = 2.2). Chang and Laird (2002) also report successful NIRS

predictions (0.86 < Rv2 < 0.91; 2.8 < RPD < 4.4) for these two properties from 108 samples

obtained from a wide range of soil groups and cropping histories. Soil pH was successfully

predicted by NIRS (Table 3); our predictions were better than those reported by Dunn et al.

(2002) [Rv2 = 0.80, RPD = 2.3] and He et al. (2007) [Rv

2 = 0.82]. Chang et al. (2001) link

the accurate prediction of pH to its significant correlation with clay content and soil organic

matter. In our study, pH was significantly correlated to soil TC (r = 0.52, P < 0.001) and

TN (r = 0.54, P < 0.001, data not shown). In contrast, Nduwamungu et al. (2009a) report a

less reliable prediction for pH due to weak correlations between soil pH and soil TC and

TN.

Excellent NIRS predictions were found for Ca and Mn (Rv2 > 0.95, RPD > 4.0).

Furthermore, predictions were successful for Mg and moderately successful for Al. The Fe

and Zn calibration equations were moderately useful, while the K and Cu calibration

equations had the lowest Rv2 (< 0.70) and RPD (< 1.75) which means that their NIRS

predictions were unacceptable.

Compared with our results, Nduwamungu et al. (2009b) report lower NIRS prediction

performance for Al, Fe, Ca, Mn, and Zn, and greater calibration accuracy for Mg and Cu

when extracted by the Mehlich 3 method and analysed by atomic absorption spectroscopy.

The ability of NIRS to predict certain soil properties may be related to their correlation with

primary soil properties such as texture (Chang et al., 2001; Nduwamungu et al., 2009a), or

organic matter content (Dalal and Henry, 1986). The CV values in the reference data used

by Nduwamungu et al. (2009b) were actually greater than those from our data set (Table 1),

except for K and Mn; however, we found better NIRS predictions for the majority of the

soil properties. Similarly, Chang et al. (2001) report similar NIRS prediction performance

Page 65: Caractérisation par spectroscopie et analyse compositionnelle des

43

for soil K and Cu to ours, extracted by Mehlich 3 and determined by ICP, but their

predictions for Fe, Ca, Mg, Mn, and Zn were less accurate despite having a reference data

set with a large variability. Hence, high CVs that reflect a large range in soil properties do

not always guarantee good NIRS predictions.

3.6 CONCLUSIONS

The current study showed that soil P-related properties of a gravely loam soil,

including M3P_Col, M3P_ICP, Cp, annual P-uptake, and annual P-budget, were not

accurately predicted by NIRS. These unsatisfactory NIRS predictions may be related to the

low correlation observed between the P-related properties and soil C content. However,

NIRS predictions were considered moderately useful for soil TP, TN, Fe, and Zn,

moderately successful for TC and Al, successful for pH and Mg, and excellent for Ca and

Mn. Although NIRS can predict several soil properties, prediction of total P was the only

soil P-related property, correlated to soil C, that was moderately useful. These findings on

homogenous textured soil samples should be validated on soils of diverse textures.

3.7 ACKNOWLEDGEMENTS

The authors acknowledge the technical assistance of Mario Laterrière, Claude

Levesque, and Danielle Mongrain (Agriculture and Agri-Food Canada, Soils and Crops

Research and Development Centre, Québec, Canada). We also thank Mervin St. Luce for

his comments on an early version of this manuscript and acknowledge the assistance of

Christina McRae, from Editworks (Nova Scotia, Canada), for the structural editing of this

manuscript. This study was funded by The Sustainable Agriculture Environmental Systems

(SAGES) initiative of Agriculture and Agri-Food Canada.

Page 66: Caractérisation par spectroscopie et analyse compositionnelle des

44

3.8 REFERENCES

Bélanger, G., G.F. Tremblay, and D. Mongrain. 2008. Yield and nutritive value of the

spring growth of an aging timothy sward. Can. J. Plant Sci. 88:457–464.

Bélanger, G., and N. Ziadi. 2008. Phosphorus and nitrogen relationships during spring

growth of an aging timothy sward. Agron. J. 100:1757–1762.

Ben-Dor, E., and A. Banin. 1995. Near infrared analysis as a rapid method to

simultaneously evaluate several soil properties. Soil Sci. Soc. Am. J. 59:364–372.

Bogrekci, I., and W.S. Lee. 2005. Spectral phosphorus mapping using diffuse reflectance of

soils and grass. Biosystems. Eng. 91:305–312.

Börjesson, T., B. Stenberg, B. Lindén, and A. Jonsson. 1999. NIR spectroscopy, mineral

nitrogen analysis and soil incubations for the prediction of crop uptake of nitrogen

during the growing season. Plant Soil 214:75–83.

Bray, R.H., and L.T. Kurtz. 1945. Determination of total, organic, and available forms of

phosphorus in soils. Soil Sci. 59:39–45.

Brunet, D., B.G. Barthès, J.-L. Chotte, and C. Feller. 2007. Determination of carbon and

nitrogen contents in Alfisols, Oxisols and Ultisols from Africa and Brazil using NIRS

analysis: Effects of sample grinding and set heterogeneity. Geoderma 139:106–117.

Chang, C.W., and D.A. Laird. 2002. Near-infrared reflectance spectroscopic analysis of soil

C and N. Soil Sci. 167:110–116.

Chang, C.W., D.A. Laird, M.J. Mausbach, and C.R.Jr. Hurburgh. 2001. Near-infrared

reflectance spectroscopy – Principal components regression analyses of soil

properties. Soil Sci. Soc. Am. J. 65:480–490.

Coûteaux, M.M., B. Berg, and P. Rovira. 2003. Near infrared reflectance spectroscopy for

determination of organic matter fractions including microbial biomass in coniferous

forest soils. Soil Biol. Biochem. 35:1587–1600.

Dalal, R.C., and R.J. Henry. 1986. Simultaneous determination of moisture, organic carbon,

and total nitrogen by near infrared reflectance, spectroscopy. Soil Sci. Soc. Am. J.

50:120–123.

Dardenne, P., G. Sinnaeve, and V. Baeten. 2000. Multivariate calibration and chemometrics

for near infrared spectroscopy: which method? J. Near Infrared Spectrosc. 8:229–237.

Dunn, B.W., H.G. Beecher, G.D. Batten, and S. Ciavarella. 2002. The potential of near-

infrared reflectance spectroscopy for soil analysis – a case study from the Riverine

Plain of south-eastern Australia. Aus. J. Exp. Agri. 42:607–614.

Fidêncio, P.H., R.J. Poppi, and J.C. de Andrade. 2002. Determination of organic matter in

soils using radial basis function networks and near infrared spectroscopy. Anal. Chim.

Acta 453:125–134.

Foley, W.J., A. McIlwee, I.R. Lawler, L.V. Aragones, A.P. Woolnough, and N. Berding.

1998. Ecological applications of near infra-red spectroscopy - a tool for rapid, cost-

effective prediction of the composition of plant and animal tissues and aspects of

animal performance. Oecologia 116:293–305.

He, Y., M. Huang, A. García, A. Hernández, and H. Song. 2007. Prediction of soil

macronutrients content using near-infrared spectroscopy. Comput. Electron. Agric.

58:144–153.

Page 67: Caractérisation par spectroscopie et analyse compositionnelle des

45

Hendershot, W.H., H. Lalande, and M. Duquette. 1993. Soil reaction and exchangeable

acidity. p. 141–145. In Carter, M.R. (ed.) Soil Sampling and Methods of Analysis for

Canadian Society of Soil Science. Lewis Publishers, Boca Raton, FL.

Isaac, R.A., and W.C. Johnson. 1976. Determination of total nitrogen in plant tissue, using

a block digestor. J. Assoc. Anal. Chem. 59:98–100.

Lachat Instruments. 2011. Methods list for automated ion analyzers (flow injection

analyses, ion chromatography). Available at

http://www.lachatinstruments.com/applications/Methods.asp (verified 27 Feb. 2012).

Lachat Instruments, Loveland, CO.

Ludwig, B., P.K. Khanna, J. Bauhus, and P. Hopmans. 2002. Near infrared spectroscopy of

forest soils to determine chemical and biological properties related to soil

sustainability. F. Ecol. Mang. 171:121–132.

Malley, D.F., E. Ben-Dor, and P.D. Martin. 2004. Application in analysis of soils. p. 729–

84. In Roberts, C.A., J. Jr. Workman, and J.B. Reeves III (ed.) Near-infrared

Spectroscopy in Agriculture. Am. Soc. Agr., Crop Sc. Soc. Am., and Soil Sc. Soc.

Am., Madison, USA.

Martens, H., and T. Naes. 2001. Multivariate calibration by data compression. p. 59–100.

In Williams, P.C., and K.H. Norris (ed.) Near Infrared Technology in the Agricultural

and Food Industries. 2nd ed. American Association of Cereal Chemists, St. Paul, MN.

McCarty, G.W., and J.B. Reeves, III. 2006. Comparison of near infrared and mid infrared

diffuse reflectance spectroscopy for field-scale measurement of soil fertility

parameters. Soil Sci. 171:94–102.

Mehlich, A. 1984. Mehlich 3 soil test extractant: A modification of Mehlich 2 extractant.

Commun. Soil Sci. Plant Anal. 15:1409–1416.

Messiga, A.J., N. Ziadi, C. Morel, and L.É. Parent. 2010. Soil phosphorus availability in

no-till versus conventional tillage following freezing and thawing cycles. Can. J. Soil

Sci. 90:419–428.

Messiga, A. J., N. Ziadi, G. Bélanger, and C. Morel. 2012. Process-based mass-balance

modeling of soil phosphorus availability in a grassland fertilized with N and P. Nutr.

Cycling Agroecosyst. 92:272-287.

Miller, C.E. 2001. Chemical principles of near-infrared technology. p. 19–8. In Williams,

P.C., and K.H. Norris (ed.) Near Infrared Technology in the Agricultural and Food

Industries 2nd ed. American Association of Cereal Chemists, St. Paul, MN.

Morel, C., H. Tunney, D. Plenet, and S. Pellerin. 2000. Transfer of phosphate ions between

soil and solution: Perspectives in soil testing. J. Environ. Qual. 29:50–59.

Morón, A., and D. Cozzolino. 2007. Measurement of phosphorus in soils by near infrared

reflectance spectroscopy: Effect of reference method on calibration. Commun. Soil

Sci. Plant Anal. 38:1965–1974.

Mouazen, A.M., B. Kuang, J. De Baerdemaeker, and H. Ramon. 2010. Comparison among

principal component, partial least squares and back propagation neural network

analyses for accuracy of measurement of selected soil properties with visible and near

infrared spectroscopy. Geoderma 158:23–31.

Murphy, J., and J.P. Riley. 1962. A modified single solution method for the determination

of phosphate in natural waters. Anal. Chim. Acta 27:31–36.

Nduwamungu, C., N. Ziadi, G.F. Tremblay, and L.-É. Parent. 2009a. Near-infrared

reflectance spectroscopy prediction of soil properties: Effects of sample cups and

preparation. Soil Sci. Soc. Am. J. 73:1896–1903.

Page 68: Caractérisation par spectroscopie et analyse compositionnelle des

46

Nduwamungu, C., N. Ziadi, L.-É. Parent, and G.F. Tremblay. 2009b. Mehlich 3 extractable

nutrients as determined by near-infrared reflectance spectroscopy. Can. J. Soil Sci.

89:579–587.

Nduwamungu, C., N. Ziadi, L.-É. Parent, G.F. Tremblay, and L. Thuriès. 2009c.

Opportunities for, and limitations of, near infrared reflectance spectroscopy

applications in soil analysis: A review. Can. J. Soil Sci. 89:531–541.

Nelson, N.S. 1987: An acid‐persulfate digestion procedure for determination of phosphorus

in sediments. Commun. Soil Sci. Plant Anal. 18:359–369.

Nie, Z., G.F. Tremblay, G. Bélanger, R. Berthiaume, Y. Castonguay, A. Bertrand, R.

Michaud, G. Allard, and J. Han. 2009. Near-infrared reflectance spectroscopy

prediction of neutral detergent-soluble carbohydrates in timothy and alfalfa. J. Dairy

Sci. 92:1702–1711.

Olsen, S.R., C.V. Cole, F.S. Watanabe, and L.A. Dean. 1954. Estimation of available

phosphorus by extraction with sodium bicarbonate. USDA Circ. 939, U.S. Govt.

Print. Office, Washington, D.C.

Reeves, III J.B., G.W. McCarty, and J.J. Meisinger. 1999. Near infrared reflectance

spectroscopy for the analysis of agricultural soils. J. Near Infrared Spectrosc. 7:179–

193.

Simard, R.R., T.S. Tran, and J. Zizka. 1991. Strontium chloride-citric acid extraction

evaluated as a soil testing procedure for phosphorus. Soil Sci. Soc. Am. J. 55:14–421.

Sørensen, L.K., and S. Dalsgaard. 2005. Determination of clay and other soil properties by

near infrared spectroscopy. Soil Sci. Soc. Am. J. 69:159–167.

SPSS. 2010. IBM SPSS Statistics 19 Core System User’s Guide. Version 19. IBM

Company, US.

Shenk, J.S., and M.O. Westerhaus. 1991. Population definition, sample selection, and

calibration procedures for near infrared reflectance spectroscopy. Crop Sci. 31:469–

474.

St. Luce, M., N. Ziadi, J. Nyiraneza, G.F. Tremblay, B.J. Zebarth, J.K. Whalen, and M.

Laterrière. 2012. Near infrared reflectance spectroscopy prediction of soil nitrogen

supply in humid temperate regions of Canada. Soil Sci. Soc. Am. J. 76:1454–1461.

Steffens, D. 1994. Phosphorus release kinetics and extractable phosphorus after long-term

fertilization. Soil Sci. Soc. Am. J. 58:1702–1708.

Stenberg, B. 2010. Effects of soil sample pretreatments and standardised rewetting as

interacted with sand classes on Vis-NIR predictions of clay and soil organic carbon.

Geoderma 158:15–22.

SYSTAT. 2012. SigmaPlot 12.0. SYSTAT Software, Inc., Point Richmond, CA.

Terhoeven-Urselmans, T., H. Schmidt, R. Georg Joergensen, and B. Ludwig. 2008.

Usefulness of near-infrared spectroscopy to determine biological and chemical soil

properties: Importance of sample pre-treatment. Soil Biol. Biochem. 40:1178–1188.

Tran, T.S., R.R. Simard, and M. Tabi. 1992. Evaluation of the electro-ultrafiltration

technique (EUF) to determine available P in neutral and calcareous soils. Commun.

Soil Sci. Plant Anal. 23:2261–2281.

Turner, B.L., B.J. Cade-Menun, L.M. Condron, and S. Newman. 2005. Extraction of soil

organic phosphorus. Talanta 66:294–306.

van Groenigen, J.W., C.S. Mutters, W.R. Horwath, and C. van Kessel. 2003. NIR and

DRIFT-MIR spectrometry of soils for predicting soil and crop parameters in a flooded

field. Plant Soil 250:155–165.

Page 69: Caractérisation par spectroscopie et analyse compositionnelle des

47

Viscarra Rossel, R.A.V., D.J.J. Walvoort, A.B. McBratney, L.J. Janik, and J.O. Skjemstad.

2006. Visible, near infrared, mid infrared or combined diffuse refelectance

spectroscopy for simultaneous assessment of various soil properties. Geoderma

131:59–75.

Williams, P.C. 2001. Implementation of near-infrared spectroscopy. p. 145–169. In

Williams, P.C., and K. Norris (ed.) Near-infrared Technology in the Agricultural and

Food Industries. 2nd ed. American Association of Cereal Chemists, St. Paul, MN.

Ziadi, N., R.R. Simard, T.S. Tran, and G. Allard. 2001. Evaluation of soil-available

phosphorus for grasses with Electro-Ultrafiltration technique and some chemical

extractions. Can. J. Soil Sci. 81:167–174.

Ziadi, N., and T. Tran. 2007. Mehlich 3-extractable elements. In: M.R. Carter et al., editors,

Soil sampling and methods of analysis. Lewis, Boca Raton, p. 81-88.

Page 70: Caractérisation par spectroscopie et analyse compositionnelle des

48

Tableau 3-1 Descriptive statistics† for the soil P-related and other properties analyzed

using reference methods.

Property Sampled years N (set #) Reference

method Min Max Mean SD CV (%)

mg kg-1

M3P_Col 2001–2007 448 (1) Colorimetry 4 102 40 19 48

M3P_ICP 2005–2007 192 (3) ICP 17 124 54 24 44

TP 2001+ 2003 + 2006 192 (2) Colorimetry 450 1242 729 136 19

mg L-1

Cp 2001–2007 448 (1) Colorimetry 0.06 1.28 0.33 0.20 61

kg ha-1

P-uptake 2001–2006 378 (5) Colorimetry 1.6 27.7 13.5 5.9 44

P-budget 2001–2006 378 (5) -26.5 42.4 9 17 189

g kg-1

TC 2001+ 2003 + 2006 192 (2) Dry combustion 20.2 31.3 25.2 2.2 9

TN 2001+ 2003 + 2006 192 (2) Dry combustion 1.8 2.8 2.2 0.2 9

pH 2001+ 2003 + 2006 192 (2) Water (1:2) 4.6 6.4 5.5 0.4 7

mg kg-1

K 2005–2007 192 (3) ICP‡ 46 332 133 61 46

Al 2005–2007 192 (3) ICP 769 1225 995 95 10

Fe 2005–2007 192 (3) ICP 162 315 233 30 13

Ca 2005 + 2007 128 (4) ICP 986 2314 1559 271 17

Mg 2005 + 2007 128 (4) ICP 134 449 280 81 29

Mn 2005 + 2007 128 (4) ICP 13 110 36 23 64

Cu 2005 + 2007 128 (4) ICP 1.1 2.6 1.6 0.3 19

Zn 2005 + 2007 128 (4) ICP 0.8 4.2 2 0.6 30 †N, number of samples; Min, minimum; Max, maximum, SD, standard deviation; CV, coefficient of variation

[(SD/mean) × 100]. ‡ICP, inductively coupled plasma.

Page 71: Caractérisation par spectroscopie et analyse compositionnelle des

49

Tableau 3-2 NIRS spectral pre-treatments and statistics† of calibration, cross-validation,

and validation for the P-related soil properties.

M3P_Col M3P_ICP Cp TP Annual

P-uptake Annual

P-budget

Statistic (mg kg-1)‡ (mg L-1)‡ (mg kg-1)‡ (kg P ha-1)‡

Pre-treatment Math treat 2,10,10,1 0,8,6,1 1,4,4,1 0,4,4,1 1,4,4,1 0,8,6,1 T 2.5 2.5 2.5 2.5 2.5 2.5 Region (nm) 400–2500 1100–2500 400–2500 1100–2500 1100–2500 400–2500 Rep File Yes No Yes No No No

Calibration Nc 350 151 338 153 295 298 T-outliers 8 3 20 1 7 4 Mean 40 54 0.30 721 13.7 8.9 SEC 16 21 0.12 61 3.1 13.2 CV (%) 40 39 40 8 23 148 Rc

2 0.43 0.23 0.46 0.78 0.70 0.40 Cross-validation

SECV 16 22 0.12 63 3.9 14.4 1-VR 0.30 0.17 0.43 0.76 0.55 0.29

Validation Nv 90 38 90 38 76 76 Mean 38.0 54.3 0.30 762 13.1 9.3 SD 18.1 24.6 0.17 156.1 5.8 15.7 SEP(C) 15.7 21.4 0.13 78.8 4.2 14.8 †Math treat, mathematical treatment; T, critical outlier value; Region, wavelength region of the spectrum that

was used; Rep File, repeatability file; Nc, number of samples used for calibration; T-outliers, outliers

eliminated during calibration; SEC, standard error of calibration; CV, coefficient of variation defined as the

ratio of SEC to the mean; Rc2, coefficient of determination of calibration; SECV, standard error of cross-

validation; 1-VR, coefficient of determination of cross-validation; Nv, number of samples used for validation;

SD, standard deviation; SEP(C), standard error of prediction corrected for the bias. ‡The units apply only to Means, SEC, SECV, SD, and SEP(C).

Page 72: Caractérisation par spectroscopie et analyse compositionnelle des

50

Tableau 3-3 NIRS spectral pre-treatment and statistics† of calibration, cross-validation, and validation for the other soil

properties.

TC TN pH K Al Fe Ca Mg Mn Cu Zn

Statistic (g kg-1)‡ (mg kg-1)‡

Pre-treatment

Math treat 2,10,10,1 2,8,6,1 2,10,10,1 1,4,4,1 1,4,4,1 2,8,6,1 1,4,4,1 2,8,6,1 2,8,6,1 0,8,6,1 0,4,4,1

T 2.5 2.5 2.5 2.5 2 2 2 2 2.5 2 2 Region (nm) 1100–2500 1100–2500 400–2500 400–2500 400–2500 1100–2500 1100–2500 400–2500 400–2500 400–2500 400–2500 Rep File Yes Yes Yes Yes No Yes No Yes No Yes Yes

Calibration

Nc 148 153 151 150 140 141 94 93 97 91 93 T-outliers 6 1 3 4 14 13 8 9 5 11 9 Mean 25.3 2.2 5.5 129 997 232 1564 287 35 1.7 1.9 SEC 0.5 0.1 0.1 25 31 13 33 13 3 0.1 0.2 CV (%) 1.9 4.5 1.8 19.4 3.1 5.6 2.1 4.5 8.5 5.8 10.5 Rc

2 0.93 0.79 0.94 0.79 0.89 0.81 0.98 0.97 0.98 0.84 0.83

Cross-validation

SECV 0.7 0.1 0.1 36 43 16 62 18 5 0.1 0.2 1-VR 0.88 0.73 0.89 0.58 0.80 0.70 0.94 0.94 0.95 0.79 0.78

Validation

Nv 38 38 38 38 38 38 25 25 25 25 25 Mean 24.9 2.2 5.4 147 989 231 1487 260 38 1.5 2.0 SD 2.23 0.22 0.42 71 81 29 285 83 23 0.22 0.52 SEP(C) 0.79 0.10 0.13 44.6 31.6 14.1 56.3 20.5 4.1 0.19 0.25 Rv

2 0.87 0.79 0.91 0.62 0.85 0.77 0.96 0.94 0.97 0.37 0.78 RPD 2.82 2.20 3.23 1.59 2.56 2.05 5.06 4.04 5.60 1.16 2.08 Prediction§ MS MU S LR MS MU E S E LR MU †Math treat, mathematical treatment; T, critical outlier value; Rep File, repeatability file; Nc, number of samples used for calibration; T-outliers, outliers

eliminated during calibration; SEC, standard error of calibration; CV, coefficient of variation defined as the ratio of SEC to the mean multiplied by 100; Rc2,

coefficient of determination of calibration; SECV, standard error of cross-validation; 1-VR, coefficient of determination of cross-validation, Nv, number of

Page 73: Caractérisation par spectroscopie et analyse compositionnelle des

51

samples used for validation, SD, standard deviation; SEP(C), standard error of prediction corrected for the bias; Rv2, coefficient of determination of validation;

RPD, ratio of standard error of prediction to standard deviation which is the SD of samples in the validation set divided by the SEP(C). ‡The units apply only to Means, SD, SEC, SECV, SD, and SEP(C). §Based on validation statistics, the NIRS predictions were considered excellent (E) when Rv

2 > 0.95 and RPD > 4; successful (S) when 0.90 ≤ Rv2 ≤ 0.95 and

3 ≤ RPD ≤ 4; moderately successful (MS) when 0.80 ≤ Rv2 < 0.90 and 2.25 ≤ RPD < 3; moderately useful (MU) when 0.70 ≤ Rv

2 < 0.80 and 1.75 ≤ RPD < 2.25;

and less reliable (LR) when Rv2 < 0.70 and RPD < 1.75 (Malley et al., 2004).

Page 74: Caractérisation par spectroscopie et analyse compositionnelle des

52

0 20 40 60 80 100

0

20

40

60

80

100

(a) M3P_Col (mg kg-1

)

y = 0.31x + 28.86

Rv

2 = 0.25, RPD = 1.15, LR

0 20 40 60 80 100 120

0

20

40

60

80

100

120

y = 0.27x + 39.55

Rv

2 = 0.24, RPD = 1.15, LR

0.0 0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

(c) Cp (mg L-1

)

y = 0.41x + 0.18

Rv

2= 0.45, RPD = 1.33, LR

400 600 800 1000 1200 1400

400

600

800

1000

1200

1400

(d) TP (mg kg-1

)

y = 0.73x + 170.98

Rv

2 = 0.75, RPD = 1.98, MU

0 5 10 15 20 25 30

0

5

10

15

20

25

30

(e) Annual P-uptake (kg ha-1

)

y = 0.60x + 6.67

Rv

2 = 0.49, RPD = 1.37, LR

(b) M3P_ICP (mg kg-1

)

-30 -20 -10 0 10 20 30 40 50

-30

-20

-10

0

10

20

30

40

50

(f) Annual P-budget (kg ha-1

)

NIR

S P

red

icte

d V

alu

es

Measured Values

y = b x + ay = x

y = 0.24x + 7.03

Rv

2 = 0.16, RPD = 1.06, LR

Page 75: Caractérisation par spectroscopie et analyse compositionnelle des

53

Figure 3-1 NIRS predicted values against measured values of (a) soil P content extracted

using the Mehlich 3 method and analysed by colorimetry (M3P_Col); (b) soil P content

extracted using the Mehlich 3 method and analysed by ICP (M3P_ICP); (c) soil P content

extracted with water and analysed by colorimetry (Cp); (d) total soil P; (e) annual timothy

crop P-uptake, and; (f) annual P-budget. Based on validation statistics reported here and in

Table 3.2, NIRS predictions were considered moderately useful (MU) when

0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25, and less reliable (LR) when Rv

2 < 0.70 and

RPD < 1.75 (Malley et al., 2004).

Page 76: Caractérisation par spectroscopie et analyse compositionnelle des

54

Prédiction du phosphore organique du sol par la spectroscopie dans le

proche infra-rouge

Dans le chapitre 3, nous avons démontré que le phosphore extrait à la solution du

Mehlich-3; méthode de référence au Québec, et à l’eau ne sont pas prédictibles par la

spectroscopie dans le proche infra-rouge (SPIR). Le P total du sol, par contre, est

prédictible par cette technique. Nous avons présumé que la corrélation du P à la matière

organique du sol affecterait le potentiel de le prédire par la SPIR, et nous avons conclu que

les modèles de prédiction obtenus dans cette étude pour un sol sableux-limoneux

podzolique devraient être validés dans d’autres sites de textures différentes. De ce fait, nous

avons évalué, dans le chapitre 4, le potentiel de la SPIR à prédire le P organique, PT, PM3

et Al, Fe, Mg et Mn extraits selon la méthode Mehlich-3 pour des Mollisols loameux et

argileux-loameux riches en P organique étant donné qu’ils étaient sous semis direct.

Page 77: Caractérisation par spectroscopie et analyse compositionnelle des

55

CHAPITRE IV: PREDICTING SOIL ORGANIC PHOSPHORUS

USING NEAR-INFRARED REFLECTANCE SPECTROSCOPY

Abdi D.a,c, Cade-Menun B.J.c, Ziadi N.b, Tremblay G.F.b, and Parent L.É.c

aAgriculture and Agri-Food Canada, Soils and Crops Research and Development Centre,

2560 Hochelaga Boulevard, Québec, QC, Canada G1V 2J3.

bAgriculture and Agri-Food Canada, Semiarid Prairie Agricultural Research Centre, P.O.

Box 1030 Swift Current, SK, Canada, S9H 3X2.

cDepartment of Soils and Agri-Food Engineering, Université Laval, Québec, QC, Canada

G1K 7P4.

sera soumis à Geoderma

Page 78: Caractérisation par spectroscopie et analyse compositionnelle des

56

4.1 RÉSUMÉ

La mesure du phosphore organique (Po) du sol se fait jusqu’à date par différence

entre le phosphore total et le phosphore inorganique. L’objectif de cette étude est d’évaluer

le potentiel de la spectroscopie dans le proche infrarouge (SPIR) à prédire (i) le Po total, et

(ii) d’autres propriétés chimiques du sol [P total (PT)], matière organique (MO), et P, Al,

Fe, Ca, Mg et Mn extraits au Mehlich-3]. Des échantillons du sol (n = 360) ont été prelevés

d’un site expérimental sous semis direct à court et à long terme en Saskatechwan. Les

équations de calibration de la SPIR ont été développées avec 80% des échantillons alors

que 20% a été utilisé pour la validation. Les résultats ont démontré que les prédictions du

Po ont été acceptables pour l’ensemble des échantillons et pour le site sous semis direct à

long terme, et fiables pour le site sous semis direct à court terme. Les prédictions ont été

acceptables pour PM3, Fe, et Mg, fiables pour la MO, et non acceptables pour PT, Al et Mn

pour les deux sites. Cette étude démontre que la SPIR est une technique prometteuse pour

quantifier le Po dans les Mollisols.

Page 79: Caractérisation par spectroscopie et analyse compositionnelle des

57

4.2 ABSTRACT

To date, there is no direct method to quantify the total concentration of organic

phosphorus (OP) in soils. Near-infrared reflectance spectroscopy (NIRS) is a direct, rapid,

inexpensive, and accurate analysis technique for a wide variety of materials and it is

increasingly used in soil science. The aim of this study was to examine the potential of

NIRS to predict (i) total soil OP, and (ii) other soil chemical properties [total P (TP),

organic matter (OM), and Mehlich-3 extractable P, Al, Fe, Ca, Mg and Mn]. Soil samples

(n = 360) were taken from an experimental site near Indian Head, SK, Canada, from short-

term (8 yr, n = 180) and long-term (31 yr, n = 180) conservation tillage plots of a field pea ̶

spring wheat rotation receiving five P fertilizer rates annually. Samples were collected at

three soil depths (0-7.5, 7.5-15, and 15-30 cm). Calibration NIRS equations were developed

using 80% of the soil samples and the partial least squares regression while the remaining

20% of samples were used for validation. The predictive ability of NIRS was evaluated

using the coefficient of determination of validation (RV2) and the ratio of standard error of

prediction to standard deviation (RPD). Results show that NIRS predictions for total OP

were classified as moderately useful (0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25) for the total

soil sample set and for the long-term no-till set, and were moderately successful

(0.80 ≤ Rv2 < 0.90 and 2.25 ≤ RPD < 3.00) for the short-term no-till set. Predictions were

moderately useful for Mehlich-3 P, Fe, and Mg, and successful for OM, but were not

acceptable for TP, Al, and Mn. This study demonstrated that NIRS is a promising analysis

technique for OP.

Page 80: Caractérisation par spectroscopie et analyse compositionnelle des

58

4.3 INTRODUCTION

Phosphorus (P) occurs in inorganic and organic forms in soils. Inorganic P as

orthophosphate (HPO42- or H2PO4

-) is the directly available form for plant or microbial

uptake. Nevertheless, organic P (OP) is widely found in the natural environment and acts as

a resource of long-term bio-available P (Oehl et al., 2001). Despite this, its role, mobility

and bio-availability are poorly understood due to analytical difficulties. Indeed, total OP

cannot be estimated directly, and instead is determined indirectly as the difference between

total P (TP) and inorganic P (IP) by ignition (Saunders and Williams, 1955) or extraction

(Hedley et al., 1982; Tiessen and Moir, 2007). These procedures involve treating soil with

several chemical reagents and are time consuming, and there is potential for error in the

steps for TP and IP, increasing the potential for errors in OP measurement.

Near infrared reflectance spectroscopy (NIRS) is a rapid, inexpensive, accurate and

environmentally friendly technique. It is used to predict the concentration of the soil

attributes of interest from an empirical model developed based on complex spectra of a

subset of soil samples. Hence, it could be a good alternative to routine soil analysis

methods. The NIRS spectra are a result of the radiation absorbed by various chemical

bonds (e.g. C-H, C-C, and O-H) found in soil constituents (Miller, 2001). Using a single

spectrum, many soil attributes can be predictable.

Some soil constituents, such as organic matter (OM), carbon, nitrogen, and pH can

be successfully predicted by NIRS (Cozzolino and Morón, 2006; Brunet et al., 2007)

because they have a theoretical basis for NIRS prediction and are considered to be primary

properties (Chang et al., 2001). Others soil constituents related to these properties, such as

aluminium (Al), iron (Fe), P, potassium (K), magnesium (Mg), manganese (Mn), calcium

(Ca), and copper (Cu) may be predictable by NIRS (Chang et al., 2001). Abdi et al. (2012)

found that available P determined using Mehlich-3 and water extraction methods for

gravelly sandy soil samples were not accurately predicted by NIRS (R2 < 0.70 and RPD <

1.75). However, NIRS prediction was moderately useful for total P (0.70 ≤ R2 < 0.80 and

1.75 ≤ RPD < 2.25). They explained these results by the fact that total P contains a certain

proportion of OP that is related to OM (Turner et al., 2005). The NIRS prediction accuracy

can be also related to soil texture type (Chang et al., 2001). To our knowledge, the NIRS

prediction potential of soil OP has not previously been assessed.

Page 81: Caractérisation par spectroscopie et analyse compositionnelle des

59

The objectives of this study were to assess the potential for NIRS to predict (i) total

OP, and (ii) other chemical attributes: TP, OM, and Mehlich-3 extractable P, Al, Fe, Ca,

Mg, and Mn for loam and clayey loam soil samples.

4.4 MATERIALS AND METHODS

4.4.1 Experimental site description

The experimental site description is provided in detail in Lafond et al. (2011), and a

brief summary follows. The experiment was located approximately 19 km south-east of

Indian Head, Saskatchewan, Canada (50.42° N, 103.58° W), on short-term no-till (ST-NT;

8 yr) and long-term no-till (LT-NT; 31 yr) plots. The soil of the LT-NT field was loamy

with a pH of 6.8, while the soil of the ST-NT field was clayey loam with a pH of 7.3. The

soil type for both fields was Orthic Black Chernozem. A continuous cropping no-till system

has been established since 1978 on the LT-NT field, and since 2001 on the adjacent ST-NT

field. Prior to this, the ST-NT field has been managed using a fallow-crop system involving

extensive tillage to 10 cm soil depth. The experimental design was a split-plot with crop

rotation (pea and spring wheat) as main plots and five P treatments (0, 11, 22, 33, and 45 kg

P2O5 ha-1) as sub-plots. Both crops were present each year, and there were two replicates

for each combination of crop, tillage and fertilizer treatment (40 plots total). Soils from this

no-till experimental site were selected for this study due to the higher soil OP content.

4.4.2 Soil sampling and analysis

Soil samples were collected in the fall of 2008 (n = 120) and in the spring and the

fall of 2009 (n = 240) at three depths (0 - 7.5, 7.5 - 15.0, and 15.0 – 30.0 cm), air-dried,

sieved, and ground (< 2 mm). Total P and OP concentrations were determined for the

whole soil sample set (n = 360). Total P was extracted using the wet acid digestion method

of Parkinson and Allen (1975). Briefly, 0.5 g of soil was mixed with 3.75 mL of H2SO4 and

3 mL of digestion solution (175 mL H2O2 + 0.21 g Se + 7 g LiSO4.H2O), and digested at

360°C for 2.5 h. Total OP was determined by the difference between 0.5 M H2SO4

extractable P in a 0.5 g soil sample ignited at 550° C and an unignited sample according to

Saunders and Williams (1955). Concentrations of TP and OP were measured in the soil

samples extracts by the colorimetric molybdate blue method (Murphy and Riley, 1962).

Soil OM was determined by loss on ignition for soil samples (n = 90) collected in 2008

Page 82: Caractérisation par spectroscopie et analyse compositionnelle des

60

from LT-NT field where pea was grown, and from ST-NT fields where pea and spring

wheat were grown. Soil P available to plants and exchangeable Al, Fe, Ca, Mg, and Mn

were determined in samples collected in 2008 (n = 120) by ICP emission spectroscopy after

a Mehlich 3 extraction (Mehlich, 1984). Briefly, 2.5 g of air-dried soil was mixed with 25

mL of a Mehlich 3 solution (0.25 M NH4NO3 + 0.015 M NH4F + 0.001 M EDTA + 0.2 M

CH3COOH + 0.013 M HNO3 buffered at pH 2.3), shaken for 5 min, and then filtered

through Whatman no. 42 paper.

4.4.3 Near-infrared reflectance spectroscopy spectrum acquisition

The absorbance of soil samples [log (1/R), where R is the reflectance] was

measured in the visible and near-infrared regions between 400 and 2498 nm at 0.5 nm

intervals using a NIRSTM DS 2500 monochromator Instrument (Foss NIRSystems Inc.,

Silver Spring, MD), with the transport cup containing approximately 25 mL of soil sample.

Each spectrum was the average of 16 co-added scans. To take into account the operator

error, a repeatability file was created by collecting 20 spectra for one randomly selected soil

sample.

4.4.4 Spectral pre-treatment

Prior to calibration, different pre-treatments were selected to improve calibration

models, including the critical T-statistics values of 2.0 and 2.5 for T-outlier detection, and

the following math treatments: 1-16-16-1, 2-32-24-1, 1-20-20-1, 2-20-20-1, 1-40-40-1, 2-

40-40-1. These four numbers (i.e., 1-16-16-1) are derivative treatments, the first indicating

the order of the derivative, the second, the gap over which the derivative is was calculated,

the third, the number of the smoothing points, and the last, the second smooth. Scatter

correction with standard normal variate and detrending (SNVD) was applied to all spectra

using the WinISI IV software to reduce scatter and particle size effect, and to remove linear

or curvilinear trend of each spectrum (Barnes et al., 1989). Trying to find the best NIRS

prediction accuracy for TP and OP, calibration equations were developed separately for

LT-NT plus ST-NT, LT-NT, and ST-NT soil sample sets.

4.4.5 Calibration, cross-validation and validation

Calibration models of soil properties have been developed using the modified partial

least squares regression (PLSR) method of WinISI IV with 80% of each soil sample set

Page 83: Caractérisation par spectroscopie et analyse compositionnelle des

61

randomly selected. The remaining 20% of soil samples were used as external validation set.

Cross-validation was performed using four groups from the calibration set to avoid over-

fitting the calibration model (Nduwamungu et al., 2009a). To select the best calibration

model, two criteria were simultaneous used: the low standard error (SE) and high

coefficient of determination (R2) in all cross-validations and in validation sets (Van Vuuren

et al., 2006).

Validation of generated calibration models was performed using WinISI IV by

comparing predicted and reference data. Calibration accuracy, i.e., closeness to reference

data, was assessed as described in Abdi et al. (2012), based on the coefficient of

determination of validation (Rv2) and the ratio of standard error of prediction to standard

deviation (RPD), which is the standard deviation of samples in the validation set (SD)

divided by the standard error of prediction corrected for the bias (SEP(C)) [RPD =

SD/SEP(C)]. Calibration equations were considered to be excellent when Rv2 > 0.95 and

RPD > 4.00; successful when 0.90 ≤ Rv2 ≤ 0.95 and 3.00 ≤ RPD ≤ 4.00; moderately

successful when 0.80 ≤ Rv2 < 0.90 and 2.25 ≤ RPD < 3.00; moderately useful when

0.70 ≤ Rv2 < 0.80 and 1.75 ≤ RPD < 2.25; and less reliable when Rv

2 < 0.70 and

RPD < 1.75 (Malley et al., 2004). The coefficient of variation (CV) in the reference set was

defined as the SD divided by the mean of chemical values, whereas the coefficient of

variation in the calibration set was computed as the ratio of standard error of calibration

(SEC) to the mean of calibration data (Williams, 2001).

4.5 RESULTS AND DISCUSSION

4.5.1 Soil reference data

Descriptive statistics for soil reference properties are provided in Tables 4.1 and 4.2.

The OP concentration was relatively variable in the soil samples collected from LT-NT and

ST-NT fields (CV = 30.8%). This variability was lower in LT-NT experiment (CV =

26.5%) and higher under ST-NT management (CV = 34.2%). Soil total P content and

variability were lower in both fields and were comparable between the two no-till systems.

For the Mehlich-3 extracted soil nutrients, their coefficients of variation decreased in the

order of P > Ca > Al > Mg > Mn > Fe. Soil organic matter content was higher (> 5 %) and

relatively homogenous distributed in both experimental sites (22.8%).

Page 84: Caractérisation par spectroscopie et analyse compositionnelle des

62

4.5.3 Spectral pre-treatment, calibration, and prediction of soil organic P

The best calibration model for OP measured in the whole soil sample set (LT-NT and

ST-NT; n = 360) was developed with the math treatment of 2-20-20-1 and the critical T

value of 2.0 (Table 4.3). Thirty four soil samples were excluded from this model, indicating

that 88% of soil samples remained in the calibration set. Based on the relatively low

standard error of calibration (26 mg kg-1) and cross-validation (34 mg kg-1) sets, the

medium coefficient of variation (29%), and the high coefficients of determination of

calibration (Rc2 = 0.89) and cross-validation (1-VR = 0.82) equations, the performance of

NIRS calibration for OP was considered acceptable. The validation of the calibration model

also showed acceptable coefficient of determination (Rv2 = 0.75) and RPD (2.01), with a

moderately useful prediction of OP for these soil samples. This NIRS prediction

performance might be related to organic bonds such as C-C and C-H (Miller, 2001;

Viscarra Rossel et al., 2006) in organic forms of P.

The OP calibration models were generated with the math treatment of 1-20-20-1 and T

value of 2.5 for the LT-NT sample set and with the math treatment of 1-40-40-1 and T

value of 2.0 for the ST-NT sample set. The first model was based on 95% of the whole

sample set of calibration (n = 180), and the second one on 88% of the sample set free of

outliers. Near-infrared reflectance spectroscopy prediction of OP in LT-NT soil samples of

the validation set (n = 36) was moderately useful (Rv2 = 0.70, RPD = 1.81; Fig. 4.1).

However, the NIRS prediction of OP in the ST-NT soil sample of the validation set (n =

36) was moderately successful (Rv2 = 0.88, RPD = 2.49; Fig. 4.1). This improvement of

NIRS prediction’s accuracy could be attributed to the higher OP mean concentration and

coefficient of variation in the ST-NT compared to the LT-NT soil sample calibration set.

According to Dardenne et al. (2000), a soil constituent with a wide dispersion is likely more

easily predictable by NIRS.

4.5.4 Spectral pre-treatment, calibration, and prediction of soil total and Mehlich-3 P

The calibration equation generated to predict TP concentration for the whole set of soil

samples (n = 360, Table 4.3) showed low coefficients of determination (RC2 = 0.77) and

variation (CV = 14%), which resulted in low coefficients of determination in cross-

validation (1-VR = 0.68) and validation (Rv2 = 0.60) models, and low RPD (1.54). Hence,

Page 85: Caractérisation par spectroscopie et analyse compositionnelle des

63

the NIRS prediction performance was not acceptable for soil TP. The same result was

obtained for the NIRS prediction of TP concentration in LT-NT (Rv2 = 0.51, RPD = 1.34)

and in ST-NT (Rv2 = 0.44, RPD = 1.34) soil samples. Abdi et al. (2012) found a moderately

useful NIRS prediction accuracy (Rv2 = 0.75, RPD = 1.98) for TP in 192 gravely sandy

loam soils, while Bogrekci and Lee (2005) successfully predicted TP (Rv2 = 0.92) in 150

fine sandy soil samples using NIRS. Thus, NIRS prediction performance for TP may be

related to soil texture.

The R2 in calibration (0.83) and validation (0.72), and the RPD value of 1.86 for M3-P

(Table 4.4) indicated that the NIRS prediction was moderately useful. Conversely, M3-P

was previously found to be not predictable by NIRS in 192 gravely sandy loam soil

samples (Abdi et al., 2012) and in 150 clayey soil samples (Nduwamungu et al., 2009b). A

possible explanation for these contradictory results is the higher value of coefficient of

variation for M3-P found in our study (CV = 119%) compared to the low values of 48 and

61% respectively found in the studies of Abdi et al. (2012) and Nduwamungu et al.

(2009b).

4.5.5 Spectral pre-treatment, calibration, and prediction of soil organic matter and

Mehlich-3 nutrients

Statistics of calibration, cross-validation, and validation for total set of soil extracted

Mehlich-3 constituents and organic matter are listed in Table 4.4. Soil organic matter was

successfully predicted by NIRS (Rv2 = 0.91, RPD = 3.02). According to Chang et al.

(2001), organic matter is considered primary property and has a theoretical basis for NIRS

prediction, which was well documented in the literature (St-Luce et al., 2014, Stevens et al.,

2013; Brian and Daniel, 2012). Near-infrared reflectance spectroscopy prediction for Ca

extracted from the whole set of loam and clay loam soil samples was moderately successful

(Rv2 = 0.86, RPD = 2.25). Moderately useful NIRS predictions were found for Fe and Mg

(Rv2 = 0.78, RPD = 2.1). However, Al and Mn were poorly predicted (Rv

2 < 0.70;

RPD < 1.75). These NIRS predictions were lower than in the study by Abdi et al. (2012)

except for Fe which was more dispersed in this study. Conversely, the range of Mn was

smaller in our study. Thus, it appears that the range of reference data may affect the

performance of NIRS calibration model as suggested by Dardenne et al. (2000).

Page 86: Caractérisation par spectroscopie et analyse compositionnelle des

64

4.6 CONCLUSION

This work showed that NIRS prediction potential for total organic P determined for

360 loam and clayey loam soils samples was moderately useful. This prediction precision

was improved to moderately successful for OP determined in clay loam soil probably due

to higher content and coefficient of variation of reference data. The NIRS predictions were

considered reliable for soil Mehlich-3 extracted P, Fe, Ca, and Mg, and not acceptable for

TP, Al, and Mn. We conclude that NIRS can be a cost-effective, time-saving and accurate

alternative technique for soil OP. Further studies are needed to evaluate the potential of

NIRS for predicting the chemical forms of soil OP.

Page 87: Caractérisation par spectroscopie et analyse compositionnelle des

65

4.7 REFERENCES

Abdi, D., Tremblay, G.F., Ziadi, N., Bélanger, G., and Parent, L.-E., 2012. Predicting soil

phosphorus-related properties using near-infrared reflectance spectroscopy. Soil Sci.

Soc. Am. J. 76: 2318–2326.

Barnes, R.J., Dhanoa, M.S., and Lister, S.J., 1989. Standard normal variate transformation

and detrending of near-infrared diffuse reflectance spectra. Appl. Spectrosc. 43:

772-777.

Bogrekci, I., and W.S. Lee. 2005. Spectral phosphorus mapping using diffuse reflectance of

soils and grass. Biosystems. Eng. 91:305–312.

Brian, K. N. and Daniel, J.A. 2012. Near Infrared Reflectance-Based Tools for Predicting

Soil Chemical Properties of Oklahoma Grazinglands. Agr. J. 104: 1122-1129.

Brunet, D., Barthès, B.G., Chotte, J.-L. and Feller, C. 2007. Determination of carbon and

nitrogen contents in Alfisols, Oxisols and Ultisols from Africa and Brazil using NIRS

analysis: Effects of sample grinding and set heterogeneity. Geoderma, 139: 106–117.

Chang, C.W., D.A. Laird, M.J. Mausbach, and Hurburgh, C.R.Jr. 2001. Near-infrared

reflectance spectroscopy – Principal components regression analyses of soil

properties. Soil Sci. Soc. Am. J. 65: 480–490.

Cozzolino, D., and Morón, A. 2006. Potential of near-infrared reflectance spectroscopy and

chemometrics to predict soil organic carbon fractions. Soil Tillage Res. 85: 78–85.

Dardenne, P., Sinnaeve, G. and Baeten, V. 2000. Multivariate calibration and chemometrics

for near infrared spectroscopy: which method? J. Near Infrared Spectrosc. 8: 229-

237.

Lafond, G.P.,Walley, F., Maya, W.E., and Holzapfel, C.B. 2011. Long term impact of no-

till on soil properties and crop productivity on the Canadian prairies. Soil Till. Res.

117: 110–123.

Malley, D.F., Ben-Dor, E. and Martin, P.D. 2004. Application in analysis of soils. p. 729–

84. In Roberts, C.A., J. Jr. Workman, and J.B. Reeves III (ed.) Near-infrared

Spectroscopy in Agriculture. Am. Soc. Agr., Crop Sc. Soc. Am., and Soil Sc. Soc.

Am., Madison, USA.

Miller, C.E. 2001. Chemical principles of near-infrared technology. p. 19–8. In Williams,

P.C., and K.H. Norris (ed.) Near Infrared Technology in the Agricultural and Food

Industries 2nd ed. American Association of Cereal Chemists, St. Paul, MN.

Murphy, J., and Riley, J.P. 1962. A modified single solution method for the determination

of phosphate in natural waters. Anal. Chim. Acta 27: 31–36.

Nduwamungu, C., Ziadi, N., Parent, L.É., Tremblay, G.F., Thuriès, L., 2009a.

Opportunities for and limitations of near infrared reflectance spectroscopy

applications in soil analysis: A review. Can. J. Soil Sci. 89: 531-541.

Nduwamungu, C., Ziadi, N. Parent, L.-É. and Tremblay, G.F. 2009b. Mehlich 3 extractable

nutrients as determined by near-infrared reflectance spectroscopy. Can. J. Soil Sci.

89: 579–587.

Oehl, F., Oberson, A., Sinaj, S. and Frossard, E. 2001. Organic Phosphorus Mineralization

Studies Using Isotopic Dilution Techniques. Soil Sci. Soc. Am. J. 65: 780–787.

Parkinson, J.A, and Allen, S.E. 1975. A wet oxidation procedure suitable for the

determination of nitrogen and mineral nutrients in biological material. Commun.

Soil Sci. Plant Anal. 6: 1-11.

Page 88: Caractérisation par spectroscopie et analyse compositionnelle des

66

St. Luce M., Ziadi N., Zebarth B.J., Grant C.A., Tremblay G.F, and Gregorich E.G. 2014.

Rapid determination of soil organic matter quality indicators using visible near

infrared reflectance spectroscopy. Geoderma 232–234: 449–458.

Saunders, W.M.H. and Williams, E.G. 1955. Observations on the determination of total

organic phosphorus in soils. J. Soil Sci. 6: 254–267.

Stevens A, Nocita M, To´th G, Montanarella L, and van Wesemael B. 2013. Prediction of

soil organic carbon at the european scale by visible and near infrared reflectance

spectroscopy. PLoS ONE 8(6): e66409. doi:10.1371/journal.pone.0066409

Tiessen, H. and Moir, J.O. 2007. Characterization of available P by sequential extraction.

In Soil Sampling and Methods of analysis. Carter, M.R., and Gregorich, E.G., (2nd

eds). pp. 293–306.

Turner, B.L., B.J. Cade-Menun, L.M. Condron, and S. Newman. 2005. Extraction of soil

organic phosphorus. Talanta, 66: 294–306.

Van Vuuren, J. A. J., Meyer, J. H. and Claassens, A. S. 2006. Potential use of near infrared

reflectance monitoring in precision agriculture. Commun. Soil Sci. Plan. 37: 2171-

2184.

Viscarra Rossel, R.A., Walvoort, D.J.J., McBratney, A.B., Janik, L.J., and Skjemstad, J.O.,

2006. Visible, near infrared, mid infrared or combined diffuse reflectance

spectroscopy for simultaneous assessment of various soil properties. Geoderma,

131: 59-75.

Williams, P.C. 2001. Implementation of near-infrared spectroscopy. p. 145–169. In

Williams, P.C., and K. Norris (ed.) Near-infrared Technology in the Agricultural

and Food Industries. 2nd ed. American Association of Cereal Chemists, St. Paul,

MN.

Page 89: Caractérisation par spectroscopie et analyse compositionnelle des

67

Tableau 4-1 Descriptive statistics for the soil organic (OP) and total (TP) P analysed for long- and short-term no-till (NT)

treatments.

OP TP

Treatment Long- and short-

term NT Long-term NT Short-term NT

Long- and short-term

NT Long-term NT Short-term NT

N 360 180 180

359 180 179

mg kg-1

Min 16.2 72.6 16.2

353.7 353.7 408.2

Max 492.7 406.1 492.7

758.1 703.0 758.1

Mean 274.9 269.1 280.6

556.7 543.3 570.1

SD 84.7 71.4 96.0

76.6 79.9 70.8

%

CV (%) 30.8 26.5 34.2

13.8 14.7 12.4

Tableau 4-2 Descriptive statistics for the soil Mehlich-3 extracted nutrients and organic matter for long- and short-term no-till

(NT) treatments.

Treatment M3-P M3-Al M3-Fe M3-Ca M3-Mg M3-Mn

OM

N 120 120 120 120 120 120 90

mg kg-1 %

Min 0.0 0.0 19.0 1046.1 290.0 7.5 3.5

Max 65.7 736.0 157.4 28643.7 4639.9 97.0 9.0

Mean 11.4 254.8 76.5 6295.2 849.7 46.5 6.0

SD 13.5 236.7 27.7 7180.9 619.9 25.5 1.4

%

CV 119.0 92.9 36.2 114.1 73.0 54.7 22.8

Page 90: Caractérisation par spectroscopie et analyse compositionnelle des

68

Tableau 4-3 Statistics of near-infrared reflectance spectroscopy calibration, cross-validation, and validation for soil (OP) and

total (TP) P analysed for long- and short-term no-till (NT) treatments.

Calibration

Cross-

validation Validation

T Math treat. Nc T-outliers Mean SD SEC R2c CV (%)

SECV 1-VR

Nv Mean SD SEP(C) R2

v RPD Predi.

Long- and short-term NT

OP (mg kg-1) 2.0 2,20,20,1 254 34 272 79 26 0.89 29

34 0.82

72 287 79 39 0.75 2.01 MU

TP (mg kg-1) 2.5 1,16,16,1 279 9 556 77 37 0.77 14

43 0.68

72 556 73 47 0.60 1.54 NA

Long-term NT

OP (mg kg-1) 2.5 1,20,20,1 137 7 265 70 31 0.80 26

32 0.79

36 292 61 34 0.70 1.81 MU

TP (mg kg-1) 2.5 1,16,16,1 143 1 540 83 36 0.80 15

45 0.71

36 552 63 47 0.51 1.34 NA

Short-term NT

OP (mg kg-1) 2.0 1,40,40,1 127 17 283 91 37 0.83 32

39 0.82

36 292 96 38 0.88 2.49 MS

TP (mg kg-1) 2.5 1,16,16,1 137 6 565 67 35 0.72 12

46 0.52

36 587 74 55 0.44 1.34 NA

Page 91: Caractérisation par spectroscopie et analyse compositionnelle des

69

Tableau 4-4 Statistics of near-infrared reflectance spectroscopy calibration, cross-validation, and validation for soil Mehlich-3

extracted nutrients and organic matter for long- and short-term no-till treatments.

Calibration

Cross-

validation Validation

T Math treat. Nc T-outliers Mean SD SEC R2c CV (%)

SECV 1-VR

Nv Mean SD SEP(C) R2

v RPD Predi.

M3-P (mg kg-1) 2.5 1,16,16,1 74 22 13 12 5 0.83 92 6.25 0.73 22 11 14 7.5 0.72 1.86 MU

M3-Al (mg kg-1) 2.0 2,32,24,1 67 29 324 215 77 0.87 66 109 0.74 19 328 253 180 0.71 1.41 NA

M3-Fe (mg kg-1) 2.5 2,32,24,1 81 15 78 23 13 0.67 30 16 0.50 24 72 27 13 0.78 2.08 MU

M3-Ca (mg kg-1) 2.0 2,32,24,1 76 20 3771 3257 951 0.92 86 1522 0.78 24 8136 8585 3815 0.86 2.25 MS

M3-Mg (mg kg-1) 2.5 1,16,16,1 85 11 710 345 106 0.96 49 144 0.82 24 914 548 258 0.78 2.12 MU

M3-Mn (mg kg-1) 2.5 1,16,16,1 80 16 52 24 12 0.77 46 13 0.70 24 41 25 15 0.66 1.67 NA

OM (%) 2.5 1,16,16,1 55 41 6 1.45 0.31 0.95 24 0.38 0.92 18 6.25 1.33 0.44 0.91 3.02 S

Page 92: Caractérisation par spectroscopie et analyse compositionnelle des

70

NIR

S P

redic

ted v

alu

es f

or

soil

org

anic

P (

mg k

g-1

)

Measured values of soil organic P (mg kg-1)

Figure 4-1 Near-infrared reflectance spectroscopy (NIRS) predicted vs. measured values

of soil organic P analysed for (a) long- and short term no-till, (b) long-term no-till,

and (c) short-term no-till treatments.

Page 93: Caractérisation par spectroscopie et analyse compositionnelle des

71

ANALYSE STATISTIQUE NON-BIAISÉE DES FORMES DU PHOSPHORE DU

SOL DÉTERMINÉES PAR RMN-31P

Dans les chapitres 3 et 4, nous avons démontré que les analyses chimiques des

formes de P total, de P inorganique et organique du sol, peuvent être remplacées en partie

par la spectroscopie dans le proche infrarouge, une nouvelle technique analytique de chimie

verte. L’objectif du chapitre 5 est de démontrer l’utilité du nouveau concept mathémathique

de l’analyse compositionnelle dans la caractérisation des formes chimiques du P par rapport

aux approches classiques. Nous avons utilisé deux bases de données publiées de deux sites

expérimentaux canadiens différents (Île-du-Prince-Édouard et Québec). Une de ces bases

de données provient du chapitre 6 de cette thèse.

Page 94: Caractérisation par spectroscopie et analyse compositionnelle des

72

Page 95: Caractérisation par spectroscopie et analyse compositionnelle des

73

CHAPITRE V: UNBIASED STATISTICAL ANALYSIS OF SOIL 31P-

NMR

Dalel Abdia,c, J. Barbara Cade-Menunb, Noura Ziadia and Léon-Étienne Parentc

aAgriculture and Agri-Food Canada, Soils and Crops Research and Development Centre,

2560 Hochelaga Boulevard, Québec, QC, Canada G1V 2J3.

bAgriculture and Agri-Food Canada, Semiarid Prairie Agricultural Research Centre, P.O.

Box 1030 Swift Current, SK, Canada, S9H 3X2.

cDepartment of Soils and Agri-Food Engineering, Université Laval, Québec, QC, Canada

G1K 7P4.

Highlights

- Soil 31P-NMR forms are compositional data

- Ordinary log transformation generated statistically erroneous results depending on

measurement scale

- Compositional analysis using clr and ilr transformations avoids statistical analysis

biases

Soumis à Geoderma, 2014

Page 96: Caractérisation par spectroscopie et analyse compositionnelle des

74

5.1 RÉSUMÉ

Les formes du phosphore (P) déterminées par la spectroscopie de la résonance magnétique

nucléaire du 31P sont des données compositionnelles. L’objectif de cette étude était de

comparer les analyses statistiques classiques et la nouvelle approche de l’analyse

compositionnelle des espèces de P. Deux bases de données publiées ont été utilisées.

L’analyse de la variance et la corrélation avec le pH du sol ont été conduites pour les

espèces du P exprimées en pourcentage du P total ou en concentrations brutes, ainsi qu’à

leurs transformations logarithmiques simples ou compositionnelles. Les valeurs statistiques

de F de l’analyse de variance et les coefficients de corrélation obtenus pour les données

brutes et celles ordinairement transformées exprimées en pourcentage ou en concentration

sont contradictoires. Cependant, les résultats statistiques obtenus avec les transformations

compositionnelles sont conformes quel que soit l’échelle de mesure. L’analyse

compositionnelle permet d’obtenir une analyse non biaisée des formes de P dans le sol.

Mots clés : espèces de P de 31P-RMN, analyse compositionnelle, log-ratio centré, log-ratio

isométrique, analyse non biaisée

Page 97: Caractérisation par spectroscopie et analyse compositionnelle des

75

5.2 ABSTRACT

Phosphorus (P) forms determined by 31P nuclear magnetic resonance spectroscopy (31P–

NMR) are compositional data. Because compositional data are intrinsically related to each

other within a closed pre-defined compositional space, a simple log transformation,

variable by variable, or any other transformation of the compositional variables may

produce statistically erroneous results. However, most studies analyze the P forms as single

components rather than parts of some whole such as total P (TP) or soil dry mass, leading

systematically to methodological biases and possibly conflicting interpretations.

Compositional data analysis using centred log-ratio (clr) or isometric log-ratio (ilr)

coordinates avoids such difficulties and preserves sub-compositional coherence in the

analysis. The objective of this study was to compare classical and compositional methods

for the statistical analysis of 31P–NMR P data expressed as proportions of TP or

concentrations relative to soil dry mass. Two published datasets were used. Analyses of

variance and regression analysis with soil pH were conducted on P species percentages

scaled on TP or as raw concentrations scaled on a soil dry-weight basis as well as their

ordinary log, centred log-ratios (clr) and isometric log-ratios (ilr). Contradictory F-statistics

values and coefficients of correlation with soil pH were obtained for the raw and ordinary

log transformed 31P-NMR P data expressed as proportions or concentrations, indicating

spurious correlations. In contrast, statistical results were the same regardless of the

measurement unit when P compound percentages were clr-transformed. Using orthogonal

ilr coordinates, 31P-NMR P data were correlated to soil properties and to each other and

synthesized into a multivariate distance without methodological bias. We conclude that the

variance and regression analyses of molecular P species are scale-dependent and that the

clr- and the ilr-transformations should be used to unbiasedly analyze the P fractions and

avoid conflicting interpretations.

Key words: 31P NMR-P species, compositional analysis, centred log-ratio, isometric log-

ratio, unbiased analysis.

Page 98: Caractérisation par spectroscopie et analyse compositionnelle des

76

5.3 INTRODUCTION

Phosphorus turnover in soils is controlled by a combination of interrelated chemical and

biological factors. At the top of system hierarchy, the C, N, S and P cycles are inter-

connected (McGill and Cole, 1981; Stevenson, 1986). Understanding relationships among

P species within the P cycle can advance our knowledge of soil P bioavailability. Inorganic

phosphorus (Pi), specifically orthophosphate, is the primary source of P for most

organisms; however, organic phosphorus (Po) is generally more abundant. The Po

encompasses a large spectrum of ionic and molecular entities that can be identified using

solution 31P-NMR spectroscopy. Based on the nature of C-P bonds, Po species are

classified into phosphonate, phosphate monoesters and phosphate diesters (Condron et al.,

2005).

When P species are determined by advanced spectroscopic techniques such as 31P-NMR

and P K-edge X-ray absorption near-edge structure spectroscopy (P-XANES), the

proportions of each species are determined as relative percentages rather than absolute

concentrations (e.g. Liu et al., 2013). Although few studies using these techniques analyze

replicated results for full statistical analyses, most studies attempt some simple analyses,

such as correlation to other soil parameters. However, the relative percentages determined

by 31P-NMR and P-XANES, being compositional data, are not normally distributed

(Aitchison, 1986), and thus require transformation prior to statistical analyses. This is often

done using simple log transformations.

When the data from these methods are analyzed statistically, the P species are

commonly expressed as proportions of total P (TP) and are thus constrained between 0 and

100%. Confidence intervals that reach below 0% or above 100% are physically impossible.

Such data are intrinsically related to each other while the proportion of one P species can be

computed by difference between 100% and the sum of the proportions of other P species

(Aitchison, 1986). Thus, a change in the percentage of any one P form must affect the

percentage of at least one other P form. In addition, P species do not follow a normal

distribution and statistical analyses may return different results depending on measurement

scale due to spurious correlations between proportions across scales (Reimann and

Filzmoser, 2000).

Page 99: Caractérisation par spectroscopie et analyse compositionnelle des

77

Simple log transformation or any other transformation of compositional variables may

generate statistically erroneous results (Filzmoser et al., 2009). Alternatively, additive log-

ratio (alr) and centred log-ratio (clr) transformations allow the projection of compositional

data into an unconstrained real space of scale-invariant variables (Aitchison, 1986).

However, the alr has been criticized as being subjective, because the results and

interpretation of univariate analysis depend on the choice of denominator (Bacon-Shone,

2011). This arbitrariness can be avoided using clr where the geometric mean of all

proportions is selected as the denominator. The clr simplifies the interpretation of the

transformed variables because one could think in terms of the original variable (Filzmoser

and Hron, 2009). Nevertheless, the clr cannot be used in multivariate analysis because the

matrix is singular (the D clr variates add up to 0; Bacon-Shone, 2011). The isometric log-

ratio (ilr) transformation (Egozcue et al., 2003) overcomes this problem using a sequential

binary partition (SBP) of balances with orthonormal basis.

The objectives of this paper were to (i) demonstrate the dependence of variance and

correlation analyses of 31P-NMR P data on measurement scale, and (ii) statistically analyse

P species using clr and ilr transformations. We hypothesized that the statistical analysis of

P forms expressed as proportions or concentrations using the classical approach of log

transformation could lead to conflicitng results, which could be avoided using clr and ilr

transformations.

5.4 MATERIALS AND METHODS

5.4.1 Datasets

We conducted univariate and multivariate analyses on two Canadian datasets reporting

soil P species. Abdi et al. (2014) analyzed the long-term effect of no-till [NT] or

mouldboard plowing [MP]) and P fertilization (0 and 35 kg P ha-1) on P species distribution

in the soil profile. The trial was conducted with a corn-soybean rotation in Québec, Canada

since 1992. The soil is a clay loam (claey, mixed, mesic Typic Humaquept). Thirty six soil

samples were collected at three depths (0-5 cm, 5-10 cm and 10-20 cm) and analyzed by

solution 31P-NMR spectroscopy. For the second study, Cade-Menun et al. (2010) examined

the P forms in a long-term experiment that compared MP to NT systems in Prince Edward

Island, Canada. The experiment was established in 1985 on sandy loam soil (Orthic

Page 100: Caractérisation par spectroscopie et analyse compositionnelle des

78

Podzol). Soil was sampled at six depths (0-5 cm, 5-10 cm, 10-20 cm, 20-30 cm, 30-40 cm

and 40-60 cm) and was also analyzed for P species.

5.4.1.1 Compositional data transformations

5.4.1.1.1 Centred log-ratio transformation

The ith clr is computed as follows (Aitchison, 1986):

(1)

where is the ith component and is the geometric mean across proportions.

The Euclidean distance between any two compositions, called the Aitchison distance, is

computed as follows:

(2)

where is the ith clr of the reference composition.

5.4.1.1.2 Isometric log-ratio transformation

The ilr is the log ratio between geometric means of two non-overlapping subsets of parts

(tagged with + and - signs) called orthonormal balances and computed as follows (Egozcue

et al., 2003):

(3)

where ilri is the ith balance between two sub-compositions, i Є [1, D-1], and are

number of components at numerator and denominator, respectively, and are

geometric means of components in subsets. Similarly to clr, ilr can scan the real space (±∞)

and is scale-invariant (the ratio between components of sets of components is a way to

eliminate the unit or scale of measurement). An advantage of ilr over clr is that the ilr

transformations have matrix rank of D-1. As a result, the ilr values associated with pre-

defined sub-compositions do not change if composition includes more sub-compositions.

Page 101: Caractérisation par spectroscopie et analyse compositionnelle des

79

The choice of non-overlapping subsets to compute geometric means is formalized by a

sequential binary partition (SBP). Orthogonal coefficients are computed as

allowing ilrs to be geometric coordinates in the Euclidean space.

The Aitchison distance A between two equal-length compositions x (diagnosed) and y

(reference or control) is computed as follows (Egozcue and Pawlowsky-Glahn, 2006):

(4)

where I is identity matrix and T indicates a transposed matrix.

The Mahalanobis distance M between two equal-length compositions x (diagnosed)

and y (reference or control) is computed as follows:

(5)

where COV is the covariance matrix.

5.4.1.1.3 Choice of SBP

Inorganic P (Pi) is a mineral product or the product of Po hydrolysis that can be

converted back to Po forms by soil microorganisms (Baldwin et al., 2005). The Pi and Po

species can be connected to each other using a mobile-fulcrums-buckets design balancing

subsets of components a priori defined in a SBP (Parent et al., 2012). The SBP can be

generated by default (Comas-Cufí and Thió-Henestrosa, 2011) or arranged hierarchically

according to some theory or practice (Egozcue and Pawlowsky-Glahn, 2006, Parent et al.,

2012). However, the choice of SBP does not influence the multivariate distance due to

balance orthogonality. The selected SBPs are presented in Tables 5.1 and 5.2 for processing

the Abdi et al. (2014) and the Cade-Menun et al. (2010) datasets, respectively.

The P compositional vector can be closed to total P or soil dry matter. In any event, the

first balances contrast the P fractions with filling values to the unit or scale of

measurement. If P fractions are scaled on soil dry matter, a filling value is computed

between the unit of measurement (e.g. kg of dry soil) and the sum of P fractions. If P

fractions are scaled on total P, a filling value is computed as residual P by difference

Page 102: Caractérisation par spectroscopie et analyse compositionnelle des

80

between total P and P fractions; residual P may also be quantified and P fractions added up

to total P. The filling value is needed to convert ilr means back to P fractions expressed on

a familiar scale or unit such as proportions. Thereafter, we contrasted Pi and Po forms.

Inorganic P is further partitioned into orthophoosphate and other Pi forms (pyro- and

polyphosphates) and the Po arranged into in subsets according to their bioavailability,

sorption and hydrolysis.

5.4.1.1.4 Ordinary logarithmic transformation

A log ratio is a log contrast or difference between two ordinary logarithmic

transformations. The ordinary logarithmic transformation returns an Euclidean distance that

differs from clr if geometric means differ between compositions (hence rejecting the ceteris

paribus assumption) as follows, where the star refers to some benchmark composition:

(6)

The ordinary logarithmic transformation returns a Mahalanobis distance that differs

from the ilr if the denominator differs between two compositions. As per example for a

dual /Po ratio, the computation is as follows after discarding the orthogonal coefficient

for simplification:

(7)

Where the difference between Pi in two samples is the difference between Pi

concentrations only if Po concentrations are the same (ceteris paribus assumption).

Similarly, difference between geometric means at numerator of two compositions is the

difference between geometric means at numerator only if geometric means at denominator

remains the same (ceteris paribus assumption). Equations 6 and 7 show that the ceteris

paribus equation is not applicable to compositional data. The Aitchison or Mahalanobis

distances across ordinary log-transformed proportions are thus inflated by compositional

discrepancies at denominator.

Page 103: Caractérisation par spectroscopie et analyse compositionnelle des

81

5.4.1.2 Statistical analysis

The 31P-NMR P forms were expressed as percentages of TP or as raw concentrations on

a soil dry-weight basis to show the influence of measurement scale on the results of

statistical analyses. For comparison, analyses of variance (ANOVA) were conducted using

clr-transformed data. Soil pH (0.01 M CaCl2) was correlated with P species percentages of

TP or as raw concentrations on a soil dry-weight basis as well as ordinary log [log (n+1)]

and isometric log-ratios to show spurious correlations. Compositional and statistical

analyses were conducted using R “compositions” (van den Boogaart et al., 2011). Rounded

zeroes for data below the detection limit for 31P-NMR were replaced according to Martín-

Fernández et al. (2003). The effects of tillage system, P fertilization and soil depth, and

their interactions were tested using Proc Mixed of SAS (SAS Institute, 2001).

5.5 RESULTS AND DISCUSSION

5.5.1 Biased analysis of variance

As a result of scale dependency and data redundancy, ANOVA may return conflicting

values for the significance of treatment effects and treatment means. Tables 5.3 and 5.4

show that treatment effects as well as their interactions differed considerably in terms of

significance whether the ANOVA was conducted using ordinary logarithmic

transformation of raw concentrations of P species or their proportions relative to total soil P

for both the Abdi et al. (2014) and Cade-Menun et al. (2010) studies.

In the Abdi et al. (2014) study, tillage, TP and FD effects were not significant

whatever the variable tested (Table 3). Among P, D, TD and TPD sources of variation,

there were 13 significant effects (P < 0.05) using log-transformed concentrations of P

species (mg P kg-1 dry soil), seven significant effects (P < 0.05) using log-transformed

proportions of P species (vs. Total P), and seven significant effects (P < 0.05) using centred

log-ratios. The interpretation of treatment effects thus differed considerably depending on

the transformation being used. For example, there was a significant effect of soil depth on

pyrophosphate concentration, but no such effect using log- and clr-transformed

pyrophosphate values. There were significant TD and TPD interactions using log-

Page 104: Caractérisation par spectroscopie et analyse compositionnelle des

82

transformed concentrations and proportions but no such significant using clr-transformed

orthophosphate values.

In the Cade-Menun et al. (2010) study (Table 5.4), there were five significant effects of

tillage treatments (P < 0.05) at five soil depths using log-transformed concentrations

compared with two significant effects (P < 0.05) using log-transformed proportions, and

five significant effects (P < 0.05) using clr-transformed values. Moreover, significant

effects were not always on the same P species at the same soil depth. For example, tillage

treatment significantly influenced (P < 0.05) the level of AMP in the 10-20 cm layer, of

myo-inositol hexakisphosphate in the 20-30 cm layer, and of -glycerophosphate in the 40-

60 cm layer, but these effects were not significant using ordinary log transformations.

Similar to the results in Tables 5.3 and 5.4, treatment means were also dependent on

the data transformation technique (Table 5.5). Moreover, the ordinary log back-transformed

treatment means were similar to the raw means (Table 5.6). In contrast, clr back-

transformed P species differed from the raw data indicating bias in raw and ordinary log-

transformed data.

Many studies investigating P cycling in soils in various environments collect replicate

field samples and conduct replicate analysis of various soil chemical parameters such as

pH, but do not conduct 31P-NMR or P-XANES analyses on these replicate samples, either

selecting a single sample or compositing the replicate samples into one sample for these

advanced analyses (e.g. Turner, 2008; Sato et al., 2009; Redel et al., 2011; Cheesman et al.,

2012; Liu et al., 2013; Wei et al., 2014). Other research groups use composite samples for

all aspects of their studies of P cycling (e.g. Turner and Engelbrecht, 2011; Vincent et al.,

2013; Hashimoto and Watanabe, 2014). The two data sets used here (Cade-Menun et al.,

2010; Abdi et al., 2014) are among the very few 31P-NMR studies to use individual samples

from replicate field plots for 31P-NMR to assess changes in soil forms with management

practices. In our opinion, all studies should try where ever possible to conduct 31P-NMR or

P-XANES analyses on replicate samples rather than composite samples, to allow the results

to be statistically analyzed. This will make the studies more scientifically rigorous, rather

than merely descriptive. However, the results from this current study clearly show that care

Page 105: Caractérisation par spectroscopie et analyse compositionnelle des

83

must be taken when transforming these results. We recommend using clr-transformed data

for statistical analysis of P forms determined by 31P-NMR or P-XANES.

5.5.2 Spurious correlations

It has long been recognized (Pearson, 1897) that the same measurement made on

different scales generates spurious correlations. More recently, Aitchison (1986) showed

that redundancy generates at least one negative correlation because any increase in one

proportion must be associated with a decrease in one or more proportions in a closed

system. When measurement scales vary from raw concentrations on dry soil basis and raw

proportions of total P to their log-transformed values, scale-dependent coefficients of

correlation can be measured by the discrepancies in correlation coefficient, sign and

significance (Table 5.7). Discrepancies are shown when correlating soil pH with (1) log-

transformed concentrations and proportions of residual P in the Abdi et al. (2014) study and

(2) orthophosphate concentrations and proportions and their log-transformed expressions in

the Cade-Menun et al. (2010) study. The pH is significantly correlated either negatively or

positively depending on measurement scale. This is why a scale-invariant expression that

avoids redundancy is required to conduct correlation analysis of compositional data.

These results raise concerns about the many studies that have correlated 31P-NMR and

P-XANES results to soil parameters such as pH with little or no data transformation (e.g.

Cheesman et al., 2012; Turner and Blackwell, 2013; Wei et al., 2014). Any relationships of

soil P forms to soil chemical parameters developed in these studies should be treated with

caution. We recommend that the authors of these and other similar studies reanalyze their

correlations after correctly transforming their data.

Because multivariate analyses are based on correlations between variables, using raw

concentrations or proportions or their log-transformation must distort the results of

multivariate analysis (Aitchison, 1986). The degree of distortion in multivariate distances

such as the Mahalanobis distance can be shown by an inflation of the Mahalanobis distance

using raw concentrations or proportions or their log-transformation (Fig. 5.1). As shown by

Eqs. 6 and 7, the distortion is attributable to the difference in the geometric means between

compositions. An unbiased multivariate distance can be obtained using ilr variables

because the ilr variables are orthogonal to each other in a Euclidean space, hence

Page 106: Caractérisation par spectroscopie et analyse compositionnelle des

84

measuring differences between compositions as straight lines. The ilr variables are the most

appropriate transformation to conduct multivariate analyses on compositional data

(Filzmoser et al., 2009). The ilr variables can be back-transformed into familiar units to

facilitate interpreting the results, solving for the D-1 ilr values and the sum of components

to the unit of measurement.

5.6 CONCLUSIONS

In this paper, we demonstrated that variance and correlation analyses of 31P-NMR P

data depended on measurement scale, which is to be expected for compositional data. The P

species could be statistically analyzed using clr transformations to facilitate interpreting

ANOVA results and ilr transformations for correlation and multivariate analyses. The clr

and ilr transformations avoid methodological biases in classical procedures that often lead

to conflicting interpretations using the same compositional vector but different scaling

procedures. The ilr transformation has considerable advantage over other data

transformation methods because P species can be arranged into interpretable orthonormal

balances according to some theory that connects the P species measured in the P cycle to

components of the C, N, and S cycles. There is a need for paradigm shift in studies on

elemental cycles like P considering that components of any cycle are connected to each

other in a coherent system analysis.

Page 107: Caractérisation par spectroscopie et analyse compositionnelle des

85

5.7 REFERENCES

Abdi, D., Cade-Menun, B.J., Ziadi, N., and Parent, L-É., 2014. Long-term impact of tillage

and P fertilization on soil P forms as deterimed by 31P-NMR spectroscopy. J.

Environ. Qual. doi:10.2134/jeq2013.10.0424.

Aitchison, J., 1986. The statistical analysis of compositional data, first ed. London, UK.

Bacon-Shone, J., 2011. A short history of compositional data analysis, in: Pawlowsky-

Glahn, V., Buccianti, A. (Eds.), Compositional data analysis: Theory and

Applications. John Wiley & Sons, New York. pp. 3-11.

Baldwin, D.S, Howitt. J.A., and Beattie, J.K., 2005. Abiotic degradation of organic

phosphorus compounds in the environment, in: Turner, B.J., Frossard, L., Baldwin,

D. (Eds.), Organic Phosphorus in the Environment. CABI Publishing, Oxfordshire,

pp. 75-88.

Cade-Menun, B.J., Carter, M.R., James, D.C., and Liu, C.W., 2010. Phosphorus forms and

chemistry in the soil profile underlong-term conservation tillage: A phosphorus-31

nuclear magnetic resonance study. J. Environ. Qual. 39: 1647-1656.

Cade-Menun, B.J., and Preston, C.M., 1996. A comparison of soil extraction procedures for 31P NMR spectroscopy. Soil Sci. 161: 770-785.

Cheesman, A.W., Turner, B.L., and Reddy, K.R., 2012. Soil phosphorus forms along a

strong nutrient gradient in a tropical ombrotrophic wetland. Soil Sci. Soc. Am. J.

76: 1496-1506.

Comas-Cufí, M., Thió-Henestrosa, S., 2011. and CoDaPack 2.0: a stand-alone, multi-

platform compositional software. Proceedings 4th International Workshop on

Compositional Data Analysis, Sant Feliu de Guíxols, Spain, 9-13 May 2011

(congress.cimne.com/codawork11/Admin/Files/FilePaper/p28.pdf).

Condron, L.M., Turner, B.L., and Cade-Menun, B.J., 2005. Chemistry and dynamics of soil

organic phosphorus. in: Sims, J.T. Sharpley, A.N. (Eds.), Phosphorus, Agriculture

and the Environment. Monograph no 46, Soil Science Society of America,

Madison, WI. pp. 87-121.

Egozcue, J.J., and Pawlowsky-Glahn, V., 2006. Simplicial geometry for compositional

data, in: Buccianti, A., Mateu-Figueras, G., Pawlowsky-Glahn, V. (Eds.),

Compositional Data Analysis in the Geosciences: From Theory to Practice. Special

Publications, 264, Geological Society, London. pp. 67–77.

Egozcue, J.J., and Pawlowsky-Glahn, V., Mateu-Figueras, G., Barceló-Vidal, C., 2003.

Isometric log-ratio transformations for compositional data analysis. Math. Geol. 35:

279-300.

Filzmoser, P., and Hron, K., 2009. Correlation analysis for compositional data. Math.

Geosc. 41: 905-919.

Filzmoser, P., Hron, K., and Reimann, C., 2009. Univariate statistical analysis of

environmental (compositional) data: Problems and possibilities. Sci. T. Environ.

407: 6100-6108.

Page 108: Caractérisation par spectroscopie et analyse compositionnelle des

86

Hashimoto, Y., and Watanabe, Y., 2014. Combined applications of chemical fractionation,

solution 31P-NMR and P K-edge XANES to determine phosphorus speciation in

soils formed on serpentine landscapes. Geoderma, 230-231, 143-150.

Liu, J., Yang, J., Cade-Menun, B.J., Liang, X., Hu, Y., Liu, C.W., Zhao, Y., Li, L., and Shi,

Y. 2013. Complementary phosphorus speciation in agricultural soils by sequential

fractionation, solution 31P nuclear magnetic resonance, and phosphorus K-edge X-

ray absorption near-edge structure spectroscopy. J. Environ. Qual. 42: 1763-1770.

Martin-Fernandez, J.A., Barcelo-Vidal, C., and Pawlowsky-Glahn, V., 2003. Dealing with

zeros and missing values in compositional data sets using nonparametric imputation.

Math. Geol. 35: 253–278.

McGill, W.B., and Cole, C.V., 1981. Comparative aspects of cycling of organic C, N, S and

P through soil organic matter. Geoderma 26, 267-286.

Parent, S.- É., Parent, L. E. Rozane, D. E. Hernandes, A., and Natale, W., 2012. Nutrient

balance as paradigm of plant and soil chemometrics, in: Issaka, R.N. (Eds.), Soil

Fertility. InTech Publishing, New York, pp. 83-114.

http://www.intechopen.com/books/soil-fertility.

Pearson, K., 1897. Mathematical contributions to the theory of evolution. On a form of

spurious correlation which may arise when indices are used in the measurement of

organs. Proceedings of the Royal Society of London, LX, pp. 489-502.

Redel, Y.D., Escudey, M., Alvear, M., Conrad, J., and Borie, F., 2011. Effects of tillage

and crop rotation on chemical phosphorus forms and some related biological

activities ina Chilean Ultisol. Soil Use Manage. 27: 221-228.

Reimann, C., and Filzmoser, P., 2000. Normal and lognormal data distribution in

geochemistry: death of a myth. Consequences for the statistical treatment of

geochemical and environmental data. Environ. Geol. 39: 1001-1014.

SAS Institute. 2001. The SAS system for Windows. Release 8.2. SAS Inst., Cary, NC.

Sato, S., Neves, E.G., Solmon, D., Liang, B., and Lehmann, J., 2009. Biogenic calcium

phosphate transformation in soils over millennial time scales. J. Soils Sediments, 9:

194-205.

Stevenson, F.J., 1986. Cycles of soils. Carbon, nitrogen, phosphorus, micronutrients, first

ed.Wiley-Interscience, New York.

Turner, B.L., 2008. Soil organic phosphorus in tropical forests: an assessment of the

NaOH-EDTA extraction procedure for quantitative analysis by solution 31P NMR

spectroscopy. Eur. J. Soil Sci. 59: 453-466.

Turner, B.L., and Blackwell, M.S.A., 2013. Isolating the influence of pH on the amounts

and forms of soil organic phosphorus. Eur. J. Soil Sci. 64: 249-259.

Turner, B.L., and Engelbrecht, B.M.J., 2011. Soil organic phosphorus in lowland tropical

rain forests. Biogeochem. 103: 297-315.

Turner, B.L., Cade-Menun, B.J., Condron, L.M., and Newman, S., 2005. Extraction of soil

organic phosphorus. Talanta, 66: 294-306.

Page 109: Caractérisation par spectroscopie et analyse compositionnelle des

87

van den Boogaart, K.G., Tolosana-Delgado, R., and Bren, M., 2011. Compositions:

Compositional data analysis, R Package Version 1.10 2.

http://cran.rproject.org/web/packages/compositions/compositions.pdf

Vincent, A.G., Vestergren, J., Gröbner, G., Persson, P., Schleucher, J., and Giesler, R.,

2013. Soil organic phosphorus transformations in a boreal forest chronosequence.

Plant Soil. 367: 149-162.

Wei, K., Chen, Z.H., Zhang, X.P., Liang, W.J., and Chen, L.J., 2014. Tillage effects on

phosphorus composition and phosphatase activities in soil aggregates. Geoderma,

217-218, 37-44.

Page 110: Caractérisation par spectroscopie et analyse compositionnelle des

88

Table 5-1 Sequential binary partition of soil 31P-NMR P species analyzed by Abdi et al. (2014).

ilr Orthoa Pyro Poly Phos Myo Neo Scyllo Gluc α-glyc β-glyc Nucl Chol DNA Res

P

r s

1 1 1 1 1 1 1 1 1 1 1 1 1 1 -1 13 1

2 1 1 1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 0 3 10

3 1 -1 -1 0 0 0 0 0 0 0 0 0 0 0 1 2

4 0 1 -1 0 0 0 0 0 0 0 0 0 0 0 1 1

5 0 0 0 1 -1 -1 -1 -1 -1 -1 -1 -1 -1 0 1 9

6 0 0 0 0 1 1 1 1 1 1 1 1 -1 0 8 1

7 0 0 0 0 1 -1 -1 -1 -1 -1 -1 -1 0 0 1 7

8 0 0 0 0 0 1 -1 -1 -1 -1 -1 -1 0 0 1 6

9 0 0 0 0 0 0 1 -1 -1 -1 -1 -1 0 0 1 5

10 0 0 0 0 0 0 0 1 -1 -1 -1 -1 0 0 1 4

11 0 0 0 0 0 0 0 0 1 -1 -1 -1 0 0 1 3

12 0 0 0 0 0 0 0 0 0 1 -1 -1 0 0 1 2

13 0 0 0 0 0 0 0 0 0 0 1 -1 0 0 1 1 aOrthophosphate (Ortho), pyrophosphate (Pyro), polyphosphate (Poly), phosphonate (Phos), myo-inositol hexakisphosphate (Myo),

neo-inositol hexakisphosphate (Neo), scyllo-inositol hexakisphosphate (Scyllo), glucose-6 phosphate (Gluc), α-glycerophosphate (α-

Glyc), β-glycerophosphate (β-Glyc), nucleotides (Nucl), choline-phosphate (Chol), deoxyribonucleic acid (DNA), residual

phosphate (Res P).

r: number of components at numerator

s: number of components at denominator

Page 111: Caractérisation par spectroscopie et analyse compositionnelle des

89

Table 5-2 Sequential binary partition of soil 31P-NMR P species analyzed by Cade-Menun et al. (2010).

ilr Orthoa Pyro Poly Phos Myo Scyllo Gluc α-glyc AMP Mono1 Mono2 Mono3 Oth.Di DNA r s

1 1 1 1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 3 11

2 1 -1 -1 0 0 0 0 0 0 0 0 0 0 0 1 2

3 0 1 -1 0 0 0 0 0 0 0 0 0 0 0 1 1

4 0 0 0 1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 1 10

5 0 0 0 0 1 1 1 1 1 1 1 1 -1 -1 8 2

6 0 0 0 0 1 1 -1 -1 -1 -1 -1 -1 0 0 2 6

7 0 0 0 0 1 -1 0 0 0 0 0 0 0 0 1 1

8 0 0 0 0 0 0 1 1 1 -1 -1 -1 0 0 3 3

9 0 0 0 0 0 0 1 -1 -1 0 0 0 0 0 1 2

10 0 0 0 0 0 0 0 1 -1 0 0 0 0 0 1 1

11 0 0 0 0 0 0 0 0 0 1 -1 -1 0 0 1 2

12 0 0 0 0 0 0 0 0 0 0 1 -1 0 0 1 1

13 0 0 0 0 0 0 0 0 0 0 0 0 1 -1 1 1 aOrthophosphate (Ortho), pyrophosphate (Pyro), polyphosphate (Poly), phosphonate (Phos), myo-inositol hexakisphosphate (Myo),

scyllo-inositol hexakisphosphate (Scyllo), glucose-1 phosphate (Gluc), α-glycerophosphate (α-Glyc), adenosine-5-monophosphate

(AMP), orthophosphate monoesters regions 1, 2, or 3 (mono1, mono2, mono3), orthophosphate diesters other than DNA (Oth.Di),

deoxyribonucleic acid (DNA)

r: number of components at numerator

s: number of components at denominator

Page 112: Caractérisation par spectroscopie et analyse compositionnelle des

90

Table 5-3 ANOVA of the effect of tillage (T), P fertilization (P) and soil depth (D) on log-

and clr-transformed P compositions (Abdi et al., 2014). P species defined in Table 1.

P species Tillage (T) Fertilization

(P)

Depth

(D)

T x

P

T x

D

F x

D

T x P

x D

log (mg P kg-1 sol)

Orthophosphate ns * ** ns * † **

Pyrophosphate ns ns * † * ns ns

Polyphosphate ns ns ns ns ns ns ns

Phosphonate ns ** ns ns ns ns ns

Myo-IP6 ns ns ns ns ns ns *

Neo-IP6 ns ns † ns ns ns ns

Scyllo-IP6 ns ns * ns † ns ns

Glucose-6 phosphate ns ns ns ns ns ns ns

α-Glycerophosphate ns ns ns ns ns ns ns

β-Glycerophosphate ns ns ns ns ns ns ns

Nucleotides ns ns * ns * ns ns

Choline-phosphate ns † ns ns ns ns ns

DNA ns ns ns ns ** ns *

Residual phosphate † ns ns ns ns ns ns

log (% total P)

Orthophosphate ns † ** ns * ns *

Pyrophosphate ns ns ns ns ns ns ns

Polyphosphate ns ns ns ns ns ns ns

Phosphonate ns ns ns ns ns ns ns

Myo-IP6 ns ns ns ns ns ns ns

Neo-IP6 ns ns ns ns ns ns ns

Scyllo-IP6 ns ns ** ns † ns †

Glucose-6 phosphate ns ns ns ns ns ns ns

α-Glycerophosphate ns † † ns ns ns ns

β-Glycerophosphate ns † † ns ns ns ns

Nucleotides ns ns † ns * ns ns

Choline-phosphate ns * ns ns ns † ns

DNA ns ns † ns * ns †

Residual phosphate ns † ns ns ns ns ns

centred log-ratio transformation

Orthophosphate ns * * ns ns ns †

Pyrophosphate ns ns † † † ns ns

Polyphosphate ns ns ns ns ns ns ns

Phosphonate ns * ns ns ns ns ns

Page 113: Caractérisation par spectroscopie et analyse compositionnelle des

91

Myo-IP6 ns ns ns ns ns ns ns

Neo-IP6 ns ns ns ns ns ns ns

Scyllo-IP6 ns ns * ns † ns ns

Glucose-6 phosphate ns ns ns ns ns ns ns

α-Glycerophosphate ns † † ns ns ns ns

β-Glycerophosphate ns † † ns ns ns ns

Choline-phosphate ns † ns ns ns ns ns

Nucleotides ns ns * ns ** ns ns

DNA ns ns ns ns ** ns ns

Residual phosphate ns ns ns ns ns ns ns

† Significant at P < 0.1

* Significant at P < 0.05

** Significant at P < 0.01

ns, nonsignificant at the 0.10 level

Page 114: Caractérisation par spectroscopie et analyse compositionnelle des

92

Table 5-4 ANOVA of the effect of tillage and soil depth on log- and clr-transformed P

compositions (Cade-Menun et al., 2010). P species defined in Table 2.

P species 0-5 cm 5-10 cm 10-20 cm 20-30 cm 30-40 cm 40-60 cm

log (mg P kg-1 soil)

Orthophosphate ns * † ns ns ns ns

Pyrophosphate ns ns ns ns ns ns ns

Polyphsophate ns ns † ns ns ns ns

Phosphonate ns * ns ns ns ns ns

Myo ns * † † ns ns ns

Scyllo ns ns ns ns ns ns ns

Glucose-1 P ns ns ns ns ns ns ns

α-GlyceroP ns ns ns † ns ns ns

AMP ns * ns ns ns ns ns

Mono1 ns ns ns ns ns ns ns

Mono2 ns ns ns ns ns ns ns

Mono3 ns ns ns ns ns ns ns

OthDdi. ns * ns ns ns ns ns

DNA ns ns ns ns ns ns ns

log (% total P)

Orthophosphate ns ns ns ns ns ns ns

Pyrophosphate ns ns ns ns ns ns ns

Polyphsophate ns ns ns ns ns ns ns

Phosphonate ns ns ns ns ns ns ns

Myo ns * * † ns ns ns

Scyllo ns ns ns ns ns ns ns

Glucose-1 P ns ns ns ns ns ns ns

α-GlyceroP ns ns ns ns ns ns ns

AMP ns ns † ns ns ns ns

Mono1 ns ns ns ns ns ns ns

Mono2 ns ns ns † ns ns ns

Mono3 ns ns ns ns ns ns ns

Oth.Di. ns ns ns ns ns ns ns

DNA ns ns ns ns ns ns ns

centred log-ratio transformation

Orthophosphate ns ns ns ns ns ns ns

Pyrophosphate ns ns ns ns ns ns ns

Polyphsophate ns ns † ns ns ns ns

Phosphonate ns ns ns ns ns ns ns

Myo ns † † * ns ns ns

Scyllo ns ns ns ns ns ns ns

Glucose-1 P ns ns ns ns ns ns ns

Page 115: Caractérisation par spectroscopie et analyse compositionnelle des

93

α-GlyceroP ns ns ns ns ns *

AMP ns ns * ns ns ns

Mono1 ns ns † ns * ns

Mono2 ns ns ns † † ns

Mono3 ns ns ns ns ns ns

Oth.Di. ns * ns ns ns ns

DNA ns ns ns ns † †

† Significant at P < 0.1

* Significant at P < 0.05

** Significant at P < 0.01

ns, nonsignificant at the 0.10 level

Page 116: Caractérisation par spectroscopie et analyse compositionnelle des

94

Table 5-5 Treatment means of P species concentrations and proportions for main

effect of tillage in the Abdi et al. (2014) and Cade-Menun et al. (2010) studies.

P species Transformation

mg P kg-1

soil

log(mg P kg-1

soil) clr

% total

P

log(% total

P) clr

mg P kg-1 soil % total P

Abdi et al. (2014) study

Orthophosphate 260.2 2.4 2.5 41.8 1.6 2.5

Pyrophosphate 6.2 0.9 -1.3 1.0 0.3 -1.3

Polyphosphate 5.7 0.8 -1.6 0.7 0.2 -1.6

Phosphonate 17.6 1.3 -0.3 2.7 0.6 -0.3

Myo-IP6 61.6 1.8 1.0 9.9 1.0 1.0

Neo-IP6 24.5 1.4 0.1 4.0 0.7 0.1

Scyllo-IP6 24.4 1.4 0.1 3.9 0.7 0.1 Glucose-6

phosphate 14.9 1.2 -0.4 2.3 0.5 -0.4

α-Glycerophosphate 10.7 1.1 -0.8 1.6 0.4 -0.8

β-Glycerophosphate 21.4 1.3 -0.1 3.2 0.6 -0.1

Nucleotides 36.5 1.6 0.4 5.5 0.8 0.4

Choline-phosphate 13.5 1.2 -0.5 2.1 0.5 -0.5

DNA 10.7 1.1 -0.7 1.7 0.4 -0.7

Residual phosphate 118.8 2.1 1.7 19.5 1.3 1.7 Cade-Menun et al. (2010) study

Orthophosphate 503.4 2.7 3.6 68.9 1.8 3.6

Pyrophosphate 8.0 1.0 -0.7 1.0 0.3 -0.7

Polyphsophate 4.9 0.8 -0.9 0.7 0.2 -0.9

Phosphonate 5.7 0.8 -0.9 0.8 0.3 -0.9

Myo 98.1 2.0 1.9 13.0 1.1 1.9

Scyllo 19.5 1.3 0.3 2.6 0.6 0.3 Glucose-1

phosphate 6.1 0.9 -0.9 0.8 0.3 -0.9

α-Glycerophosphate 14.1 1.2 -0.1 1.8 0.4 -0.1

AMP 11.1 1.1 -0.3 1.4 0.4 -0.3

Mono1 5.7 0.8 -0.9 0.8 0.3 -0.9

Mono2 42.5 1.6 1.0 5.5 0.8 1.0

Mono3 5.1 0.8 -1.0 0.7 0.2 -1.0

Oth.Di. 5.6 0.8 -0.9 0.7 0.2 -0.9

DNA 9.9 1.0 -0.4 1.3 0.4 -0.4

Page 117: Caractérisation par spectroscopie et analyse compositionnelle des

95

Table 5-6 Back-transformed treatment means of P species concentrations and

proportions for main effect of tillage in the Abdi et al. (2014) and Cade-Menun et al.

(2010) studies.

P species Back-transformation

mg P kg-1

soil

log(mg P kg-1

soil) clr

% total

P

log(% total

P) Clr

mg P kg-1 soil % total P

Abdi et al. (2014) study

Orthophosphate 259.3 259.3 262.1 41.4 41.4 41.8

Pyrophosphate 6.6 6.6 6.0 1.0 1.0 1.0

Polyphosphate 5.7 5.7 4.5 0.9 0.9 0.7

Phosphonate 17.8 17.8 17.2 2.8 2.8 2.7

Myo-IP6 61.3 61.3 62.2 9.8 9.8 9.9

Neo-IP6 24.5 24.5 25.0 3.9 3.9 4.0

Scyllo-IP6 24.4 24.4 24.1 3.9 3.9 3.8 Glucose-6

phosphate 14.8 14.8 14.7 2.4 2.4 2.3

α-Glycerophosphate 10.4 10.4 10.1 1.7 1.7 1.6

β-Glycerophosphate 20.8 20.8 20.3 3.3 3.3 3.2

Nucleotides 35.9 35.9 34.3 5.7 5.7 5.5

Choline-phosphate 13.8 13.8 13.1 2.2 2.2 2.1

DNA 11.1 11.1 10.9 1.8 1.8 1.7

Residual phosphate 121.5 121.5 122.4 19.4 19.4 19.5

Cade-Menun et al. (2010) study

Orthophosphate 503.4 503.4 515.0 68.9 68.9 69.6

Pyrophosphate 8.0 8.0 7.1 1.0 1.0 1.0

Polyphsophate 4.9 4.9 5.6 0.7 0.7 0.8

Phosphonate 5.7 5.7 5.7 0.8 0.8 0.8

Myo 98.1 98.1 95.3 13.0 13.0 12.9

Scyllo 19.5 19.5 18.4 2.6 2.6 2.5 Glucose-1

phosphate 6.1 6.1 5.8 0.8 0.8 0.8

α-Glycerophosphate 14.1 14.1 13.1 1.8 1.8 1.8

AMP 11.1 11.1 10.2 1.4 1.4 1.4

Mono1 5.7 5.7 5.6 0.8 0.8 0.8

Mono2 42.5 42.5 37.5 5.5 5.5 5.1

Mono3 5.1 5.1 5.3 0.7 0.7 0.7

Oth.Di. 5.6 5.6 5.5 0.7 0.7 0.7

DNA 9.9 9.9 9.5 1.3 1.3 1.3

Page 118: Caractérisation par spectroscopie et analyse compositionnelle des

96

Table 5-7 Correlation between soil pH and P species expressed as raw or log-

transformed concentrations or proportions or as isometric log ratios (ilr). P species and

ilrs defined in Tables 1 (Abdi et al., 2014) and 2 (Cade-Menun et al., 2010).

P species mg P kg-1

soil

% total P Log (mg P kg-1

soil)

Log (% total

P)

Ilr

r r

Abdi et al. (2014) study

Orthophosphate -0.26 0.37† -0.32† 0.33* ilr1 -0.49**

Pyrophosphate 0.28 0.49† 0.32† 0.54* ilr2 0.53**

Polyphosphate -0.19 -0.01 -0.06 0.07 ilr3 -0.25

Phosphonate -0.44† -0.05 -0.43* -0.07 ilr4 0.17

Myo-IP6 -0.68† -0.37† -0.69* -0.36* ilr5 0.14

Neo-IP6 -0.64† -0.04 -0.64* -0.04 ilr6 -0.77**

Scyllo-IP6 -0.85† -0.45† -0.82* -0.49* ilr7 0.41*

Glucose-6 phosphate -0.67† -0.31* -0.65* -0.34* ilr8 0.62**

α-Glycerophosphate -0.79† -0.42† -0.83* -0.45* ilr9 0.05

β-Glycerophosphate -0.76† -0.64† -0.78* -0.66* ilr10 0.28

Nucleotides -0.43† -0.10 -0.48* -0.13 ilr11 0.38*

Choline-phosphate -0.91† -0.74† -0.89* -0.74* ilr12 -0.11

DNA 0.06 0.49† 0.05 0.51* ilr13 -0.54**

Residual phosphate -0.35† 0.39† -0.38* 0.38*

Cade-Menun et al. (2010) study

Orthophosphate 0.54* -0.66** 0.53** -0.65** ilr1 -0.28

Pyrophosphate 0.58* 0.42* 0.57** 0.42* ilr2 -0.59**

Polyphsophate 0.37 0.05 0.15 -0.02 ilr3 0.26

Phosphonate 0.68** 0.22 0.64** 0.22 ilr4 -0.27

Myo 0.72** 0.46* 0.67** 0.50* ilr5 0.36†

Scyllo 0.78** 0.64** 0.71** 0.66** ilr6 0.24

Glucose-1 phosphate 0.41† 0.09 0.47* 0.09 ilr7 -0.38†

α-Glycerophosphate 0.57* 0.39† 0.56** 0.42* ilr8 0.32

AMP 0.79** 0.70** 0.75** 0.73** ilr9 -0.51*

Mono1 0.43† -0.30 0.37† -0.30 ilr10 -0.49*

Mono2 0.85** 0.58** 0.71** 0.62** ilr11 -0.60**

Mono3 0.70** 0.64** 0.69** 0.65** ilr12 0.63**

Oth.Di. 0.65** 0.28 0.62** 0.29 ilr13 -0.06

DNA 0.56* 0.24 0.52* 0.24

† Significant at P < 0.1

* Significant at P < 0.05

** Significant at P < 0.01

Page 119: Caractérisation par spectroscopie et analyse compositionnelle des

97

Figure 5-1 Relationship between Mahalanobis distance from ilr with (a) ordinary log

transformed 31P NMR-P species concentration, and (b) raw of 31P NMR-P species

concentrations (data from Abdi et al., 2014).

Page 120: Caractérisation par spectroscopie et analyse compositionnelle des

98

Effet à long-terme du travail du sol et de la fertilisation phosphatée sur

les formes de P déterminées par RMN-31P

Le chapitre 5 a démontré l’importance du traitement des formes du P par la nouvelle

approche mathématique de l’analyse compositionnelle qui permet de générer des résultats

non biaisés, et par conséquent, des interprétations cohérentes. Nous présentons, dans

l’annexe, un modèle de balance que nous avons développé en utilisant les coordonnées du

log ratio isométrique pour décrire les relations entre des pools de P. Nous avons utilisé des

données publiées pour des Mollisols canadiens cultivés. Ce travail a été présenté lors d’un

congrès international sur l’analyse des données compositionnelles. Dans le chapitre 6, nous

avons utilisé la spectroscopie de la résonance magnétique nucléaire du 31P pour

charactériser les espèces ioniques et moléculaires du P, et étudié leur distribution dans le

profil du sol perturbé par des pratiques culturales en utilisant l’analyse compositionnelle.

Page 121: Caractérisation par spectroscopie et analyse compositionnelle des

99

CHAPITRE VI: LONG-TERM IMPACT OF TILLAGE

PRACTICES AND P FERTILIZATION ON SOIL P FORMS AS

DETERMINED BY 31P NUCLEAR MAGNETIC RESONANCE

SPECTROSCOPY

Dalel Abdi, Barbara J. Cade-Menun*, Noura Ziadi and Léon-Étienne Parent

D. Abdi and N. Ziadi, Agriculture and Agri-Food Canada, Soils and Crops Research and

Development Centre, 2560 Hochelaga Boulevard, Québec, QC, Canada G1V 2J3; B. J.

Cade-Menun, Agriculture and Agri-Food Canada, Semiarid Prairie Agricultural Research

Centre, P.O. Box 1030 Swift Current, SK, Canada, S9H 3X2; D. Abdi and L.É. Parent,

Department of Soils and Agri-Food Engineering, Université Laval, Québec, QC, Canada

G1K 7P4.

Journal of Environmental Quality, 2014. 43 (4): 1431-1441.

Page 122: Caractérisation par spectroscopie et analyse compositionnelle des

100

6.1 RÉSUMÉ

Les pratiques de conservation du sol sont de plus en plus utilisées pour réduire

l’érosion, améliorer la capacité de rétention de l’eau et la teneur en matière organique.

L’objectif de cette étude était d’évaluer l’effet du travail du sol (labour concentionel; MP,

et semis direct, NT) à long-terme et la fertilisation phosphatée (0 et 35 kg P ha-1) sur la

distribution des espèces de P dans le profil du sol. Les échantillons de sol ont été prelevés

d’un site expérimental établi au Québec, Canada, sur une rotation maïs-soya dans trois

profondeurs (0-5, 5-10, et 10-20 cm). Les résultats des analyses chimiques des échantillons

du sol ont montré que le PM3 et les orthophosphates s’accumulaient à la surface du sol

fertilisé sous semis direct, alors que les formes organiques du P (monoesters et nucléotides)

s’accumulaient en profondeur du sol non labouré. Nous avons conclu que le semis direct et

la fertilisation phosphatée changeaient la distribution des formes de P tout au long du profil

du sol et pourrait augmenter le risque de perte de P.

Page 123: Caractérisation par spectroscopie et analyse compositionnelle des

101

6.2 ABSTRACT

Conservation tillage practices have become increasingly common in recent years

to reduce soil erosion, improve water conservation, and increase soil organic matter.

However, research suggests that conservation tillage can stratify soil test phosphorus (P),

but little is known about the effects on soil organic P. This study was conducted to assess

the long term effects of tillage practices (no-till [NT] and mouldboard plowing, [MP]),

and P fertilization (0 and 35 kg P ha-1) on the distribution of P species in the soil profile.

Soil samples from a long term corn-soybean rotation experiment in Québec, Canada, were

collected from three depths (0-5, 5-10, and 10-20 cm). These were analysed for total P

(TP), total C (TC), total N (TN), pH, and Mehlich-3 P (PM3); P forms were characterized

with solution phosphorus-31 nuclear magnetic resonance spectroscopy (31P-NMR).

Results showed a stratification of TP, TC, TN, pH, PM3 and Mehlich-3 extractable

aluminum (Al) and magnesium (Mg) under NT management. The PM3 and

orthophosphate concentrations were greater at the soil surface (0 – 5 cm) of the NT-P35

treatment. Organic P forms (orthophosphate monoesters scyllo-IP6 and nucleotides) had

accumulated in the deep layer of NT treatment possibly due to preferential movement. We

found evidence that the NT system and P fertilization changed the distribution of P forms

along the soil profile, potentially increasing soluble inorganic P loss in surface runoff and

organic P in drainage, and decreasing bioavailability of both inorganic and organic P in

deeper soil layers.

Abbreviations: α-Glyc, α-glycerophosphate; β-Glyc, β-glycerophosphate; Chol-P, choline

phosphate; clr, centered log ratio; Gluc-6P, glucose-6-phosphate; IP6, inositol

hexakisphosphate; MP, mouldboard plow; Nucl, nucleotides; NT, no-till; Ortho,

orthophosphate; PM3, soil P content extracted using the Mehlich-3 method; P0, soil

treatment with 0 kg P ha−1; P35, soil treatment with 35 kg P ha−1; Phos, phosphonate; Poly,

polyphosphate; Pyro, pyrophosphate; Res P, residual P; TC, total carbon; TN, total

nitrogen; TP, total P.

Page 124: Caractérisation par spectroscopie et analyse compositionnelle des

102

6.3 INTRODUCTION

Phosphorus is one of the essential and commonly limiting macronutrients for plant

growth, but is a major cause of freshwater eutrophication (Elser et al., 2007). Conservation

tillage practices (minimum or no-till; NT) are characterized by minimal soil disturbance and

mixing, and have been used more frequently in recent years to reduce off-site losses of

nutrients associated with eroded particles, including P (Chichester and Richardson, 1992;

Shepard, 2005). Some benefits of NT over conventional tillage include reduced wind and

water erosion (Dick, 1992; Olson and Ebelhar, 2009) and greater soil biological activity

(Duiker and Beegle, 2006). Further advantages of NT for crop growth and yields (Lafond et

al., 2011) and economic performance (Holm et al., 2006) are well recognized.

However, by maintaining crop residues and fertilizers on the soil surface and

reducing their mixing into the plow layer, the relatively immobile nutrients that do not

readily move down the soil profile will remain at or near the soil surface (Sharpley, 2003;

Sharpley and Smith, 1994). Therefore, NT management systems often result in high

concentrations of nutrients at the soil surface (0 -5 cm) but sharply decreasing

concentrations below this depth (Selles et al., 1999). Studies have shown that NT

management has induced the stratification of soil organic carbon ( Poirier et al., 2009;

Zibilske et al., 2002), nitrate (Grant and Lafond, 1994; Lupwayi et al., 2006), potassium

(Fernández et al., 2008) and extractable P ( Sharpley, 2003; Sharpley and Smith, 1994;

Zibilske et al., 2002).

Stratification of P is of particular environmental and agronomic concern. Indeed,

high concentrations at the soil surface potentially increase the loss of dissolved P in runoff

(Cade-Menun et al., 2013; Kleinman et al., 2009), which is readily available to aquatic

organisms, while low concentrations at depth may limit plant P uptake by decreasing

available P in the rooting zone (Lupwayi et al., 2006). Dissolved P could be lost from soil to

water through drains (Haygarth et al., 1998), where dissolved organic P can represent a large

fraction (McDowell and Koopmans, 2006) and could be used by algae, especially when low

dissolved inorganic P concentrations limit growth (Whitton et al., 1991). Therefore, it is

important to characterize P stratification and to determine its possible causes. This can allow

Page 125: Caractérisation par spectroscopie et analyse compositionnelle des

103

producers to apply fertilizer appropriately for plant requirements while minimizing the

potential for P loss.

Phosphorus stratification may be due to the minimal mixing of surface-applied

fertilizers under NT management (Sharpley, 2003), and/or due to leaching of P from crop

residues that are retained on the soil surface (Hussain et al., 1999). To indicate the P source,

Cade-Menun et al. (2010) identified specific inorganic and organic P forms and

characterized their distribution patterns with soil depth in a long-term tillage study, using

solution 31P nuclear magnetic resonance (31P -NMR) spectroscopy. To date, few studies have

investigated the effects of tillage systems (Redel et al., 2011) and P fertilization (Ahlgren et

al., 2013) on the stratification of 31P-NMR P forms in soil profiles, and their interaction

effect has never been studied.

Solution 31P-NMR spectroscopy is by far the most widely used spectroscopic

technique for the speciation of soil organic P, because it provides the most detailed and

accurate information in most circumstances (Doolette and Smernik, 2011). However, few

31P-NMR studies to date have used replicate field samples, allowing for statistical analysis

of P forms. Caution must be used during statistical analysis of 31P-NMR data, because this

technique produces results that are not normally distributed and must be transformed prior to

statistical analysis. The 31P–NMR P forms are generally computed as proportions of total P

and hence are constrained between 0 and 100%. Therefore, following Aitchison (1986) and

Egozcue et al. (2003), P compounds can be defined as compositional data, i.e. strictly

positive data characterized by a constant sum and considered to be parts of a whole that only

provides relative information. Because compositional data are intrinsically related to each

other and have logistic-normal distribution, they should be expressed in terms of log ratios

(Aitchison, 1986). A simple log transformation, variable by variable, or any other

transformation of the single compositional variables may provide statistically erroneous

results, and thus may lead to contradictory interpretations (Filzmoser et al., 2009).

Compositional data analysis using log ratio transformation can avoid these difficulties.

The objectives of this study were: (i) to use 31P-NMR spectroscopy to identify the P

forms along the soil profile under a long-term tillage and P fertilization study in Québec,

Canada; (ii) to use compositional analysis to examine the effect of tillage systems and P

Page 126: Caractérisation par spectroscopie et analyse compositionnelle des

104

fertilization on their distribution patterns with depth; and (iii) to determine the possible causes

and risks for the stratification of soil P forms.

6.4 MATERIELS AND METHODS

6.4.1 Experimental site

The long-term crop rotation experiment was established in 1992 at the l’Acadie

Research Station (45°18’N; 73°21’W), Agriculture and Agri-Food Canada. The soil is a clay

loam (clayey, mixed, mesic Typic Humaquept) with 364 g kg-1 clay and 204 g kg-1 sand in the

Ap horizon. This deep soil originates from a fluvial deposit and evolved from a fine-textured,

greyish-to-brown parent material. The site is tile-drained with slope less than 1% and was

cropped with alfalfa (Medicago sativa) before 1992. Corn (Zea mays L.) was grown in 1992 to

1994, followed by soybean (Glycine max L.) in 1995; after that, corn and soybean were grown

in an annual rotation until 2010.

The chemical characteristics of the topsoil when the experiment was established were,

on average: soil organic matter (SOM) 38 g kg-1, PM3 135 kg ha-1, Mehlich-3 saturation ratio

(P/Al) 4.3%, and pH 6.3 (1:2 soil/water) (Légère et al., 2008; Tremblay et al., 2003). The mean

annual temperature in the area of the study is 6.3°C, and the mean total annual precipitation is

1100 mm (Poirier et al., 2009).

The experimental set-up was a split plot with two tillage practices (NT and MP)

assigned to main plots and nine fertilization combinations consisting of three N (0, 80 and 160

kg N ha-1) and three P (0, 17.5 and 35 kg P ha-1) applications to subplots. Experimental

treatments were replicated in four blocks, with individual plots measuring 25 m long and 4.5 m

wide. The MP treatment consisted of one moldboard plowing operation in the fall after harvest

to a depth of 20 cm, followed by disking and harrowing to 10 cm each spring before seeding.

For the NT treatment, plots had previously been ridge-tilled from 1992 to 1997 and were flat

direct-seeded from 1998 onward. For direct seeding, crop residues were left on the ground after

harvest. There were six rows per subplot unit, and corn and soybean were sown at rates of 74 x

103 and 45 x 104 plants ha-1, respectively. Mineral fertilizers (N and P) were band-applied (5

cm from the seeding row) only during the corn phase of the rotation (11 years) using a disk

opener (3 – 4 cm deep), according to local recommendations. The P treatments were applied in

a single application at planting as triple super-phosphate (0-46-0). Nitrogen treatments were

Page 127: Caractérisation par spectroscopie et analyse compositionnelle des

105

first band-applied at seeding at rates of 0, 48 and 48 as urea, and completed with 0, 32, and 112

kg N ha-1 side-dressed as ammonium nitrate at approximately the eight-leaf stage.

6.4.2 Soil sampling and chemical analysis

Although the experiment was conducted with four replicate blocks per treatment, only

three blocks with subplots receiving 160 kg N ha-1 and 0 or 35 kg P ha-1 were selected for this

study. Soil profiles were sampled to depths of 0-5 cm, 5-10 cm and 10-20 cm during fall 2010

for a total of 36 samples (2 tillage x 2 P-fertilization x 3 replicates x 3 depths). Samples were

air-dried and sieved to < 2 mm. Soil pH was measured in distilled water with 1:2 soil to

solution ratio (Hendershot et al., 2008). Soils were extracted by shaking 2.5 g of soil with 25 ml

of Mehlich-3 solution (pH 2.3) for 5 min (Mehlich, 1984) and the concentrations of P, Al, iron

(Fe), calcium (Ca), and magnesium (Mg) were determined with an Inductively Coupled

Plasma Optical Emission Spectrometer (ICP-OES; Model 4300DV, Perkin Elmer, Shelton,

CT). Total soil P was determined as described in Nelson (1987). Briefly, 0.1 g of finely ground

soil (0.2 mm) was mixed in a 50-mL boiling flask with 0.5 g K2S2O8 and 10 mL 0.9 M H2SO4,

and then digested at 121◦C in an autoclave for 90 min. The solution was analyzed by the

ammonium molybdate-ascorbic acid method (Murphy and Riley, 1962). Total C and N were

determined by dry combustion on 0.20 mm ground soils with a LECO CNS-1000 analyzer

(LECO Corp., St. Joseph, MI).

6.4.3 Solution 31P-NMR spectroscopy

Samples were analyzed by solution 31P-NMR spectroscopy using a modified version

of the Cade-Menun and Preston (1996) procedure. This involved shaking 2.5 g of soil with

25 ml of combined 0.25 mol L-1 NaOH and 0.05 mol L-1 Na2EDTA solutions for 6 h,

followed by centrifugation for 20 min at approximately 1500 x g. A 1-mL aliquot was

removed and diluted to 10 ml with deionized water for determination of TP, Al, Fe, Ca, and

manganese (Mn) by ICP-OES. The remaining supernatants were frozen and freeze-dried.

Freeze-dried samples were re-dissolved in 1.0 mL D2O, 0.6 mL of 10 mol L-1 NaOH, and

0.6 mL of the NaOH-EDTA extracting solution. The samples were centrifuged (1500 x g)

for 20 min to remove particles > 0.1 µm in diameter and transferred to 10 mm NMR tubes.

Solution 31P-NMR spectra were acquired on a 600 MHz spectrometer (202.5 MHz for P;

INOVA; Varian, Palo Alto, CA) equipped with a 10 mm broadband probe. The NMR

Page 128: Caractérisation par spectroscopie et analyse compositionnelle des

106

experimental parameters were: pulse width 18 µs (90o), acquisition 0.675 s, delay time 4.32

s; 2200-4300 scans (4-6 h); no proton decoupling. Although spin-lattice relaxation (T1)

times were not measured for these samples, this delay time was estimated to be sufficient

based on the ratio of P/(Fe+Mn) in the extracts (McDowell et al., 2006; Cade-Menun and

Liu, 2013).

Chemical shifts of signals were determined in parts per million (ppm) relative to an

external orthophosphoric acid standard (85%), and the orthophosphate peak was

standardized to 6 ppm for each sample. Signals were assigned to P compounds based on the

literature (Cade-Menun, 2005; Cade-Menun et al., 2010; Turner et al., 2012). Peak areas

were calculated by integration on spectra processed with 1 Hz and 7 Hz line broadening,

using NMR Utility Transform Software (NUTS, Acorn NMR, Livermore CA, 2000 edition).

6.4.4 Compositional data analysis

Soil chemical properties, excluding pH, were closed to total mass of soil and

represented as pertaining to a sample space called the simplex SD which is defined as

follows:

SD = {TP + Al + Fe + Ca + Mg + TC + TN + Res = 100%} [1]

where Res represent the residual mass in the soil. Total P comprised PM3 and its

complementary value to total P. To conduct statistical analyses in this study, we used the

centred log-ratio (clr) from Eq. [2] to transform, and back-transform, the soil chemical

properties (excluding pH), and the P forms computed as relative percentages of TP or as

raw concentrations on a dry-weight basis using the R ‘compositions’ package (van den

Boogaart et al., 2011). The clr transformation computes the geometric mean (g(x)) across

components ( ix ) as follows:

[2]

The clr transformation simplifies the interpretation of the transformed variables

because one could think in terms of the original variables of the simplex (Filzmoser and

Hron, 2009).

)(ln)(

xg

xxclr i

i

Page 129: Caractérisation par spectroscopie et analyse compositionnelle des

107

The R ‘robCompositions’ package was used to replace the rounded zeros for

concentrations of P failing to be detected (values below detection limit of the 31P-NMR), and

the whole composition was adjusted accordingly (Martín-Fernández et al., 2003).

Confidence intervals were computed about clrs means then back-transformed to familiar

measurement units using the R ‘compositions’ package (van den Boogaart et al., 2011).

6.4.5 Statistical analysis

Data were tested for normality using the SAS univariate procedure (SAS Institute,

2001). Analyses of variance (ANOVA) for centred log-ratio transformed 31P-NMR P

forms, PM3, TP, Al, Fe, Ca, Mg, TC and TN, and pH were conducted using Proc Mixed of

SAS (SAS Institute, 2001) to test the effects of tillage, P fertilization, depth, and their

interactions. Least Significant Difference (LSD) was computed to separate treatments

means when ANOVA test was significant (p < 0.05) or considered as a trend (0.05 < p <

0.1).

6.5 RESULTS

6.5.1 Chemical soil properties

The statistical ANOVA results for chemical soil properties are shown in Table 1.

There was a significant interaction of tillage, fertilization, and depth for TP (Table 1, Fig. 1).

The LSD results showed significant differences for the amount of TP between the top soil

layer (0-5 cm) and the subjacent layers (5-10 cm and 10-20 cm) under NT-P0, and between

the 0 to10 cm and 10 to 20 cm soil layers in the NT-P35 treatment. Similarly, a significant

tillage x P fertilization x depth interaction (p < 0.05) was obtained for PM3 (Table 1). The

highest concentrations for PM3 were recorded under NT in P fertilized treatment at 0 to 5 cm

(Fig. 1). Concentrations of Al and Mg were affected by tillage x soil depth interaction (Fig. 2).

However, Fe was affected only by the depth variation, and was significantly lower at 0-5 cm

(212 ± 14mg kg-1) than at the deeper layers (5-10 cm, 226± 14 mg kg-1; 10-20 cm, 229± 17

mg kg-1). There were no significant treatment effects for Ca (Table 1), which was 2432 ± 131

mg kg-1 under MP and 2768 ± 244 mg kg-1 under NT. Significant differences in TC and TN

were detected in NT treatment among the soil depths, where they were higher in the top 5 cm

soil layer (Fig. 2). Soil pH was significantly higher in the NT treatment at 0 to 5 cm (6.33 ±

Page 130: Caractérisation par spectroscopie et analyse compositionnelle des

108

0.14; p < 0.01) compared to 5 to 10 cm (5.95 ± 0.25) and 10 to 20 cm (5.96 ± 0.32). None of

the soils parameters tested, except PM3, showed any effect of P fertilization.

6.5.2 Identification of phosphorus forms by 31P nuclear magnetic resonance

spectroscopy

The 31P-NMR peaks for P compounds detected in this study fall between 25 and -25

ppm (Table 2). An example spectrum from MP-P35 treatment at 0 to 5 cm depth is shown in

Fig. 3. Three groups of inorganic P forms were detected: orthophosphate at 6.00 ± 0.01 ppm

chemical shift, pyrophosphate at -4.02 ± 0.02 ppm, and polyphosphates between -4.23 and -

24.76 ppm, with the polyphosphate end group detected at -3.98 ± 0.05 ppm.

Organic P compound classes detected by solution 31P-NMR included phosphonates

from 25 to 7.80 ppm, orthophosphates monoesters at 7.70 to 6.20 ppm and at 5.78 to 3.5

ppm, and orthophosphates diesters between 2.20 and -3.40 ppm (Cade-Menun, 2005; Cade-

Menun et al., 2010). An example spectrum of the monoester region from MP-P35 treatment

at 0 to 5 cm depth is shown in Fig. 4. For all treatments and depths, the 31P-NMR spectra

indicated that orthophosphate monoesters were dominated by stereoisomers of inositol

hexakisphosphate (IP6). Three of these (myo, neo, and scyllo) were detected in all

treatments, whereas D-chiro-IP6 was detected in NT-P0 in one block at 0 to 5 cm depth, and

in NT-P35 at the three depths. The most abundant stereoisomer of IP6, myo-IP6, gave four

characteristics signals in the ratio 1:2:2:1 at 5.48 (± 0.02), 4.51 (± 0.02), 4.11 (± 0.02), and

4.02 (± 0.02) ppm (Turner et al., 2003; Cade-Menun, 2005). The identification of these

peaks as myo-IP6 was confirmed by rerunning two samples after spiking with phytic acid

(McDowell et al., 2007). Two signals at 6.41 (± 0.01) ppm and 4.27 (± 0.01) ppm in 1:4

ratio were assigned to the 4 equatorial/2-axial conformation of neo-IP6 based on Turner et

al. (2012). The peaks for D-chiro IP6 were detected in 2:2:2 ratio in the 2-equatorial/4-axial

conformation at 6.23 ppm, 4.75 (± 0.01) ppm, and 4.34 (± 0.03), ppm (Turner et al., 2012).

The signal from scyllo-IP6 occurred at 3.71 ± 0.02 ppm (Cade-Menun et al., 2010).

Glucose 6-phosphate was detected at 5.12 ± 0.03 ppm, while the 4.88 ± 0.01 ppm and 4.55

± 0.02 ppm peaks were assigned to α-glycerophosphate (α-glyc) and β-glycerophosphate

(β-glyc), respectively in a 1:2 ratio. Choline phosphate (chol-P, 3.85 ± 0.03 ppm) and two

signals of nucleotides (4.33 ± 0.01, 4.16 ± 0.02 ppm) were also detected in every soil

Page 131: Caractérisation par spectroscopie et analyse compositionnelle des

109

sample. Peak locations for β-glyc and chol-P were confirmed by spiking samples with these

two P forms.

Unidentified groups of the orthophosphate monoesters were divided into three

general groups based on their locations (Cade-Menun, 2005; Hill and Cade-Menun, 2009).

The Monoester 1 group contained unidentified peaks between 6.48 and 7 ppm, and included

a peak at 6.51± 0.04 ppm. Peaks in the Monoester 2 region were detected at 5.68 ± 0.02

ppm, 5.26 ± 0.01 ppm, and 4.75 ± 0.03 ppm. The Monoester 3 group included a peak at 3.79

± 0.15 ppm observed in the soil samples containing D-chiro IP6. An unknown peak was

detected at 4.93 ± 0.02 ppm. It currently has not been specifically identified; unpublished

work (Cade-Menun, 2014) suggests that it may be myo-1-IP, but this requires confirmation

with spiking experiments, which were not done for the current manuscript. The DNA peak

was found in most samples at -0.75 ± 0.03 ppm. The remaining orthophosphate diesters

were divided into two groups. Peaks for “Other diester 1”, were detected between 3.34 and

0.41 ppm, and may include diagnostic peaks for phospholipids (Cade-Menun et al., 2010).

The “Other diester 2” group was observed between 1.76 and -3.72 ppm. The unidentified

groups of the orthophosphate monoesters (Monoester 1, 2, and 3) and diester (Other diester

1 and 2), the unknown compound, and D-chiro IP6 that was not detected in all soil samples

were amalgamated and identified as a residual phosphate fraction “Res” of TP.

6.5.3 Distribution of 31P nuclear magnetic resonance phosphorus forms

The statistical results for the relative percentages of soil P forms determined by 31P-

NMR are presented in Table 3. The only P form to show an interaction of tillage, depth and

fertilization was orthophosphate, which was significant at p < 0.1. Orthophosphate was the

prominent fraction of extracted P in all analyzed soil samples (Table 4). The highest

concentration (365 ± 21 mg kg-1), corresponding to the relative percentage of extracted P of

49.7%, was found in the NT-P35 treatment at 0 to 5 cm depth and was significantly different

from the values at 5 to 20 cm (p < 0.01, Fig. 1). In contrast, the lowest concentration (206 ±

26 mg kg-1), which is equivalent to 37.7 % of extracted P, was observed in the deep layer (10-

20 cm) under MP-P0.

Significant effects of the interaction between tillage and depth were observed for

pyrophosphate, scyllo-IP6, nucleotides and DNA, while significant effects of the interaction

Page 132: Caractérisation par spectroscopie et analyse compositionnelle des

110

between tillage and fertilization occurred only for pyrophosphate (Table 3). Significant

differences were observed between the top soil (0-5 cm) and the deep (5-20 cm) layers

under no-till for pyrophosphate, scyllo-IP6 and nucleotides and between 0 to 10 cm and 10

to 20 cm for DNA (Fig. 5). In MP treatment, the DNA content increased significantly from

the top 5 cm layer (9.3 ± 0.4 mg kg-1) to 5 to 10 cm (11.6 ± 0.1 mg kg-1), and 10 to 20 cm

(11.9 ± 0.2 mg kg-1, Fig. 5).

Where there were no interactions with other treatments, the percentage of

phosphonate was significantly higher under the P35 treatment and the percentages of α and

β-glycerophosphate and choline phosphate were significantly higher under the P0 treatment

(Table 4). The percentages of α and β-glycerophosphate were the only P forms affected by

depth without an interaction to another treatment (Table 3), and were significantly higher at

10-20 cm than higher in the soil profile (Table 4). There were no significant differences

between treatments for polyphosphate, myo-IP6, neo-IP6, glucose-6P and the residue

fraction of P (Table 3).

6.6 DISCUSSION

Our results showed that TP significantly varied with soil depth under NT treatment,

where it accumulated in the top unfertilized soil layer (0–5 cm) and in the 0 to 10 cm layer

where P fertilizer was applied. This agrees with Redel et al. (2011) who found a significant

difference between NT and conventional tillage in TP amount in the 0 to 20 soil layer.

Additionally, pH, PM3, Mg, and TC were significantly higher in the topsoil (0–5 cm) of NT

treatment than the deeper layers (5–20 cm), and TN accumulated in the surface 10 cm. The

opposite trend was observed for Al. Conversely, the distributions of TC and TN were

homogenous along the soil profile under MP management where they slowly increased (Fig.

2). This suggests that stratification in NT results from the retention of crop residues at the

soil surface, where decomposition led to the release of soluble P and the accumulation of TC

and TN (Poirier et al., 2009), and to a pH increase (Hussain et al., 1999; Paul et al., 2001),

which produced a decrease in extractable Al (Shang et al., 1992) and an increase in

exchangeable Mg (Hussain et al. 1999). The stratification of PM3 in the 0 to 5 cm layer of

the NT-P0 treatment, in contrast with MP-P0 (Fig. 2), suggests that the build-up of organic

matter and the subsequent leaching of P during 18 years of soil cultivation played an

Page 133: Caractérisation par spectroscopie et analyse compositionnelle des

111

important role in P stratification. This accumulation of PM3 seemed to be accentuated in the

P-fertilized treatment (Fig. 1) due to a lack of mixing of soil with applied fertilizer. The

depletion of available P in lower layers of NT (5-20 cm), due in part to the higher bulk

density, may result in reduced P uptake by plants (Lupwayi et al., 2006). Thus, crop yield

was greater under MP than under zero tillage (Légère et al., 2008; Messiga et al., 2012). The

same trend for PM3 has been reported elsewhere for available P extracted with different

methods, when compared NT and MP systems applied on P fertilized soil ( Cade-Menun et

al., 2010; Duiker and Beegle, 2006; Hussain et al., 1999; Vu et al., 2009). The increase of

PM3 with P fertilization, regardless of soil tillage treatment and depths, was also observed in

samples collected in 2009 from the same research plot (Sheng et al., 2013). Likewise, the

31P-NMR spectra showed concomitant stratification of orthophosphate in NT-P35 (Fig. 1).

Shi et al. (2012) found greater alkaline phosphatase activity, which mineralized more

orthophosphate monoesters at the surface soil under NT management in greater amounts

compared to MP of the same research site in 2009. This could easily explain the

significantly higher amount of orthophosphate on the NT soil surface compared to

conventional tillage, as could stratification of the orthophosphate applied in fertilizer.

Stratification also occurred under NT for scyllo-IP6 and the microbe-associated

compounds: pyrophosphate, nucleotides and DNA (Fig. 5). The origin of scyllo-IP6 remains

unknown, although the fact that it is a stereoisomer of myo-IP6 (Turner et al., 2002) points

to epimerization reactions (L'Annunziata, 1975) or microbial production (Caldwell and

Black, 1958). However, no significant differences between the treatments were found for

myo-IP6 and neo-IP6, which were uniformly distributed across treatments (Table 2). Under

NT treatment, the pyrophosphate and DNA concentrations were greater at the surface (0–5

cm) soil layer than the deeper layers (Fig. 5), whereas DNA significantly accumulated

below the 5 cm layer under MP. Total C and N were similarly affected (Fig. 2), suggesting

that DNA was synthesized in greater amounts under NT owing to the higher organic matter

accumulated at the soil surface (Condron et al., 2005) in comparison to conventional

tillage. It was unlikely to have accumulated due to sorption to the mineral surface, because

the soil pH was higher than the isoelectric point (pH 5.0) of DNA (Condron et al., 2005).

The opposite trend was observed for the nucleotides, which may be derived from the

hydrolysis of RNA during NMR analysis (He et al., 2011), or may be present naturally in

Page 134: Caractérisation par spectroscopie et analyse compositionnelle des

112

soil (Vestergren et al., 2012). Under direct drill, scyllo-IP6 and nucleotides accumulated

significantly in the deeper layers (Fig. 5), possibly due to their preferential movement

through the soil column (Condron et al., 2005). The accumulation of scyllo-IP6 could be

attributed also to the higher amount of Mehlich-3 extractable Al in the 5 to 20 cm soil layer

(Fig. 2). Indeed, Shang et al. (1992) found that inositol hexakisphosphate sorption was

dependent on the contents of Al oxides and the adsorption rate increased when the pH

decreased. Mononucleotides, and the RNA from which they likely originated in the soil

sample, do not sorb strongly to the soil; their increase may reflect higher microbial

biomass, but this was not analyzed for this study.

The α- and β-glycerophosphates and choline-P have been identified as degradation

products of phospholipids of cellular membranes during NMR analysis (Doolette et al.,

2009; He et al., 2011; Young et al., 2013). Both α and β-glycerophosphate increased

significantly in the soil profile regardless of treatment, suggesting that without degradation

phospholipids would also have increased. Little is known about these compounds in soils

(Cade-Menun et al., 2010), so it is difficult to explain the factors and processes controlling

their presence. The percentages of α and β-glycerophosphate and choline-P were reduced

under P fertilization in general. However, phosphonates, which are more stable in soils and

thus less bioavailable (Condron et al., 2005), were detected at a higher percentage in soils

receiving 35 kg P ha-1 than in unfertilized soils.

Overall, it appears that labile inorganic P accumulated at the surface of no-till soil

and decreased with depth, while organic P, such as scyllo-IP6 and nucleotides, increased

deeper in the soil profile. Consequently, NT may increase the potential for soluble

inorganic P loss in surface runoff (Kleinman et al., 2009), and lead to the loss of organic

monoesters draining through different hydrological pathways (Condron et al., 2005). These

hypotheses were further confirmed by Dodd et al. (2014) who showed that NT system

largely increased soluble P concentrations in the surface run-off and dissolved reactive P in

leachate compared with MP.

As knowledge on molecular species increases for the soil P cycle, more elaborate

theories could be built relating inorganic and organic P species into a framework of

balances using isometric log ratios (Filzmoser et al., 2009), similarly to balance designs in

Page 135: Caractérisation par spectroscopie et analyse compositionnelle des

113

biochemistry (Aslam et al., 2013) and ionomics (Parent et al., 2013). An optimum balance

between molecular species and beneficial cultural practices could be defined to maximize

crop yield while minimizing P losses.

6.7 CONCLUSIONS

In this study, we showed that long-term zero tillage resulted in the stratification of

TP, TC, TN, pH, Al, Mg, and PM3 in the soil profile, owing to the build-up of organic

matter at soil surface. The accumulation of soluble inorganic P in the top 5 cm was higher

with P fertilizer application due in part to the lack of mixing of fertilizer. Additionally, no-

till treatment led to the accumulation of 31P-NMR monoesters such as the scyllo-IP6 and the

nucleotides in deeper layers. This stratification may increase the potential of soluble

inorganic P loss by surface runoff, and organic P by leaching and drainage. The relative

percentage of simple orthophosphate monoesters such as the β-glycerophosphate and

choline-P were lower under P fertilization, whereas the percentage of phosphonates was

higher. Further investigations are needed to unravel the interactions between P species and

to understand P changes as perturbed by agricultural management practices.

6.8 ACKNOWLEDGMENTS

This study was funded by The Sustainable Agriculture Environmental Systems

(SAGES) initiative of Agriculture and Agri-Food Canada and by the Natural Sciences and

Engineering Research Council of Canada (NSERC-DG 2254). The authors are very

grateful to S. Côté and S. Michaud for their technical support during sampling operations

and laboratory analysis. The NMR work was conducted at the Stanford Magnetic

Resonance Laboratory at the Stanford University School of Medicine. We gratefully thank

Dr. Corey Liu for his assistance with the P-NMR analysis.

Page 136: Caractérisation par spectroscopie et analyse compositionnelle des

114

6.9 REFERENCES

Ahlgren, J., F. Djodjic, G. Börjesson, and L. Mattsson. 2013. Identification and

quantification of organic phosphorus forms in soils from fertility experiments. Soil

Use Manage. 29:24-35.

Aitchison, J. 1986. The statistical analysis of compositional data. Monographs on Statistics

and Applied Probability. 1sted. London, UK.

Aslam, S., P. Garnier, C. Rumpel, S. E. Parent, and P. Benoit. 2013. Adsorption and

desorption behavior of selected pesticides as influenced by decomposition of maize

mulch. Chemosphere 91: 1447–55.

Cade-Menun, B.J. 2005. Characterizing phosphorus in environmental and agricultural

samples by 31P nuclear magnetic resonance spectroscopy. Talanta 66:359-371.

Cade-Menun, B.J., and C.W. Liu. 2014. Solution 31P-NMR spectroscopy of soils from

2005-2013: A review of sample preparation and experimental parameters. Soil Sci.

Soc. Am. J. 78 (1):19-37.

Cade-Menun, B.J., and C.M. Preston. 1996. A comparison of soil extraction procedures for

31P NMR spectroscopy. Soil Sci. 161:770-785.

Cade-Menun, B.J., M.R. Carter, D.C. James, and C.W. Liu. 2010. Phosphorus forms and

chemistry in the soil profile under long-term conservation tillage: A phosphorus-31

nuclear magnetic resonance study. J. Environ. Qual. 39:1647-1656.

Cade-Menun, B.J., G. Bell, S. Baker-Ismail, Y. Fouli, K. Hodder, D.W. McMartin, C.

Perez-Valdivia, and K. Wu. 2013. Nutrient loss from Saskatchewan cropland and

pasture in spring snowmelt runoff. Can. J. Soil Sci. 93:445-458.

Caldwell, A.G., and C.A. Black. 1958. Inositol hexaphosphate: II. Synthesis by soil

microorganisms. Soil Sci. Soc. Am. J. 22:293-296.

Chichester, F.W., and C.W. Richardson. 1992. Sediment and nutrient loss from clay soils as

affected by tillage. J. Environ. Qual. 21:587-590.

Condron, L.M., B.L. Turner, and B.J. Cade-Menun. 2005. The chemistry and dynamics of

soil organic phosphorus. In: J.T. Sims and A.N. Sharpley, editors, Phosphorus:

agriculture and the environment. Wisconsin, USA. p. 87-121.

Dick, R.P. 1992. A review: long-term effects of agricultural systems on soil biochemical

and microbial parameters. Agric. Ecosyst. Environ. 40:25-36.

Dodd, R.J., R.W. McDowell, and L.M. Condron. 2014. Is tillage an effective method to

decrease phosphorus loss from phosphorus enriched pastoral soils? Soil Tillage Res.

135:1-8.

Doolette, A.L., and R.J. Smernik. 2011. Soil organic phosphorus speciation using

spectroscopic techniques. In: E.K. Bünemann, A. Oberson and E. Frossard editors,

Phosphorus in Action. Springer Heidelberg Dordrecht London, New York. p. 3-36.

Doolette, A.L., R.J. Smernik, and W.J. Dougherty. 2009. Spiking improved solution

phosphorus-31 nuclear magnetic resonance identification of soil phosphorus. Soil

Sci. Soc. Am. J. 73:919-927.

Duiker, S.W., and D.B. Beegle. 2006. Soil fertility distributions in long-term no-till,

chisel/disk and moldboard plow/disk systems. Soil Tillage Res. 88:30-41.

Egozcue, J.J., V. Pawlowsky-Glahn, G. Mateu-Figueras, and C. Barceló-Vidal. 2003.

Isometric logratio transformations for compositional data analysis. Math. Geol.

35:279-300.

Page 137: Caractérisation par spectroscopie et analyse compositionnelle des

115

Elser, J.J., M.E.S. Bracken, E.E. Cleland, D.S. Gruner, W.S. Harpole, H. Hillebrand, J.T.

Ngai, E.W. Seabloom, J.B. Shurin, and J.E. Smith. 2007. Global analysis of

nitrogen and phosphorus limitation of primary producers in freshwater, marine and

terrestrial ecosystems. Ecol. Lett. 10:1135-1142.

Fernández, F.G., S.M. Brouder, C.A. Beyrouty, J.J. Volenec, and R. Hoyum. 2008.

Assessment of plant-available potassium for no-till, Rainfed Soybean. Soil Sci.

Soc. Am. J. 72:1085-1095.

Filzmoser, P., and K. Hron. 2009. Correlation analysis for compositional data. Math.

Geosci. 41:905-919.

Filzmoser, P., K.Hron, and C. Reimann. 2009. Univariate statistical analysis of

environmental (compositional) data: Problems and possibilities. Sci. Total Environ.

407:6100-6108.

Grant, C.A., and G.P. Lafond. 1994. The effects of tillage systems and crop rotations on

soil chemical properties of a Black Chernozemic soil. Can. J. Soil Sci. 74:301-306.

Haygarth, P.M., L. Hepworth, and S.C. Jarvis. 1998. Forms of phosphorus transfer in

hydrological pathways from soil under grazed grassland. Eur. J. Soil Sci. 49:65-72.

He, Z., D.C. Olk, and B.J. Cade-Menun. 2011. Forms and lability of phosphorus in humic

acid fractions of hord silt loam soil. Soil Sci. Soc. Am. J. 75:1712-1722.

Hendershot, W.H., H. Lalande, and M. Duquette. 2008. Soil reaction and exchangeable

acidity. In: Carter MR and EG Gregorich editors, Soil Sampling and Methods of

Analysis. Can Soc Soil Sci, CRC Press Inc., Boca Raton, FL. p 173–178.

Hill, J.E., and B.J. Cade-Menun. 2009. Phosphorus-31 nuclear magnetic resonance

spectroscopy transect study of poultry operations on the Delmarva Peninsula. J.

Environ. Qual. 38:130-138.

Holm, F.A., R.P. Zentner, A.G. Thomas, K. Sapsford, A. Légère, B.D. Gossen, O. Olfert,

and J.Y. Leeson. 2006. Agronomic and economic responses to integrated weed

management systems and fungicide in a wheat-canola-barley-pea rotation. Can. J.

Plant Sci. 86:1281-1295.

Hussain, I., K.R. Olson, and S.A. Ebelhar. 1999. Long-term tillage effects on soil chemical

properties and organic matter fractions. Soil Sci. Soc. Am. J. 63:1335-1341.

Kleinman, P.A., A. Sharpley, L. Saporito, A. Buda, and R. Bryant. 2009. Application of

manure to no-till soils: phosphorus losses by sub-surface and surface pathways.

Nutr. Cycl. Agroecosyst. 84:215-227.

L'Annunziata, M.F. 1975. The origin and transformations of the soil inositol phosphate

isomers. Soil Sci. Soc. Am. J. 39:377-379.

Lafond, G.P., F. Walley, W.E. May, and C.B. Holzapfel. 2011. Long term impact of no-till

on soil properties and crop productivity on the Canadian prairies. Soil Tillage Res.

117:110-123.

Légère, A., F.C. Stevenson., and N. Ziadi. 2008. Contrasting responses of weed

communities and crops to 12 years of tillage and fertilization treatments. Weed

Tech. 22:309-317.

Lupwayi, N.Z., G.W. Clayton, J.T. O’Donovan, K.N. Harker, T.K. Turkington, and Y.K.

Soon. 2006. Soil nutrient stratification and uptake by wheat after seven years of

conventional and zero tillage in the Northern Grain belt of Canada. Can. J. Soil Sci.

86:767-778.

Page 138: Caractérisation par spectroscopie et analyse compositionnelle des

116

Martín-Fernández, J.A., C. Barceló-Vidal, and V. Pawlowsky-Glahn. 2003. Dealing with

zeros and missing values in compositional data sets using nonparametric

imputation. Math. Geol. 35:253-278.

McDowell, R.W., and G.F. Koopmans. 2006. Assessing the bioavailability of dissolved

organic phosphorus in pasture and cultivated soils treated with different rates of

nitrogen fertiliser. Soil Biol. Biochem. 38:61-70.

McDowell, R.W., I. Stewart, and B.J. Cade-Menun. 2006. An examination of spin–lattice

relaxation times for analysis of soil and manure extracts by liquid state phosphorus-

31 nuclear magnetic resonance spectroscopy. J. Environ. Qual. 35:293-302.

McDowell, R.W., B.J. Cade-Menun, and I. Stewart. 2007. Organic phosphorus speciation

and pedogenesis: analysis by solution 31P nuclear magnetic resonance

spectroscopy. Europ. J. Soil Sci. 58:1348-1357.

Mehlich, A. 1984. Mehlich 3 soil test extractant: A modification of Mehlich 2 extractant.

Commun. Soil Sci. Plant Anal. 15:1409-1416.

Messiga, A.J., N. Ziadi, C. Morel, C. Grant, G. Tremblay, G. Lamarre, and L.-E. Parent.

2012. Long term impact of tillage practices and biennial P and N fertilization on

maize and soybean yields and soil P status. Field Crops Res. 133:10-22.

Murphy, J., and J.P. Riley. 1962. A modified single solution method for the determination

of phosphate in natural waters. Anal. Chim. Acta 27:31-36.

Nelson, N.S. 1987. An acid‐persulfate digestion procedure for determination of phosphorus

in sediments. Commun. Soil Sci. Plant Anal. 18:359-369.

Olson, K.R., and S.A. Ebelhar. 2009. Impacts of conservation tillage systems on long-term

crop yields. J. Agron. 8:14-20.

Parent, L. E., S.-É. Parent, V. Hébert-Gentile, K. Naess, and L. Lapointe. 2013. Mineral

balance plasticity of cloudberry (Rubus chamaemorus) in Quebec-Labrador. Am. J.

Plant Sci. 4:1508-1520.

Paul, K.I., A.S. Black, and M.K. Conyers. 2001. Effect of plant residue return on the

development of surface soil pH gradients. Biol. Fertil. Soils 33:75-82.

Poirier, V., D.A. Angers, P. Rochette, M.H. Chantigny, N. Ziadi, G. Tremblay, and J.

Fortin. 2009. Interactive effects of tillage and mineral fertilization on soil carbon

profiles. Soil Sci. Soc. Am. J. 73:255-261.

Redel, Y.D., M. Escudey, M. Alvear, J. Conrad, and F. Borie. 2011. Effects of tillage and

crop rotation on chemical phosphorus forms and some related biological activities

in a Chilean Ultisol. Soil Use Manage. 27:221-228.

SAS Institute. 2001. The SAS system for Windows. Release 8.2. SAS Inst., Cary, NC.

Selles, F., B.G. McConkey, and C.A. Campbell. 1999. Distribution and forms of P under

cultivator- and zero-tillage for continuous- and fallow-wheat cropping systems in

the semi-arid Canadian prairies. Soil Tillage Res. 51:47-59.

Shang, C., J.W.B. Stewart, and P.M. Huang. 1992. pH effect on kinetics of adsorption of

organic and inorganic phosphates by short-range ordered aluminum and iron

precipitates. Geoderma 53:1-14.

Sharpley, A.N. 2003. Soil mixing to decrease surface stratification of phosphorus in

manured soils. J. Environ. Qual. 32:1375-1384.

Sharpley, A.N., and S.J. Smith. 1994. Wheat tillage and water quality in the Southern

plains. Soil Tillage Res. 30:33-48.

Page 139: Caractérisation par spectroscopie et analyse compositionnelle des

117

Sheng, M., R. Lalande, C. Hamel, and N. Ziadi. 2013. Effect of long-term tillage and

mineral phosphorus fertilization on arbuscular mycorrhizal fungi in a humid

continental zone of Eastern Canada. Plant Soil:1-15.

Shepard, R. 2005. Nutrient management planning: Is it the answer to better management? J.

Soil Water Conserv. 60:171-176.

Shi, Y., R. Lalande, N. Ziadi, M. Sheng, and Z. Hu. 2012. An assessment of the soil

microbial status after 17 years of tillage and mineral P fertilization management.

Appli. Soil Ecol. 62:14-23.

Tremblay, G., L. Robert, P. Filion, G. Govaerts, R. Mongeau, J. Filiatrault, J.M. Beausoleil,

G. Moreau, T.S. Tran, 2003. Régies culturales et fertilisations azotée et phosphatée

dans une rotation maïs-soya. Bulletin Technique No. 3.05. CÉROM, Saint-Bruno-

de-Montarville, Quebec City, QC, 8 pp.

Turner, B.L., N. Mahieu, and L.M. Condron. 2003. Quantification of myo-inositol

hexakisphosphate in alkaline soil extracts by solution 31P nmr spectroscopy and

spectral deconvolution. Soil Sci. 168:469-478.

Turner, B.L., M.J. Papházy, P.M. Haygarth, and I.D. McKelvie. 2002. Inositol phosphates

in the environment. Phil Trans Roy Soc London B: 357:449-469.

Turner, B.L., A.W. Cheesman, H.Y. Godage, A.M. Riley, and B.V.L. Potter. 2012.

Determination of neo- and d-chiro-inositol hexakisphosphate in soils by solution

31P NMR spectroscopy. Environ. Sci. Technol. 46:4994-5002.

van den Boogaart, K.G., R. Tolosana-Delgado, and M. Bren. 2011. Compositions:

compositional data analysis, R Package Version 1.10-2.

http://CRAN.Rproject.org/package=compositions.

Vestergren, J., A.G. Vincent, M. Jansson, P. Persson, U. Ilstedt, G. Gröbner, R. Giesler,

and J. Schleucher. 2012. High-resolution characterization of organic phosphorus in

soil extracts using 2D 1H–31P nmr correlation spectroscopy. Environ. Sci. Technol.

46:3950-3956.

Vu, D.T., C. Tang, and R.D. Armstrong. 2009. Tillage system affects phosphorus form and

depth distribution in three contrasting Victorian soils. Soil Res. 47:33-45.

Whitton, B.A., S.L.J. Grainger, G.R.W. Hawley, and J.W. Simon. 1991. Cell-bound and

extracellular phosphatase activities of cyanobacterial isolates. Microbial Ecol.

21:85-98.

Young, E.O., D.S. Ross, B.J. Cade-Menun, and C.W. Liu. 2013. Phosphorus speciation in

riparian soils: A phosphorus-31 nuclear magnetic resonance spectroscopy and

enzyme hydrolysis study. Soil Sci. Soc. Am. J. 77:1636-1647.

Zibilske, L.M., J.M. Bradford, and J.R. Smart. 2002. Conservation tillage induced changes

in organic carbon, total nitrogen and available phosphorus in a semi-arid alkaline

subtropical soil. Soil Tillage Res. 66:153-163.

Page 140: Caractérisation par spectroscopie et analyse compositionnelle des

118

Table 6-1 Analysis of variance for the effects of tillage, P fertilization and depth on clr transformed concentrations of soil total P

(TP), Mehlich-3 extractable P (PM3), aluminium (Al), iron (Fe), calcium (Ca), magnesium (Mg), total carbon (TC) and total

nitrogen (TN), and pH.

Sources of

variation

TP† PM3 Al Fe Ca Mg TC TN pH

Tillage (T) NS‡ NS NS NS NS NS NS NS NS

Phosphorus (P) NS ** NS NS NS NS NS NS NS

Depth (D) NS *** *** ** NS * ** **** ****

T × P NS NS NS NS NS NS NS NS NS

T × D **** ** *** NS NS **** ** * *

P × D NS * NS NS NS NS NS NS NS

T × P × D * * NS NS NS NS NS NS NS

* Significant at the 0.05 probability level.

** Significant at the 0.01 probability level.

*** Significant at the 0.001 probability level.

**** Significant at the 0.10 probability level.

† PM3, Mehlich-3–extractable P; TC, total C; TN, total N; TP, total P.

‡ Nonsignificant.

Page 141: Caractérisation par spectroscopie et analyse compositionnelle des

119

Table 6-2 Chemical shift of P forms detected in the 31P-NMR spectrum of the soil as

affected by tillage and P fertilization management and depth.

Category P form or compound class Chemical shift

ppm

Inorganic P orthophosphate 6.00

pyrophosphate 4.02

polyphosphates 4.23 to 24.76

polyphosphate end group 3.98

Organic P phosphonates 25.0–7.8 (signals at 20.36,

18.76)

Orthophosphate monoesters Myo-IP6 5.48, 4.51, 4.11, 4.02

Neo- IP6 6.41, 4.27

D-chiro IP6 6.23, 4.75, 4.34

Scyllo- IP6 3.71

glucose 6-phosphate 5.12

-glycerophosphate 4.88

-glycerophosphate 4.55

choline phosphate 3.85

nucleotides 4.33, 4.16

monoester 1 7.0–6.48 (signal at 6.51)

monoester 2 5.68, 5.26, 4.75

monoester 3 3.79

unknown 4.93

Orthophosphate diesters DNA 0.75

other diester 1 3.34–0.41

other diester 2 1.76 to 3.72

Page 142: Caractérisation par spectroscopie et analyse compositionnelle des

120

Table 6-3 Analysis of variance for the effects of tillage, P fertilization, and depth on centered log ratio–transformed soil P forms

determined by 31P nuclear magnetic resonance spectroscopy.

Sources of

variation

Inorganic P† Organic P‡

Ortho Pyro Poly Phos myo-IP6 neo-

IP6

scyllo-IP6 Gluc-6P -Glyc -Glyc Chol-

P

Nucl DNA Res

P

Tillage (T) NS§ NS NS NS NS NS NS NS NS NS NS NS NS NS

Phosphorus (P) * NS NS * NS NS NS NS **** **** **** NS NS NS

Depth (D) * **** NS NS NS NS * NS **** **** NS * NS NS

T × P NS **** NS NS NS NS NS NS NS NS NS NS NS NS

T × D NS **** NS NS NS NS **** NS NS NS NS ** ** NS

P × D NS NS NS NS NS NS NS NS NS NS NS NS NS NS

T × P × D **** NS NS NS NS NS NS NS NS NS NS NS NS NS

* Significant at the 0.05 probability level.

** Significant at the 0.01 probability level.

**** Significant at the 0.10 probability level.

† Ortho, orthophosphate; Poly, polyphosphate; Pyro, pyrophosphate.

‡ -Glyc, -glycerophosphate; -Glyc, -glycerophosphate; Chol-P, choline-phosphate; Gluc-6P, glucose-6 phosphate; myo-IP6, myo-inositol

hexakisphosphate; neo-IP6, neo-inositol hexakisphosphate; Nucl, nucleotides; Phos, phosphonate; scyllo-IP6, scyllo-inositol hexakisphosphate; Res P,

residual organic P (unidentified organic P forms).

§ Nonsignificant.

Page 143: Caractérisation par spectroscopie et analyse compositionnelle des

121

Table 6-4 Back-centered log ratio–transformed soil P forms determined by 31P nuclear magnetic resonance spectroscopy as

affected by tillage, P fertilization, and depth.

Treatment

Inorganic P‡ Organic P§

Ortho Pyro Poly Phos myo-

IP6

neo-IP6 scyllo-

IP6

Gluc-

6P -

Glyc

-Glyc Nucl Chol-

P

DNA Res P

% total P

Tillage

MP 42.3 ± 1.4a¶ 1.0 ± 0.1a 0.8 ±

0.2a

2.5 ±

0.2a

10.0 ±

0.3a

4.0 ± 0.1a 4.0 ± 0.2a 2.5 ±

0.1a

1.5 ±

0.1a

2.9 ±

0.3a

5.8 ±

0.4a

2.2 ±

0.2a

1.8 ±

0.1a

18.8

± 0.6a

NT 41.2 ± 1.5a 0.9 ± 0.1a 0.6 ±

0.1a

2.9 ±

0.2a

9.8 ±

0.3a

4.0 ± 0.1a 3.7 ± 0.3a 2.2 ±

0.1a

1.8 ±

0.1a

3.6 ±

0.2a

5.2 ±

0.5a

2.0 ±

0.2a

1.7 ±

0.1a

20.3 ±

0.9a

Phosphorus

P0 38.1 ± 1.0b 1.1 ± 0.1a 0.7 ±

0.1a

2.4 ±

0.2b

10.3 ±

0.4a

4.0 ± 0.1a 4.3 ± 0.2a 2.5 ±

0.2a

1.8 ±

0.1a#

3.6 ±

0.2a#

6.2 ±

0.3a

2.5 ±

0.2a

1.9 ±

0.2a

20.9 ±

0.9a

P35 45.6 ± 1.3a 0.9 ± 0.1a 0.8 ±

0.2a

3.0 ±

0.2a

9.5 ±

0.1a

3.9 ± 0.0a 3.5 ± 0.3a 2.2 ±

0.1a

1.5 ±

0.1b#

2.9 ±

0.2b#

4.8 ±

0.5a

1.7 ±

0.1b

1.6 ±

0.1a

18.1 ±

0.5a

Depth

0–5 cm 44.0 ± 1.8a 1.1 ± 0.1a 0.9 ±

0.2a

2.5 ±

0.2a

9.9 ±

0.3a

3.9 ± 0.1a 3.5 ± 0.3b 2.2 ±

0.1a

1.5 ±

0.1b

3.0 ±

0.3b

4.8 ±

0.5b

1.9 ±

0.2a

1.7 ±

0.1a

19.2 ±

1.1a

5–10 cm 41.3 ± 1.8b 0.9 ± 0.1b 0.6 ±

0.2a

3.0 ±

0.3a

9.9 ±

0.4a

3.9 ± 0.1a 4.2 ± 0.3a 2.3 ±

0.2a

1.6 ±

0.1b#

3.1 ±

0.3b#

5.9 ±

0.5a

2.3 ±

0.3a

1.7 ±

0.2a

19.3 ±

0.8a

10–20 cm 40.1 ± 1.9b 0.9 ± 0.2b 0.7 ±

0.1a

2.7 ±

0.3a

10.0 ±

0.4a

4.0 ± 0.1a 4.0 ± 0.3a 2.5 ±

0.1a

1.8 ±

0.2a#

3.6 ±

0.3a#

5.8 ±

0.5a

2.1 ±

0.2a

1.8 ±

0.2a

20.1 ±

0.9a

† MP, moldboard plow; NT, no-till; P0, soil treatment with 0 kg P ha1; P35, soil treatment with 35 kg P ha1.

‡ Ortho, orthophosphate; Poly, polyphosphate; Pyro, pyrophosphate.

§-Glyc, -glycerophosphate; -Glyc, -glycerophosphate; Chol-P, choline-phosphate; Gluc-6P, glucose-6 phosphate; myo-IP6, myo-inositol

hexakisphosphate; neo-IP6, neo-inositol hexakisphosphate; Nucl, nucleotides; Phos, phosphonate; scyllo-IP6, scyllo-inositol hexakisphosphate; Res P,

residual organic P (unidentified organic P forms).

¶ For each treatment, different letters indicate significantly different according to LSD (0.05).

# For each treatment, different letters indicate significantly different according to LSD (0.1).

Page 144: Caractérisation par spectroscopie et analyse compositionnelle des

122

Figure 6-1 Distribution of total P (TP), Mehlich-3 extractable P (PM3) and orthophosphate

concentrations at various soil depths under (a, c, e) mouldboard plow (MP) and (b, d, f) no-

till (NT) treatments. P0 and P35 represent additions of 0 and 35 kg P ha−1, respectively.

Values are means of three replications. For each treatment, different letters indicate

significantly different means among soil depth according to LSD (0.05). † For each

treatment, different letters indicate significantly different means among depth according to

LSD (0.1).

Page 145: Caractérisation par spectroscopie et analyse compositionnelle des

123

Figure 6-2 Distribution of (a) total carbon (TC) and (b) total nitrogen (TN) content, and (c)

Al Mehlich-3 and (d) Mg Mehlich-3 at various soil depths under mouldboard plow (MP)

and no-till (NT) treatments. Values are means of three replicates. For each treatment,

different letters indicate significantly different means among soil depth according to LSD

(0.05).

Page 146: Caractérisation par spectroscopie et analyse compositionnelle des

124

Figure 6-3 Phosphorus-31 nuclear magnetic resonance spectroscopy spectrum showing the

range of P compounds detected at the 0 to 5 cm depth of the mouldboard plow fertilized

treatment (Oth.D1, other diester 1; Oth.D2, other diester 2).

Page 147: Caractérisation par spectroscopie et analyse compositionnelle des

125

Figure 6-4 Phosphorus-31 nuclear magnetic resonance spectroscopy spectrum showing the

P compounds detected in the monoester region at the 0 to 5 cm depth of mouldboard plow

fertilized treatment. (A) neo-IP6; (B) orthophosphate; (C) myo-IP6; (D) glucose-6P; (E)

unknown; (F) α-glycerophosphate; (G) β-glycerophosphate; (H) nucleotides; (I) choline-P;

(J) scyllo-IP6; (M1) monoester 1; (M2) monoester 2.

Page 148: Caractérisation par spectroscopie et analyse compositionnelle des

126

Figure 6-5 Distributions of (a) pyrophosphate, (b) scyllo-IP6, (c) DNA and (d) nucleotides

concentrations at various soil depths under mouldboard plow (MP) and no-till (NT)

treatments. Values are means of three replicates. For each treatment, different letters

indicate significantly different means among depths according to LSD (0.05).

Page 149: Caractérisation par spectroscopie et analyse compositionnelle des

127

CHAPITRE VII: CONCLUSIONS ET RECOMMANDATIONS

Le phosphore est un élément nutritif indispensable pour la production végétale. Il

est présent sous différentes formes qui coexistent par des mécanismes physiques, chimiques

et biologiques dans un cycle fermé. Dans les sols cultivés, les pratiques culturales

perturbent le cycle de P et peuvent affecter la production agricole et/ou contribuer à

l’eutrophisation des eaux de surface. La compréhension de la dynamique et des

changements des formes du P dans les écosystèmes agricoles est un élément clé pour une

meilleure gestion de P. Les méthodes utilisées jusqu’à date pour caractériser les

changements des pools de P se basent sur une description opérationnelle. D’autre part, les

fractions de P sont définies comme étant des données compositionnelles; soient des

données strictement positives comprises entre 0 et une quelconque unité de mesure. Selon

Aitchison (1986), l’application des statistiques conventionnelles sur ces données peut

générer des résultats erronés menant à des interprétations contradictoires. En outre, les

techniques couramment utilisées pour mesurer le P du sol sont laborieuses, relativement

lentes, coûteuses, et nécessitent l’utilisation de plusieurs extractifs chimiques. L’objectif

général de cette thèse a été de mesurer les formes du P du sol et d’étudier leurs

changements selon les pratiques culturales moyennant de nouveaux outils.

Nos objectifs étaient de (i) évaluer le potentiel de la spectroscopie dans le proche

infrarouge (SPIR) à prédire le P total, le P disponible extrait à la solution Mehlich-3 (PM3)

et à l’eau (Cp), (ii) évaluer le potentiel de la SPIR à prédire le P organique, (iii) démontrer

que l’analyse compositionnelle permet d’étudier les formes chimiques du P sans biais, et

(iv) utiliser la spectroscopie de résonance magnétique nucléaire du 31P et l’analyse

compositionnelle pour identifier les espèces ioniques et moléculaires du P et caractériser

leur distribution selon le système du travail du sol et la fertilisation phosphatée.

Ces objectifs ont été atteints et les conclusions suivantes relatives à chacun d’eux

sont tirées :

Les résultats de la première étude de cette thèse (troisième chapitre) ont démontré

que le P total est modérément prédictible par la spectroscopie dans le proche-infrarouge

pour un sol sablo-loameux au Québec. Ce modèle de prédiction pourrait être appliqué dans

des situations où les exigences de précision sont relativement faibles. Néanmoins, la SPIR

Page 150: Caractérisation par spectroscopie et analyse compositionnelle des

128

ne peut pas être une alternative à l’analyse chimique conventionnelle du P disponible à la

plante tel que mesuré par les méthodes de Mehlich-3 (PM3) ou à l’eau (Cp).

Dans la deuxième étude (quatrième chapitre), les résultats ont montré que le PT ne

peut pas être prédictible par la SPIR dans des sols loameux et argileux-loameux,

contrairement au PM3. Ces résultats contradictoires entre les deux études peuvent

s’expliquer en partie par la sensibilité de cette technique spectroscopique à la texture du sol,

à la méthode de référence utilisée et à la variation de la teneur du P dans le sol. De ce fait,

ces modèles de prédiction par la SPIR devraient être validés dans d’autres sites de texture

différentes et teneurs variables en P. Un résultat important dans ce chapitre est que le P

organique est prédictible directement par la SPIR dans ces sols de texture moyennement

grossière à moyennement fine dû probablement à son lien avec la matière organique. Ceci

constitue un progrès dans les techniques analytiques du Po permettant de promouvoir sa

caractérisation et son étude dans les écosystèmes, étant donné que la SPIR est une

technique rapide, directe, économique et durable de point de vue environnemental. La

troisième étude de la thèse a permis de démontrer que l’analyse statistique conventionnelle

des espèces de P déterminées par la RMN-31P est biaisée. En effet, nous avons démontré

que les résultats des analyses de variance et de corrélation des espèces de P mesurées en

proportions ou en concentrations, brutes ou log transformées, sont souvent différents

menant à des interprétations incohérentes. Ceci est le résultat de la redondance, la

dépendance de l’échelle, et la distribution non-normale des données compositionnelles.

L’utilisation des transformations du log ratio centré (clr) ou isométrique (ilr) a permis

d’éviter ce biais statistique et d’avoir des résultats fiables et cohérents. Ainsi, nous avons

révélé dans ce chapitre l’importance de l’analyse compositionnelle pour une étude non

biaisée des formes de P dans le sol, et nous recommandons son utilisation pour l’étude des

formes de P dans l’environnement.

Dans la dernière étude de cette thèse, nous avons caractérisé les espèces ioniques et

moléculaires de P et leur distribution dans des sols collectés dans une rotation maïs soya

moyennant la résonance magnétique nucléaire du 31P. Les résultats ont démontré que

l’accumulation de P dans la couche superficielle du semis direct est principalement due aux

ions orthophosphates, et elle est plus importante dans les sols fertilisés. Cependant, les

formes organiques s’accumulaient en profondeur (5–20 cm) sous forme d’inositols

Page 151: Caractérisation par spectroscopie et analyse compositionnelle des

129

monoesters susceptibles d’atteindre les cours d’eaux adjacents par drainage. Sur la base de

ces résultats, nous recommandons de déterminer les apports de P au sol sous semis direct

selon des analyses de P par profondeur (0-5, 5-10 et 10-20 cm) pour éviter l’accumulation

du P en surface et s’assurer d’apporter les quantités en P disponible nécessaires aux plantes.

Page 152: Caractérisation par spectroscopie et analyse compositionnelle des

130

Page 153: Caractérisation par spectroscopie et analyse compositionnelle des

131

CHAPITRE VIII: ANNEXE

COMPOSITIONAL ANALYSIS OF POOLS IN CANADIAN

MOLLISOLS

D. ABDI1,2, N. ZIADI2 and L-É. PARENT1

1Agriculture and Agri-Food Canada, 2560 Hochelaga Boulevard, Quebec City, Quebec,

Canada, G1V 2J3; 2Département des sols et de génie agroalimentaire, Université Laval,

Quebec City, QC, Canada, G1K 7P4.

Proceeding of 4th International Workshop on

Compositional Data Analysis, Spain, 2011

Page 154: Caractérisation par spectroscopie et analyse compositionnelle des

132

8.1 RÉSUMÉ

Le cycle du phosphore dans les Mollisols des prairies canadiennes est perturbé par

les pratiques culturales. Les modèles utilisés jusqu’à date ne prennent pas en considération

les intéractions entre les différents pools du P. L’analyse compositionnelle utilisant les

coordonnées du log-ratio isométrique (ilr) est appropriée pour modeler ces intéractions.

L’objectif de l’étude était de modéler les changements des pools de P dans des Mollisols en

fonction des pratiques culturales et du temps en utilisant les ilr. Deux bases de données

publiées ont été utilisées. Les résultats ont démontré que la rotation culturale et la

fertilisation changeaient majoritairement la balance entre les pools inorganique et

organique, et celle entre les pools les plus disponibles et les moins disponibles du P

inorganique en augmentant le risque de perte de P. Les résultats ont montré aussi des

changements importants dans l’horizon Ah au cours du temps, dus à la perte en P

organique.

Page 155: Caractérisation par spectroscopie et analyse compositionnelle des

133

8.2 ABSTRACT

During cultivation, the internal phosphorus cycle of Mollisols (Chernozems) of the

Canadian Prairies is perturbed by crop sequences including wheat phases, tillage practices, and

regular applications of fertilizers. To monitor these changes, a proximate sequential phosphorus

(P) fractionation procedure was developed by Hedley et al. (1982) to extract inorganic and

organic P fractions as very labile (resin-P), labile (NaHCO3-P), slowly available (NaOH-P), and

very slowly available (HCl-P) pools. Models used so far to monitor P pools do not address the

interactive behaviour of P fractions constrained to a closed compositional space. Compositional

data analysis using isometric log ratio (ilr) coordinates is appropriate for modelling the

interactive P pools using sequential binary partitions of P pools. Our objective was to model

changes of P pools in Mollisols in response to management and time using ilr coordinates. We

used a dataset with treatments and another where a Mollisol was analyzed at time zero and 4, 65,

and 90 yr after sod breakup. Seven P fractions were assigned to P reactivity groups to compute

six ilr coordinates. The ilr2 contrasting inorganic (geochemical) and organic (biological) P pools

and ilr4 between the most readily available and less P bioavailable pools were the most sensitive

to crop sequence and fertilization. Using composition at time zero as reference, the Aitchison

distance reached a plateau after the 4th year in the Bm horizon compared to continuous change in

the Ah horizon. Time changed the P balance of cultivated Mollisols primarily in the inorganic vs.

organic P pools. The risks of yield loss and environmental damage can be minimized using soil

tests that quantify the rapidly bioavailable inorganic P pools and crop management strategies that

promote biological P pools.

Page 156: Caractérisation par spectroscopie et analyse compositionnelle des

134

8.3 INTRODUCTION

Mollisols (Chernozems) form an important soil group used for large scale grain

production in the Canadian Prairies. They are naturally fertile soils well supplied with plant

nutrients such as phosphorus. These soils have been perturbed by tillage practices and

applications of phosphatic fertilizers and manure P. As a result, soil and crop management

influenced the internal soil P cycling of Mollisols. Phosphorus fractionation procedures can

quantify the P pools likely to change slowly or rapidly in soils under perturbation.

Hedley et al. (1982) proposed a sequential extraction procedure to chemically assess the

availability of soil organic and inorganic P forms. Cross and Schlesinger (1995) classified the

Hedley et al. (1982) interactive P pools into rapidly plant-available (resin-P and NaHCO3-P) and

refractory (NaOH-P, sonic P, HCl-P, and residual P) pools. The oxalate-extractable P (Pox)

estimates the inorganic P accumulation from fast and slow reactions with iron (Feox) and

aluminium (Alox) hydroxides (Lookman et al., 1996). Inositol phosphates that may account for

more than 50% of soil organic P may also react with Fe and Al compounds in soils (Shang et al.,

1990, 1992; Celi et al., 1999). Refractory and residual P pools contribute little to soil P cycling at

time scale required for soil management. Since organic P usually declines in soils following

cultivation (Stevenson, 1986) chemical fractionation can assess long-term change in P pools in

response to land use or soil management (Frossard et al., 2000).

Raw soil P fractions have been used to describe P distribution and model P dynamics in

soils in state-space (Shuai and Yost, 2004), path (Tiessen et al., 1983), variance, regression and

correlation analyses (e.g. McKenzie et al., 1992; Tiessen et al., 1984). Compositional data such

as raw P fractions have severe limitations for linear modelling since they are constrained to a

close space of strictly positive data spoiled by redundancy and spurious correlations. Gaussian

laws cannot be applied to those data since it is impossible to obtain analytical data less than zero

or more than 100%. For these reasons, P pools bear relative information about pool exchange

processes. Raw P concentration data must be log ratio transformed before analysis by linear

statistical procedures conceived for the real space made of both negative and positive values

(Aitchison, 1986).

Page 157: Caractérisation par spectroscopie et analyse compositionnelle des

135

Isometric log ratio (ilr) transformation with orthonormal bases are sequential binary

partitions (SBP) of compositional data (Egozcue and Pawlowski-Glahn, 2005). For D

phosphorus fractions, there are (D-1) SBP’s. The conceptual model of Tiessen et al. (1984)

shows mass exchange among P pools. A conceptual framework for binary partitions includes

partitions between rapidly and slowly bioavailable P pools (Hedley et al., 1982) and between

geochemical (inorganic) and biological (organic) P pools (Cross and Schlesinger, 1995). The

Aitchison distance can be computed across ilr coordinates as a distance between perturbed and

reference compositions.

The objective of this study is to present the conceptual model that describes relationships

among soil P pools and to decompose these relationships into sequential binary partitions and ilr

coordinates. Time and treatment variations in P pools are analyzed using datasets on the effect of

crop sequence and fertilization on P pools and on time change in P pools in two Mollisols of the

Canadian Prairies.

8.4 MATERIALS AND METHODS

McKenzie et al. (1992) fractionated soil P in dryland grain crop sequences (continuous

wheat, wheat-fallow and wheat-wheat-fallow) on a Lethbridge sandy clay loam (Calcareous Dark

Brown Chernozem) given nitrogen and/or phosphorus fertilizers for 14 to 19 years. On the other

hand, time change in P pools balance was modelled by Tiessen et al. (1983) in a Blaine Lake silt

loam (Orthic Black Chernozemic) following chronosequence starting with native prairie (time

zero) over 4-, 60- and 90-years of crop sequences.

The P pools were quantified using a modified Hedley et al. (1982) procedure. Sonic

P pools that account for a small fraction of total P pools were amalgamated with their

respective P pools. Residue inorganic and organic P pools that are undefined P pools were

amalgamated into a single residue P pool. A modified Tiessen et al. (1984) conceptual

model relating P pools is presented in Figure 6.1. There geochemical and biological P pools

(Cross and Schlesingere, 1995) can be partitioned into slowly and rapidly bioavailable P

pools (Hedley et al., 1982). Residue P represents the slowly bioavailable P pools as

illustrated in Tiessen et al. (1984). The main P pools for partitions were thus the

geochemical, biological, and residue P pools. The first sequential binary partition (SBP)

Page 158: Caractérisation par spectroscopie et analyse compositionnelle des

136

was between residue P and other P fractions (Table 6.1). The second partition was between

the geochemical and biological P pools. In the geochemical pool, there was only one

unidirectional transformation, the one from primary minerals to solution P, leading to the

third partition. Other P pools are bidirectional between rapid or slow geochemical or

biological species (Figure 6.1).

8.4.1 Isometric log ratio transformation and the Aitchison distance

A D-part composition can be described by its parts as follows

(Aitchison, 1986):

(1)

Where is the closure operator to unit . The isometric log ratio coordinate is

computed from SBP’s as follows (Egozcue and Pawlowski-Glahn, 2005):

)(

)(ln*

xg

xg

sr

rsxi (2)

Where is the geometric mean of P fractions in group and is the

geometric mean of P fractions in group . The ilr sign indicates in what direction P pools

change in response to treatment or over time. The Aitchison distance between two

compositions is computed as follows across D-1 compositional dimensions (Egozcue and

Pawlowsky-Glanh, 2006):

(3)

(4)

Page 159: Caractérisation par spectroscopie et analyse compositionnelle des

137

Where is the reference composition.

8.4.2 The Mackenzie et al. (1992) dataset

The response of Mollisol P pools to perturbation by cultural practices and by nitrogen (N)

and phosphorus (P) fertilization regimes investigated by McKenzie et al. (1992) is presented in

Table 6.2. In absolute terms, the P fertilization influenced more markedly the resin and NaHCO3,

and NaOH pools compared to other pools.

The effect sizes of treatments on P pools are reflected by ilr values (Table 6.3). The effect

size of added P varied with fertilization and crop sequence. The effect of added P was most

prominent in ilr2, ilr3 and ilr4 for continuous wheat and the wheat-wheat-fallow sequence while

treatment effects were much smaller in the wheat-fallow sequence (Table 3). The dominance of

wheat in the sequence affected markedly the balance between geochemical and biological pools

and that between P in primary minerals and other inorganic pools.

A clearer picture of P pools change is given by subtracting from the effect of treatments

that of no cultivation (Table 6.4). As shown by the Aitchison distance (Table 6.4), check and

added N were closest to uncultivated conditions across crop sequences, especially where wheat

was dominant. Cultivation, crop sequence and fertilization increased the inorganic and organic

pools compared to residue P, indicating more P bioavailability. The geochemical pool increased

over the biological one across crop sequences (ilr2). The slowly and rapidly bioavailable

inorganic P pools largely increased compared to P from primary minerals (ilr3) under continuous

wheat and wheat-wheat-fallow compared to the wheat-fallow sequences. In general, the balance

between slowly and rapidly bioavailable inorganic P pools (ilr5) slightly decreased compared to

the uncultivated reference composition. Differences were also small in the balance between

slowly and rapidly bioavailable organic P pools (ilr6) and the uncultivated reference

composition.

The ilr4 is the most important balance for plant nutrition and the environment. Where no

P was added, the wheat-fallow sequence showed the highest P bioavailability but also the highest

risk for eutrophication of surface waters by dissolved P. Where P was added, crop sequences

showed similar ilr4 values and Aitchison distances. These results indicate that routine soil tests

representative of ilr4 are useful tools to address both agronomic and environmental issues.

Page 160: Caractérisation par spectroscopie et analyse compositionnelle des

138

However, the biological dimension of the system (ilr2) can be manipulated on the long run by

crop production systems that can decrease ilr4 while maintaining the balance between organic

pools (ilr6).

8.4.3 The Tiessen et al. (1983) dataset

Time change of P pools in horizons Ah and Bm has been studied by Tiessen et al. (1983)

in a Mollisol after breakup of the natural prairie ecosystem followed by cultivation for 90 years

(Table 6.5). The most prominent change was a decrease in the biological pools in both soil

horizons.

After computing ilr coordinates at each time step, the degree of change in P balances can

be measured as changes in each ilr coordinate and globally as the Aitchison distance (Figure 2).

The slowly and rapidly bioavailable P pools decreased in both horizons compared to residue P

(ilr1) and primary minerals P (ilr3) just after sod breakup and varied chaotically thereafter. The

inorganic P pools decreased then increased (ilr2) probably due to humification following sod

breakup and to cultivation thereafter. Resin P increased as a result of mineralization of root

organic matter in the Bm horizon and decreased generally as a result of crop P uptake. Rapidly

bioavailable inorganic (ilr5) and organic (ilr6) P pools decreased compared to slowly

bioavailable ones probably as a result or lesser microbial turnover following cultivation.

Above all, the Aitchison distance across ilr coordinates increased more rapidly in the Bm

than the Ah horizons due to more rapid change in the most readily and the most slowly available

P pools (Table 6.5). Thereafter, P balances remained stationary in the Bm horizon while there

was considerable depletion of organic and residue P pools in the Ah horizon. Hence, few years

after sod breakup, major changes in P pools occurred in the Ah horizon of this Mollisol where

soil conservation practices must have largest effect on plant nutrition and environmental quality.

8.5 CONCLUSION

Using isometric log ratios to model P dynamics, the effect of crop sequences and

fertilization on Mollisol P pools was primarily related to the dominance of wheat in the sequence

and to P fertilization. More crops and less fallow in the sequence maintained soil P pools closer

to uncultivated conditions.

Page 161: Caractérisation par spectroscopie et analyse compositionnelle des

139

The change of P pools over time was shown by an early change in balance where slowly

to rapidly bioavailable P forms generally decreased compared to the recalcitrant residue and

primary minerals P. Overall, the organic P pools decreased in the long run in both Ah and Bm

horizons. Time change started rapidly then stabilized in the Bm horizon. The soil monotonically

moved away from native conditions in the Ah horizon primarily due to loss of organic P.

Page 162: Caractérisation par spectroscopie et analyse compositionnelle des

140

8.6 REFERENCE

Aitchison, J. (1986). The Statistical Analysis of Compositional Data. Monographs on Statistics and

Applied Probability. Chapman & Hall Ltd., London (UK). 416 p.

Celi, L., S. Lamacchia, F. A. Marsan, and E. Barberis (1999). Interaction of inositol hexaphosphate on

clays: adsorption and charging phenomena. Soil Science 164, 574-484.

Cross, A. F. and W. H. Schlesinger (1995). A literature review and evaluation of the Hedley

fractionation: Applications to the biogeochemical cycle of soil phosphorus in natural ecosystems.

Geoderma 64, 197-214.

Egozcue, J. J. and V. Pawlowski-Glahn (2006). Simplicial geometry for compositional data. p. 145-

159 in A Buccianti, G Mateu-Figueras, V Pawlowski-Glahn (eds) Compositional data analysis in the

geosciences: from theory to practice. Geol. Soc., London, Spec. Publ. 264.

Egozcue, J. J. and V. Pawlowski-Glahn (2005). Groups of parts and their balances in compositional

data analysis. Mathematical Geology 37, 795-828.

Frossard, E., L. M. Condron, A. Oberson, S. Sinaj, and J. C. Fardeau (2000). Processes governing

phosphorus availability in temperate soils. Journal of Environmental Quality 29, 15-23.

Hedley, M. J., J. W. B. Stewart, and B. S. Chauhan (1982). Changes in inorganic and organic soil

phosphorus fractions by cultivation practices and by laboratory incubations. Soil Science Society of

America Journal 46, 970-976.

Lookman, R., K. Jansen, R. Merckx, and K. Vlassak. (1996). Relationship between soil properties and

phosphate saturation parameters a transect study in northern Belgium. Geoderma 69 (3-4), 265-274.

McKenzie, R. H., J. W. B. Stewart, J. F. Dormaar, and G. B. Schaalje (1992). Long-term crop rotation

and fertilizer effects on phosphorus transformations: I. In a Chernozemic soil. Canadian Journal of

Soil Science 72, 569-579.

Shang, C., J. W. B. Stewart, and P. M. Huang (1992). pH effect on kinetics of adsorption of organic

and inorganic phosphates by short-range ordered aluminum and iron precipitates. Geoderma 53, 1-

14.

Shang, C., P. M. Huang, and J. W. B. Stewart (1990). Kinetics of adsorption of organic and inorganic

phosphates by short-range ordered precipitate of aluminium. Canadian Journal of Soil Science 70,

461-470.

Shuai, X. and R. S. Yost (2004). State-space modeling to simplify soil phosphorus fractionation. Soil

Science Society of America Journal 68, 1437-1444.

Stevenson, F. J. (1986). Cycles of soils. Carbon, nitrogen, phosphorus, sulfur, micronutrients. Wiley-

Interscience, NY. 380 p.

Tiessen, H., J. W. B. Stweart, and C. V. Cole (1984). Pathways of Phosphorus Transformations in

Soils of Differing Pedogenesis. Soil Science Society of America Journal 48, 853-858.

Tiessen, H., J. W. B. Stewart, and J. O. Moir (1983). Changes in organic and inorganic phosphorus

composition of two grassland soils and their particle size fractions during 60-90 years of cultivation.

Journal of Soil Science 34, 815-823.

Page 163: Caractérisation par spectroscopie et analyse compositionnelle des

141

Table 8-1 Sequential binary partitions of soil P fractions (r is number of P fractions with plus

sign and s is number of P fractions with minus sign).

ilr Inorganic P Organic P Residue r s Balance

coefficient Resin NaHCO3 NaOH H2SO4 NaHCO3 NaOH

1 1 1 1 1 1 1 -1 6 1 0.926

2 1 1 1 1 -1 -1 0 4 2 1.155

3 1 1 1 -1 0 0 0 3 1 0.866

4 1 -1 -1 0 0 0 0 1 2 0.816

5 0 1 -1 0 0 0 0 1 1 0.707

6 0 0 0 0 1 -1 0 1 1 0.707

Page 164: Caractérisation par spectroscopie et analyse compositionnelle des

142

Table 8-2 Mollisol P fractions following crop sequence and NP fertilization (data from

McKenzie et al., 1992).

Treatment Inorganic P fractions Organic P fractions Residue P

Resin NaHCO3 NaOH HCl NaHCO3 NaOH

mg kg-1

Uncultivated control

None 8 6 12 218 7.5 73 254

Continuous wheat

Check (no N nor P) 19 8 20 201 4.4 52 219

Added N 15 8 22 195 8.9 65 215

Added N and P 78 26 44 215 9.5 66 220

Added P 73 19 36 212 6.3 62 223

Wheat-wheat-fallow sequence

Check (no N nor P) 19 8 23 199 5.0 53 214

Added N 15 7 21 203 5.1 62 218

Added N and P 51 16 32 216 4.9 57 222

Added P 61 15 30 213 4.4 50 220

Wheat -fallow sequence

Check (no N nor P) 43 14 36 217 7.8 59 210

Added N 37 13 33 212 8.4 61 213

Added N and P 59 18 38 221 9.0 64 213

Added P 68 17 39 226 6.8 57 215

Page 165: Caractérisation par spectroscopie et analyse compositionnelle des

143

Table 8-3 Ilr coordinates of P pools following crop sequence and fertilization (data from

McKenzie et al., 1992).

Treatment Ilr1 Ilr2 Ilr3 Ilr4 Ilr5 Ilr6

Uncultivated control

None -2.342 -0.251 -2.828 -0.048 -0.490 -1.609

Continuous wheat

Check (no N nor P) -2.095 0.709 -2.278 0.332 -0.648 -1.746

Added N -1.961 0.124 -2.292 0.100 -0.715 -1.406

Added N and P -1.412 1.122 -1.361 0.682 -0.372 -1.371

Added P -1.589 1.224 -1.516 0.838 -0.452 -1.617

Wheat-wheat-fallow sequence

Check (no N nor P) -2.031 0.662 -2.229 0.275 -0.747 -1.669

Added N -2.089 0.433 -2.379 0.174 -0.777 -1.766

Added N and P -1.734 1.236 -1.719 0.664 -0.490 -1.735

Added P -1.757 1.384 -1.693 0.862 -0.490 -1.719

Wheat -fallow sequence

Check (no N nor P) -1.634 0.895 -1.777 0.531 -0.668 -1.431

Added N -1.682 0.736 -1.847 0.474 -0.659 -1.402

Added N and P -1.513 0.950 -1.613 0.664 -0.528 -1.387

Added P -1.563 1.217 -1.601 0.793 -0.587 -1.503

Page 166: Caractérisation par spectroscopie et analyse compositionnelle des

144

Table 8-4 Ilr differences in P pools between treatments and uncultivated check (data from

McKenzie et al., 1992).

Treatment Ilr1 Ilr2 Ilr3 Ilr4 Ilr5 Ilr6

Aitchison

distance

Distance from uncultivated control

Continuous wheat

Check (no N nor P) 0.247 0.961 0.551 0.380 -0.158 -0.137 1.475

Added N 0.381 0.376 0.536 0.148 -0.225 0.203 0.687

Added N and P 0.930 1.373 1.468 0.730 0.118 0.238 5.510

Added P 0.753 1.475 1.312 0.886 0.038 -0.008 5.251

Wheat-wheat-fallow sequence

Check (no N nor P) 0.311 0.913 0.600 0.323 -0.257 -0.060 1.464

Added N 0.253 0.684 0.449 0.222 -0.287 -0.157 0.890

Added N and P 0.608 1.487 1.109 0.712 0.000 -0.126 4.333

Added P 0.585 1.635 1.136 0.911 0.000 -0.110 5.146

Wheat-fallow sequence

Check (no N nor P) 0.708 1.146 1.051 0.579 -0.178 0.178 3.319

Added N 0.660 0.988 0.981 0.522 -0.169 0.207 2.718

Added N and P 0.829 1.201 1.215 0.712 -0.038 0.222 4.164

Added P 0.779 1.468 1.227 0.841 -0.097 0.106 4.998

Page 167: Caractérisation par spectroscopie et analyse compositionnelle des

145

Table 8-5 Effect of time on P pools in a Mollisol (data from Tiessen et al., 1984).

Time Inorganic P Organic P Residue P

Resin NaHCO3 NaOH H2SO4 NaHCO3 NaOH

year mg kg-1

Ah horizon

0 25.8 13.5 31.0 174 49.5 167 337

4 22.5 13.4 32.7 177 52.0 177 416

65 14.8 10.4 32.6 200 31.7 124 318

90 21.6 11.2 32.0 196 19.5 75 273

Bm horizon

0 9.1 5.3 18.4 199 16.6 63 236

4 8.0 3.3 18.4 190 16.5 72 290

65 6.3 4.2 22.8 221 15.1 66 242

90 7.1 3.9 17.9 228 9.7 47 254

Page 168: Caractérisation par spectroscopie et analyse compositionnelle des

146

Figure 8-1 Conceptual relational model between P pools in Mollisols (modified from

Tiessen et al., 1984).

Page 169: Caractérisation par spectroscopie et analyse compositionnelle des

147

0.000

0.200

0.400

0.600

0.800

1.000

1.200

0 20 40 60 80 100

Time elapsed until breakup (year)

Ait

chis

on d

ista

nce

Ah

Bm

Figure 8-2 Time change in P balance distances from initial conditions in a Blaine lake soil

(data from Tiessen et al., 1983).