Sol – Wikipédia

 Sol – Wikipédia

mélange de matière organique, de minéraux, de gaz, de liquides et d’organismes qui, ensemble, soutiennent la vie

Ceci est un diagramme et une photographie connexe des couches de sol du substratum rocheux au sol.

Sol est un mélange de matière organique, de minéraux, de gaz, de liquides et d’organismes qui, ensemble, soutiennent la vie. Le corps du sol de la Terre, appelé pédosphère, a quatre fonctions importantes:

Toutes ces fonctions, à leur tour, modifient le sol et ses propriétés.

Le sol est aussi communément appelé Terre ou alors saleté; certaines définitions scientifiques distinguent saleté de sol en restreignant le premier terme spécifiquement au sol déplacé.

La pédosphère s’interface avec la lithosphère, l’hydrosphère, l’atmosphère et la biosphère.[1] Le terme pédolithe, couramment utilisé pour désigner le sol, se traduit par terre Pierre dans le sens pierre fondamentale, du grec ancien πέδον terre, Terre. Le sol se compose d’une phase solide de minéraux et de matière organique (la matrice du sol), ainsi que d’une phase poreuse qui contient des gaz (l’atmosphère du sol) et de l’eau (la solution du sol).[2][3] En conséquence, les pédologues peuvent envisager les sols comme un système à trois états de solides, de liquides et de gaz.[4]

Le sol est le produit de plusieurs facteurs: l’influence du climat, du relief (élévation, orientation et pente du terrain), des organismes et des matériaux d’origine du sol (minéraux d’origine) interagissant avec le temps.[5] Il subit continuellement un développement par le biais de nombreux processus physiques, chimiques et biologiques, qui comprennent l’altération et l’érosion associée. Compte tenu de sa complexité et de sa forte connectivité interne, les écologistes des sols considèrent le sol comme un écosystème.[6]

La plupart des sols ont une densité apparente sèche (densité du sol prenant en compte les vides à l’état sec) comprise entre 1,1 et 1,6 g / cm3, alors que la densité des particules du sol est beaucoup plus élevée, de l’ordre de 2,6 à 2,7 g / cm3.[7] Peu de sol de la planète Terre est plus ancien que le Pléistocène et aucun n’est plus ancien que le Cénozoïque,[8] bien que les sols fossilisés soient préservés d’aussi loin que l’Archéen.[9]

La science du sol a deux branches d’étude de base: l’édaphologie et la pédologie. Édaphologie étudie l’influence des sols sur les êtres vivants.[10]Pédologie se concentre sur la formation, la description (morphologie) et la classification des sols dans leur environnement naturel.[11] En termes d’ingénierie, le sol est inclus dans le concept plus large de régolithe, qui comprend également d’autres matériaux meubles qui se trouvent au-dessus du substrat rocheux, comme on peut le trouver sur la Lune et sur d’autres objets célestes.[12]

Processus[[[[Éditer]

Le sol fonctionne comme une composante majeure de l’écosystème terrestre. Les écosystèmes du monde sont profondément affectés par les processus mis en œuvre dans le sol, avec des effets allant de l’appauvrissement de la couche d’ozone et du réchauffement climatique à la destruction de la forêt tropicale et à la pollution de l’eau. En ce qui concerne le cycle du carbone de la Terre, le sol agit comme un important réservoir de carbone,[13] et c’est potentiellement l’un des plus réactifs aux perturbations humaines[14] et le changement climatique.[15] À mesure que la planète se réchauffe, il a été prédit que les sols ajouteront du dioxyde de carbone à l’atmosphère en raison d’une activité biologique accrue à des températures plus élevées, une rétroaction positive (amplification).[16] Cette prédiction a cependant été remise en question en tenant compte des connaissances plus récentes sur le renouvellement du carbone du sol.[17]

Le sol agit comme un milieu d’ingénierie, un habitat pour les organismes du sol, un système de recyclage des nutriments et des déchets organiques, un régulateur de la qualité de l’eau, un modificateur de la composition atmosphérique et un milieu pour la croissance des plantes, ce qui en fait un fournisseur extrêmement important de services écosystémiques. .[18] Étant donné que le sol a une vaste gamme de niches et d’habitats disponibles, il contient la majeure partie de la diversité génétique de la Terre. Un gramme de sol peut contenir des milliards d’organismes, appartenant à des milliers d’espèces, pour la plupart microbiennes et encore largement inexplorées.[19][20] Le sol a une densité procaryote moyenne d’environ 108 organismes par gramme,[21] alors que l’océan n’a pas plus de 107 organismes procaryotes par millilitre (gramme) d’eau de mer.[22]Le carbone organique contenu dans le sol est finalement renvoyé dans l’atmosphère par le processus de respiration effectué par des organismes hétérotrophes, mais une partie substantielle est retenue dans le sol sous forme de matière organique du sol; le travail du sol augmente généralement le taux de respiration du sol, conduisant à l’épuisement de la matière organique du sol.[23] Puisque les racines des plantes ont besoin d’oxygène, l’aération est une caractéristique importante du sol. Cette ventilation peut être réalisée via des réseaux de pores du sol interconnectés, qui absorbent et retiennent également l’eau de pluie, la rendant facilement disponible pour l’absorption par les plantes. Étant donné que les plantes ont besoin d’un approvisionnement en eau presque continu, mais que la plupart des régions reçoivent des précipitations sporadiques, la capacité de rétention d’eau des sols est vitale pour la survie des plantes.[24]

Les sols peuvent éliminer efficacement les impuretés,[25] tuer les agents pathogènes,[26] et dégrader les contaminants, cette dernière propriété étant appelée atténuation naturelle.[27] En règle générale, les sols maintiennent une absorption nette d’oxygène et de méthane et subissent un rejet net de dioxyde de carbone et d’oxyde nitreux.[28] Les sols offrent aux plantes un soutien physique, de l’air, de l’eau, une modération de la température, des nutriments et une protection contre les toxines.[29] Les sols fournissent des nutriments facilement disponibles aux plantes et aux animaux en convertissant la matière organique morte en diverses formes de nutriments.[30]

Composition[[[[Éditer]

Composants d’un sol limoneux en pourcentage de volume

Eau (25%)

Gaz (25%)

Sable (18%)

Limon (18%)

Argile (9%)

Matière organique (5%)

Un sol typique contient environ 50% de solides (45% de matières minérales et 5% de matière organique) et 50% de vides (ou pores) dont la moitié est occupée par l’eau et l’autre moitié par le gaz.[31] Le pourcentage de contenu minéral et organique du sol peut être traité comme une constante (à court terme), tandis que le pourcentage de teneur en eau et en gaz du sol est considéré comme très variable, de sorte qu’une augmentation de l’un est simultanément compensée par une réduction de l’autre.[32] L’espace poreux permet l’infiltration et le mouvement de l’air et de l’eau, qui sont tous deux essentiels à la vie dans le sol.[33]Le compactage, un problème courant avec les sols, réduit cet espace, empêchant l’air et l’eau d’atteindre les racines des plantes et les organismes du sol.[34]

Avec suffisamment de temps, un sol indifférencié fera évoluer un profil de sol qui se compose de deux ou plusieurs couches, appelées horizons de sol. Celles-ci diffèrent par une ou plusieurs propriétés telles que leur texture, leur structure, leur densité, leur porosité, leur consistance, leur température, leur couleur et leur réactivité.[8] Les horizons diffèrent grandement en épaisseur et manquent généralement de frontières nettes; leur développement dépend du type de matériau d’origine, des processus qui modifient ces matériaux d’origine et des facteurs de formation du sol qui influencent ces processus. Les influences biologiques sur les propriétés du sol sont les plus fortes près de la surface, tandis que les influences géochimiques sur les propriétés du sol augmentent avec la profondeur. Les profils de sols matures comprennent généralement trois horizons maîtres de base: A, B et C. Le solum comprend normalement les horizons A et B. La composante vivante du sol est en grande partie confinée au solum, et est généralement plus proéminente dans l’horizon A. Il a été suggéré que le pédon, une colonne de sol s’étendant verticalement de la surface au matériau d’origine sous-jacent et suffisamment grande pour montrer les caractéristiques de tous ses horizons, pourrait être subdivisée en humipédon (la partie vivante, où résident la plupart des organismes du sol, correspondant à la forme d’humus), les copedon (en position intermédiaire, là où la plupart des altérations des minéraux ont lieu) et le lithopédon (en contact avec le sous-sol).[36]

La texture du sol est déterminée par les proportions relatives des particules individuelles de sable, de limon et d’argile qui composent le sol. L’interaction des particules minérales individuelles avec la matière organique, l’eau, les gaz via des processus biotiques et abiotiques amène ces particules à floculer (coller ensemble) pour former des agrégats ou des peds.[37] Lorsque ces agrégats peuvent être identifiés, un sol peut être considéré comme développé et peut être décrit plus en détail en termes de couleur, de porosité, de consistance, de réaction (acidité), etc.

L’eau est un agent critique dans le développement du sol en raison de son implication dans la dissolution, la précipitation, l’érosion, le transport et le dépôt des matériaux qui composent un sol.[38] Le mélange d’eau et de matières dissoutes ou en suspension qui occupent l’espace poreux du sol s’appelle la solution du sol. Étant donné que l’eau du sol n’est jamais de l’eau pure, mais qu’elle contient des centaines de substances organiques et minérales dissoutes, elle peut être appelée plus précisément la solution du sol. L’eau est essentielle à la dissolution, à la précipitation et au lessivage des minéraux du profil du sol. Enfin, l’eau affecte le type de végétation qui pousse dans un sol, ce qui à son tour affecte le développement du sol, une rétroaction complexe qui est illustrée dans la dynamique des modèles de végétation en bandes dans les régions semi-arides.[39]

Les sols fournissent aux plantes des nutriments, dont la plupart sont maintenus en place par des particules d’argile et de matière organique (colloïdes)[40] Les nutriments peuvent être adsorbés sur des surfaces minérales argileuses, liés dans des minéraux argileux (absorbés) ou liés dans des composés organiques en tant que partie des organismes vivants ou de la matière organique du sol mort. Ces nutriments liés interagissent avec l’eau du sol pour tamponner la composition de la solution du sol (atténuer les changements dans la solution du sol) lorsque les sols se mouillent ou se dessèchent, que les plantes absorbent les nutriments, que les sels sont lessivés ou que des acides ou des alcalis sont ajoutés.[41][42]

La disponibilité des éléments nutritifs des plantes est affectée par le pH du sol, qui est une mesure de l’activité des ions hydrogène dans la solution du sol. Le pH du sol est fonction de nombreux facteurs de formation du sol et est généralement plus bas (plus acide) là où l’altération est plus avancée.[43]

La plupart des éléments nutritifs des plantes, à l’exception de l’azote, proviennent des minéraux qui composent la matière mère du sol. Une partie de l’azote provient de la pluie sous forme d’acide nitrique dilué et d’ammoniac,[44] mais la majeure partie de l’azote est disponible dans les sols en raison de la fixation de l’azote par les bactéries. Une fois dans le système sol-plante, la plupart des nutriments sont recyclés par les organismes vivants, les résidus végétaux et microbiens (matière organique du sol), les formes liées aux minéraux et la solution du sol. Les organismes vivants du sol (microbes, animaux et racines de plantes) et la matière organique du sol sont d’une importance cruciale pour ce recyclage, et donc pour la formation et la fertilité des sols.[45] Les enzymes microbiennes du sol peuvent libérer des nutriments à partir de minéraux ou de matière organique destinés à être utilisés par les plantes et autres micro-organismes, les séquestrer (les incorporer) dans des cellules vivantes, ou provoquer leur perte du sol par volatilisation (perte dans l’atmosphère sous forme de gaz) ou par lessivage.[46]

Formation[[[[Éditer]

La formation du sol, ou pédogenèse, est l’effet combiné de processus physiques, chimiques, biologiques et anthropiques agissant sur le matériau d’origine du sol. On dit que le sol se forme lorsque la matière organique s’est accumulée et que les colloïdes sont lavés vers le bas, laissant des dépôts d’argile, d’humus, d’oxyde de fer, de carbonate et de gypse, produisant une couche distincte appelée horizon B. Il s’agit d’une définition quelque peu arbitraire car des mélanges de sable, de limon, d’argile et d’humus soutiendront l’activité biologique et agricole avant cette date.[47] Ces constituants sont déplacés d’un niveau à un autre par l’eau et l’activité animale. En conséquence, des couches (horizons) se forment dans le profil du sol. L’altération et le mouvement des matériaux dans un sol provoquent la formation d’horizons de sol distinctifs. Cependant, les définitions plus récentes du sol englobent les sols sans aucune matière organique, tels que les régolithes qui se sont formés sur Mars.[48] et des conditions analogues dans les déserts de la planète Terre.[49]

Un exemple de développement d’un sol commencerait par l’altération du substrat rocheux coulé de lave, qui produirait le matériau d’origine purement minéral à partir duquel se forme la texture du sol. Le développement du sol se ferait le plus rapidement à partir de la roche nue de flux récents dans un climat chaud, sous des pluies abondantes et fréquentes. Dans ces conditions, les plantes (dans un premier temps les lichens et cyanobactéries fixateurs d’azote puis les plantes épilithiques supérieures) s’établissent très rapidement sur la lave basaltique, même s’il y a très peu de matière organique.[50] Les plantes sont soutenues par la roche poreuse car elle est remplie d’eau nutritive qui transporte les minéraux dissous des roches. Les crevasses et les poches, la topographie locale des roches, contiendraient des matériaux fins et hébergeraient des racines de plantes. Les racines des plantes en développement sont associées à des champignons mycorhiziens altérant les minéraux[51] qui aident à briser la lave poreuse, et par ces moyens la matière organique et un sol minéral plus fin s’accumulent avec le temps. Ces étapes initiales de développement du sol ont été décrites sur les volcans,[52] inselbergs,[53] et moraines glaciaires.[54]

Le déroulement de la formation du sol est influencé par au moins cinq facteurs classiques qui sont étroitement liés dans l’évolution d’un sol. Ce sont: le matériau d’origine, le climat, la topographie (relief), les organismes et le temps.[55] Lorsqu’ils sont réorganisés selon le climat, le relief, les organismes, le matériau d’origine et le temps, ils forment l’acronyme CROPT.[56]

Propriétés physiques[[[[Éditer]

Les propriétés physiques des sols, par ordre d’importance décroissante pour les services écosystémiques tels que la production végétale, sont la texture, la structure, la densité apparente, la porosité, la consistance, la température, la couleur et la résistivité.[57] La texture du sol est déterminée par la proportion relative des trois types de particules minérales du sol, appelées sol sépare: le sable, le limon et l’argile. À la prochaine plus grande échelle, les structures du sol appelées peds ou plus communément agrégats de sol sont créés à partir du sol se sépare lorsque les oxydes de fer, les carbonates, l’argile, la silice et l’humus enrobent les particules et les font adhérer à des structures secondaires plus grandes et relativement stables.[58] La densité apparente du sol, lorsqu’elle est déterminée dans des conditions d’humidité normalisées, est une estimation du compactage du sol.[59] La porosité du sol se compose de la partie vide du volume du sol et est occupée par les gaz ou l’eau. La consistance du sol est la capacité des matériaux du sol à adhérer ensemble. La température et la couleur du sol sont auto-définies. La résistivité fait référence à la résistance à la conduction des courants électriques et affecte le taux de corrosion des structures métalliques et en béton enfouies dans le sol.[60] Ces propriétés varient selon la profondeur d’un profil de sol, c’est-à-dire à travers les horizons du sol. La plupart de ces propriétés déterminent l’aération du sol et la capacité de l’eau à s’infiltrer et à être retenue dans le sol.[61]

Humidité du sol[[[[Éditer]

Humidité du sol fait référence à la teneur en eau du sol. Il peut être exprimé en volume ou en poids. La mesure de l’humidité du sol peut être basée sur in situ sondes (p. ex. sondes capacitives, sondes à neutrons) ou méthodes de télédétection.

L’eau qui pénètre dans un champ est éliminée d’un champ par ruissellement, drainage, évaporation ou transpiration.[62] Le ruissellement est l’eau qui coule à la surface jusqu’au bord du champ; le drainage est l’eau qui coule à travers le sol vers le bas ou vers le bord du champ sous terre; la perte d’eau par évaporation d’un champ est la partie de l’eau qui s’évapore dans l’atmosphère directement à partir de la surface du champ; la transpiration est la perte d’eau du champ par son évaporation de la plante elle-même.

L’eau affecte la formation, la structure, la stabilité et l’érosion du sol, mais est une préoccupation majeure en ce qui concerne la croissance des plantes.[63] L’eau est essentielle aux plantes pour quatre raisons:

  1. Il constitue 80 à 95% du protoplasme de la plante.
  2. C’est essentiel pour la photosynthèse.
  3. C’est le solvant dans lequel les nutriments sont transportés vers, dans et à travers la plante.
  4. Il fournit la turgescence grâce à laquelle la plante se maintient en bonne position.

De plus, l’eau modifie le profil du sol en dissolvant et en re-déposant des solutés et colloïdes minéraux et organiques, souvent à des niveaux inférieurs, un processus appelé lessivage. Dans un sol limoneux, les solides constituent la moitié du volume, le gaz un quart du volume et l’eau un quart du volume dont seulement la moitié sera disponible pour la plupart des plantes, avec une forte variation selon le potentiel matriciel.[65]

Un champ inondé drainera l’eau gravitationnelle sous l’influence de la gravité jusqu’à ce que les forces adhésives et cohésives de l’eau résistent à un drainage supplémentaire, point auquel on dit qu’il a atteint la capacité du champ. À ce stade, les plantes doivent appliquer une aspiration pour puiser de l’eau dans un sol. L’eau que les plantes peuvent puiser dans le sol s’appelle l’eau disponible. Une fois que l’eau disponible est épuisée, l’humidité restante est appelée eau non disponible car la plante ne peut pas produire une aspiration suffisante pour aspirer cette eau. À 15 bar d’aspiration, point de flétrissement, les graines ne germent pas, les plantes commencent à se flétrir puis meurent à moins qu’elles ne le soient. capable de récupérer après le réapprovisionnement en eau grâce à des adaptations spécifiques à l’espèce.[71] L’eau se déplace dans le sol sous l’influence de la gravité, de l’osmose et de la capillarité.[72] Lorsque l’eau pénètre dans le sol, elle déplace l’air des macropores interconnectés par flottabilité et brise les agrégats dans lesquels l’air est emprisonné, un processus appelé extinction.[73]
La vitesse à laquelle un sol peut absorber l’eau dépend du sol et de ses autres conditions. Au fur et à mesure qu’une plante grandit, ses racines éliminent d’abord l’eau des plus grands pores (macropores). Bientôt, les pores plus grands ne retiennent que l’air, et l’eau restante se trouve uniquement dans les pores de taille intermédiaire et de plus petite taille (micropores). L’eau dans les plus petits pores est si fortement maintenue à la surface des particules que les racines des plantes ne peuvent pas l’arracher. Par conséquent, toute l’eau du sol n’est pas disponible pour les plantes, avec une forte dépendance à la texture.[74] Lorsqu’il est saturé, le sol peut perdre des nutriments à mesure que l’eau s’écoule.[75] L’eau se déplace dans un champ drainant sous l’influence de la pression où le sol est localement saturé et par capillarité tirant vers les parties les plus sèches du sol.[76] La plupart des besoins en eau des plantes sont fournis par l’aspiration causée par l’évaporation des feuilles des plantes (transpiration) et une fraction inférieure est fournie par l’aspiration créée par les différences de pression osmotique entre l’intérieur de la plante et la solution du sol.[77][78] Les racines des plantes doivent chercher de l’eau et pousser de préférence dans des microsites de sol plus humides,[79] mais certaines parties du système racinaire sont également capables de réhumidifier les parties sèches du sol.[80] Une eau insuffisante endommagera le rendement d’une culture.[81] La majeure partie de l’eau disponible est utilisée dans la transpiration pour attirer les nutriments dans la plante.[82]

L’eau du sol est également importante pour la modélisation du climat et la prévision numérique du temps. Le Système mondial d’observation du climat a spécifié l’eau du sol comme l’une des 50 variables climatiques essentielles (ECV).[83] L’eau du sol peut être mesurée in situ avec un capteur d’humidité du sol ou peut être estimée à partir de données satellitaires et de modèles hydrologiques. Chaque méthode présente des avantages et des inconvénients, et par conséquent, l’intégration de différentes techniques peut réduire les inconvénients d’une seule méthode donnée.[84]

Rétention d’eau[[[[Éditer]

L’eau est retenue dans un sol lorsque la force adhésive d’attraction des atomes d’hydrogène de l’eau pour l’oxygène des particules du sol est plus forte que les forces de cohésion ressenties par l’hydrogène de l’eau pour les autres atomes d’oxygène de l’eau. Lorsqu’un champ est inondé, l’espace poreux du sol est complètement rempli d’eau. Le champ se drainera sous la force de la gravité jusqu’à ce qu’il atteigne ce qu’on appelle la capacité du champ, à quel point les plus petits pores sont remplis d’eau et les plus grands d’eau et de gaz.[86] La quantité totale d’eau retenue lorsque la capacité du champ est atteinte est fonction de la surface spécifique des particules de sol.[87] En conséquence, les sols à haute teneur en argile et à haute teneur organique ont des capacités de champ plus élevées.[88] L’énergie potentielle de l’eau par unité de volume par rapport à l’eau pure dans les conditions de référence est appelée potentiel d’eau. Le potentiel hydrique total est une somme du potentiel matriciel qui résulte de l’action capillaire, du potentiel osmotique pour un sol salin et du potentiel gravitationnel lorsqu’il s’agit de la direction verticale du mouvement de l’eau. Le potentiel hydrique dans le sol a généralement des valeurs négatives, et par conséquent il est également exprimé en succion, qui est définie comme le moins du potentiel hydrique. L’aspiration a une valeur positive et peut être considérée comme la force totale requise pour tirer ou pousser l’eau hors du sol. Le potentiel d’eau ou l’aspiration est exprimé en unités de kPa (103pascal), bar (100 kPa) ou cm H2O (environ 0,098 kPa). Logarithme commun d’aspiration en cm H2O s’appelle pF.[89] Par conséquent, pF 3 = 1000 cm = 98 kPa = 0,98 bar.

Les forces avec lesquelles l’eau est retenue dans les sols déterminent sa disponibilité pour les plantes. Les forces d’adhérence retiennent l’eau fortement sur les surfaces minérales et humiques et moins fortement sur elle-même par des forces cohésives. La racine d’une plante peut pénétrer un très petit volume d’eau qui adhère au sol et être initialement capable d’aspirer de l’eau qui n’est que légèrement maintenue par les forces de cohésion. Mais au fur et à mesure que la gouttelette est aspirée, les forces d’adhérence de l’eau pour les particules du sol produisent une aspiration de plus en plus élevée, jusqu’à 1500 kPa (pF = 4,2).[90] À une aspiration de 1500 kPa, la quantité d’eau du sol est appelée point de flétrissement. À cette aspiration, la plante ne peut pas soutenir ses besoins en eau car l’eau est toujours perdue de la plante par la transpiration, la turgescence de la plante est perdue et elle se flétrit, bien que la fermeture stomatique puisse diminuer la transpiration et ainsi retarder le flétrissement en dessous du point de flétrissement, en particulier. sous adaptation ou acclimatation à la sécheresse.[91] Le niveau suivant, appelé séchage à l’air, se produit à une aspiration de 100 000 kPa (pF = 6). Enfin, la condition de séchage à l’étuve est atteinte à une aspiration de 1 000 000 kPa (pF = 7). Toute eau en dessous du point de flétrissement est appelée eau indisponible.

Lorsque la teneur en humidité du sol est optimale pour la croissance des plantes, l’eau des pores de grande et moyenne taille peut se déplacer dans le sol et être facilement utilisée par les plantes.[74] La quantité d’eau restant dans un sol drainé à la capacité du champ et la quantité disponible sont des fonctions du type de sol. Le sol sableux retiendra très peu d’eau, tandis que l’argile en retiendra le maximum.[88] L’eau disponible pour le limon limoneux pourrait être de 20% alors que pour le sable, elle pourrait être de seulement 6% en volume, comme indiqué dans ce tableau.

Point de flétrissement, capacité au champ et eau disponible de différentes textures de sol (unité:% en volume)[93]
Texture du sol Point de flétrissement Capacité sur le terrain Eau disponible
Le sable 3,3 9,1 5,8
loam sableux 9,5 20,7 11.2
Terreau 11,7 27,0 15,3
Limon limoneux 13,3 33,0 19,7
Terreau d’argile 19,7 31,8 12,1
Argile 27,2 39,6 12,4

Les valeurs ci-dessus sont des valeurs moyennes pour les textures du sol.

L’écoulement de l’eau[[[[Éditer]

L’eau se déplace dans le sol sous l’effet de la gravité, de l’osmose et de la capillarité. À une aspiration de zéro à 33 kPa (capacité du champ), l’eau est poussée à travers le sol à partir du point de son application sous la force de gravité et le gradient de pression créé par la pression de l’eau; c’est ce qu’on appelle un débit saturé. À une aspiration plus élevée, le mouvement de l’eau est tiré par capillarité d’un sol plus humide vers un sol plus sec. Ceci est causé par l’adhésion de l’eau aux solides du sol et est appelé écoulement insaturé.[95]

L’infiltration et le mouvement de l’eau dans le sol sont contrôlés par six facteurs:

  1. Texture du sol
  2. Structure du sol. Les sols à texture fine et à structure granulaire sont les plus favorables à l’infiltration d’eau.
  3. La quantité de matière organique. La matière grossière est préférable et si elle est en surface, elle aide à prévenir la destruction de la structure du sol et la création de croûtes.
  4. Profondeur du sol jusqu’aux couches imperméables telles que les dures ou le substrat rocheux
  5. La quantité d’eau déjà présente dans le sol
  6. La température du sol. Les sols chauds absorbent l’eau plus rapidement tandis que les sols gelés peuvent ne pas être capables d’absorber selon le type de gel.

Les taux d’infiltration d’eau varient de 0,25 cm par heure pour les sols à haute teneur en argile à 2,5 cm par heure pour le sable et les structures de sol bien stabilisées et agrégées. L’eau traverse le sol de manière inégale, sous la forme de soi-disant «doigts de gravité», en raison de la tension superficielle entre les particules d’eau.[98][99]

Les racines des arbres, vivants ou morts, créent des canaux préférentiels pour l’écoulement des eaux de pluie dans le sol,[100] grossissant les taux d’infiltration de l’eau jusqu’à 27 fois.[101]

Les inondations augmentent temporairement la perméabilité du sol dans les lits des rivières, aidant à recharger les aquifères.[102]

L’eau appliquée à un sol est poussée par des gradients de pression depuis le point de son application où elle est saturée localement, vers des zones moins saturées, comme la zone vadose.[103][104] Une fois que le sol est complètement mouillé, toute eau supplémentaire se déplace vers le bas ou s’infiltre hors de la gamme des racines des plantes, transportant avec elle de l’argile, de l’humus, des nutriments, principalement des cations et divers contaminants, y compris des pesticides, des polluants, des virus et des bactéries, pouvant causer contamination des eaux souterraines.[105][106] Par ordre de solubilité décroissante, les nutriments lessivés sont:

  • Calcium
  • Magnésium, soufre, potassium; selon la composition du sol
  • Azote; généralement peu, à moins que l’engrais nitrate n’ait été appliqué récemment
  • Phosphore; très peu car ses formes dans le sol sont de faible solubilité.

Aux États-Unis, l’eau de percolation due aux précipitations varie de presque zéro centimètre juste à l’est des montagnes Rocheuses à cinquante centimètres ou plus par jour dans les Appalaches et sur la côte nord du golfe du Mexique.

L’eau est tirée par action capillaire en raison de la force d’adhérence de l’eau aux solides du sol, produisant un gradient d’aspiration du sol humide vers un sol plus sec[109] et des macropores aux micropores.[[[[citation requise] L’équation dite de Richards permet de calculer le taux de changement temporel de la teneur en humidité dans les sols en raison du mouvement de l’eau dans les sols non saturés.[110] Fait intéressant, cette équation attribuée à Richards a été initialement publiée par Richardson en 1922.[111] L’équation de la vitesse de l’humidité du sol,[112] qui peut être résolu en utilisant la méthode d’écoulement de zone de vadose à teneur en eau finie,[113][114] décrit la vitesse de l’écoulement de l’eau à travers un sol non saturé dans le sens vertical. La solution numérique de l’équation de Richardson / Richards permet de calculer le débit d’eau insaturée et le transport de soluté à l’aide d’un logiciel tel que Hydrus,[115] en donnant au sol les paramètres hydrauliques des fonctions hydrauliques (fonction de rétention d’eau et fonction de conductivité hydraulique insaturée) et les conditions initiales et aux limites. Un écoulement préférentiel se produit le long des macropores interconnectés, des crevasses, des canaux de racine et de ver, qui drainent l’eau par gravité.[116][117]
De nombreux modèles basés sur la physique du sol permettent maintenant une représentation de l’écoulement préférentiel comme un double continuum, une double porosité ou une double perméabilité, mais ceux-ci ont généralement été «boulonnés» à la solution de Richards sans aucun fondement physique rigoureux.[118]

Absorption d’eau par les plantes[[[[Éditer]

Les moyens par lesquels les plantes l’acquièrent et leurs nutriments sont tout aussi importants pour le stockage et le mouvement de l’eau dans le sol. La plupart de l’eau du sol est absorbée par les plantes comme absorption passive causée par la force de traction de l’eau qui s’évapore (transpire) de la longue colonne d’eau (flux de sève de xylème) qui mène des racines de la plante à ses feuilles, selon la théorie de la cohésion-tension .[119] Le mouvement ascendant de l’eau et des solutés (élévation hydraulique) est régulé dans les racines par l’endoderme[120] et dans le feuillage végétal par conductance stomatique,[121] et peut être interrompu dans les vaisseaux du xylème des racines et des pousses par cavitation, également appelée embolie au xylème.[122] De plus, la forte concentration de sels dans les racines des plantes crée un gradient de pression osmotique qui pousse l’eau du sol dans les racines.[123] L’absorption osmotique devient plus importante pendant les périodes de faible transpiration de l’eau causée par des températures plus basses (par exemple la nuit) ou une humidité élevée, et l’inverse se produit à des températures élevées ou à une faible humidité. Ce sont ces processus qui provoquent respectivement la guttation et le flétrissement.[125]

L’extension des racines est vitale pour la survie des plantes. Une étude d’une seule plante de seigle d’hiver cultivée pendant quatre mois dans un pied cube (0,0283 mètre cube) de sol limoneux a montré que la plante développait 13 800 000 racines, soit un total de 620 km de longueur et 237 mètres carrés de surface; et 14 milliards de racines de cheveux d’une longueur totale de 10 620 km et d’une superficie totale de 400 mètres carrés; pour une surface totale de 638 m². La superficie totale du sol limoneux a été estimée à 52 000 mètres carrés. En d’autres termes, les racines n’étaient en contact qu’avec 1,2% du sol. Cependant, l’extension des racines doit être considérée comme un processus dynamique, permettant à de nouvelles racines d’explorer un nouveau volume de sol chaque jour, augmentant considérablement le volume total de sol exploré sur une période de croissance donnée, et donc le volume d’eau absorbé par la racine. système sur cette période.[127] L’architecture racinaire, c’est-à-dire la configuration spatiale du système racinaire, joue un rôle prépondérant dans l’adaptation des plantes à l’eau du sol et à la disponibilité des nutriments, et donc dans la productivité des plantes.[128]

Les racines doivent chercher de l’eau car l’écoulement insaturé de l’eau dans le sol ne peut se déplacer qu’à un rythme allant jusqu’à 2,5 cm par jour; en conséquence, ils meurent et grandissent constamment car ils recherchent des concentrations élevées d’humidité du sol.[129] Une humidité insuffisante du sol, au point de provoquer un flétrissement, causera des dommages permanents et les rendements des cultures en souffriront. Lorsque le sorgho-grain a été exposé à une aspiration du sol aussi faible que 1300 kPa pendant l’émergence de l’épi aux stades de floraison et de germination, sa production a été réduite de 34%.

Consommation et efficacité d’utilisation de l’eau[[[[Éditer]

Seule une petite fraction (0,1% à 1%) de l’eau utilisée par une usine est conservée dans l’usine. La majorité est finalement perdue par transpiration, tandis que l’évaporation de la surface du sol est également importante, le rapport transpiration: évaporation variant selon le type de végétation et le climat, atteignant un sommet dans les forêts tropicales humides et plongeant dans les steppes et les déserts.[131] La transpiration et la perte d’humidité du sol par évaporation sont appelées évapotranspiration. L’évapotranspiration plus l’eau contenue dans la plante correspond à une consommation consommatrice, ce qui est presque identique à l’évapotranspiration.[132]

L’eau totale utilisée dans un champ agricole comprend le ruissellement de surface, le drainage et la consommation. The use of loose mulches will reduce evaporative losses for a period after a field is irrigated, but in the end the total evaporative loss (plant plus soil) will approach that of an uncovered soil, while more water is immediately available for plant growth.[133]Water use efficiency is measured by the transpiration ratio, which is the ratio of the total water transpired by a plant to the dry weight of the harvested plant. Transpiration ratios for crops range from 300 to 700. For example, alfalfa may have a transpiration ratio of 500 and as a result 500 kilograms of water will produce one kilogram of dry alfalfa.

Soil gas[[[[edit]

The atmosphere of soil, or soil gas, is very different from the atmosphere above. The consumption of oxygen by microbes and plant roots, and their release of carbon dioxide, decrease oxygen and increase carbon dioxide concentration. Atmospheric CO2 concentration is 0.04%, but in the soil pore space it may range from 10 to 100 times that level, thus potentially contributing to the inhibition of root respiration.[135] Calcareous soils regulate CO2 concentration by carbonate buffering, contrary to acid soils in which all CO2 respired accumulates in the soil pore system.[136] At extreme levels CO2 is toxic.[137] This suggests a possible negative feedback control of soil CO2 concentration through its inhibitory effects on root and microbial respiration (also called ‘soil respiration’).[138] In addition, the soil voids are saturated with water vapour, at least until the point of maximal hygroscopicity, beyond which a vapour-pressure deficit occurs in the soil pore space.[33] Adequate porosity is necessary, not just to allow the penetration of water, but also to allow gases to diffuse in and out. Movement of gases is by diffusion from high concentrations to lower, the diffusion coefficient decreasing with soil compaction.[139] Oxygen from above atmosphere diffuses in the soil where it is consumed and levels of carbon dioxide in excess of above atmosphere diffuse out with other gases (including greenhouse gases) as well as water.[140] Soil texture and structure strongly affect soil porosity and gas diffusion. It is the total pore space (porosity) of soil, not the pore size, and the degree of pore interconnection (or conversely pore sealing), together with water content, air turbulence and temperature, that determine the rate of diffusion of gases into and out of soil.[140]Platy soil structure and soil compaction (low porosity) impede gas flow, and a deficiency of oxygen may encourage anaerobic bacteria to reduce (strip oxygen) from nitrate NO3 to the gases N2, N2O, and NO, which are then lost to the atmosphere, thereby depleting the soil of nitrogen.[142] Aerated soil is also a net sink of methane CH4[143] but a net producer of methane (a strong heat-absorbing greenhouse gas) when soils are depleted of oxygen and subject to elevated temperatures.[144]

Soil atmosphere is also the seat of emissions of volatiles other than carbon and nitrogen oxides from various soil organisms, e.g. roots,[145] bacteria,[146] fungi,[147] animals.[148] These volatiles are used as chemical cues, making soil atmosphere the seat of interaction networks[149][150] playing a decisive role in the stability, dynamics and evolution of soil ecosystems.[151] Biogenic soil volatile organic compounds are exchanged with the aboveground atmosphere, in which they are just 1–2 orders of magnitude lower than those from aboveground vegetation.[152]

We humans can get some idea of the soil atmosphere through the well-known ‘after-the-rain’ scent, when infiltering rainwater flushes out the whole soil atmosphere after a drought period, or when soil is excavated,[153] a bulk property attributed in a reductionist manner to particular biochemical compounds such as petrichor or geosmin.

Solid phase (soil matrix)[[[[edit]

Soil particles can be classified by their chemical composition (mineralogy) as well as their size. The particle size distribution of a soil, its texture, determines many of the properties of that soil, in particular hydraulic conductivity and water potential,[154] but the mineralogy of those particles can strongly modify those properties. The mineralogy of the finest soil particles, clay, is especially important.[155]

Chemistry[[[[edit]

The chemistry of a soil determines its ability to supply available plant nutrients and affects its physical properties and the health of its living population. In addition, a soil’s chemistry also determines its corrosivity, stability, and ability to absorb pollutants and to filter water. It is the surface chemistry of mineral and organic colloids that determines soil’s chemical properties.[156] A colloid is a small, insoluble particle ranging in size from 1 nanometer to 1 micrometer, thus small enough to remain suspended by Brownian motion in a fluid medium without settling.[157] Most soils contain organic colloidal particles called humus as well as the inorganic colloidal particles of clays. The very high specific surface area of colloids and their net electrical charges give soil its ability to hold and release ions. Negatively charged sites on colloids attract and release cations in what is referred to as cation exchange. Cation-exchange capacity (CEC) is the amount of exchangeable cations per unit weight of dry soil and is expressed in terms of milliequivalents of positively charged ions per 100 grams of soil (or centimoles of positive charge per kilogram of soil; cmolc/kg). Similarly, positively charged sites on colloids can attract and release anions in the soil giving the soil anion exchange capacity (AEC).

Cation and anion exchange[[[[edit]

The cation exchange, that takes place between colloids and soil water, buffers (moderates) soil pH, alters soil structure, and purifies percolating water by adsorbing cations of all types, both useful and harmful.

The negative or positive charges on colloid particles make them able to hold cations or anions, respectively, to their surfaces. The charges result from four sources.

  1. Isomorphous substitution occurs in clay during its formation, when lower-valence cations substitute for higher-valence cations in the crystal structure.[159] Substitutions in the outermost layers are more effective than for the innermost layers, as the electric charge strength drops off as the square of the distance. The net result is oxygen atoms with net negative charge and the ability to attract cations.
  2. Edge-of-clay oxygen atoms are not in balance ionically as the tetrahedral and octahedral structures are incomplete.[160]
  3. Hydroxyls may substitute for oxygens of the silica layers, a process called hydroxylation. When the hydrogens of the clay hydroxyls are ionised into solution, they leave the oxygen with a negative charge (anionic clays).[161]
  4. Hydrogens of humus hydroxyl groups may also be ionised into solution, leaving, similarly to clay, an oxygen with a negative charge.[162]

Cations held to the negatively charged colloids resist being washed downward by water and out of reach of plants’ roots, thereby preserving the fertility of soils in areas of moderate rainfall and low temperatures.[163][164]

There is a hierarchy in the process of cation exchange on colloids, as they differ in the strength of adsorption by the colloid and hence their ability to replace one another (ion exchange). If present in equal amounts in the soil water solution:

Al3+ replaces H+ replaces Ca2+ replaces Mg2+ replaces K+ same as NH4+ replaces Na+[165]

If one cation is added in large amounts, it may replace the others by the sheer force of its numbers. This is called law of mass action. This is largely what occurs with the addition of cationic fertilisers (potash, lime).[166]

As the soil solution becomes more acidic (low pH, meaning an abundance of H+, the other cations more weakly bound to colloids are pushed into solution as hydrogen ions occupy exchange sites (protonation). A low pH may cause hydrogen of hydroxyl groups to be pulled into solution, leaving charged sites on the colloid available to be occupied by other cations. This ionisation of hydroxyl groups on the surface of soil colloids creates what is described as pH-dependent surface charges.[167] Unlike permanent charges developed by isomorphous substitution, pH-dependent charges are variable and increase with increasing pH.[42] Freed cations can be made available to plants but are also prone to be leached from the soil, possibly making the soil less fertile.[168] Plants are able to excrete H+ into the soil through the synthesis of organic acids and by that means, change the pH of the soil near the root and push cations off the colloids, thus making those available to the plant.[169]

Cation exchange capacity (CEC)[[[[edit]

Cation exchange capacity should be thought of as the soil’s ability to remove cations from the soil water solution and sequester those to be exchanged later as the plant roots release hydrogen ions to the solution. CEC is the amount of exchangeable hydrogen cation (H+) that will combine with 100 grams dry weight of soil and whose measure is one milliequivalents per 100 grams of soil (1 meq/100 g). Hydrogen ions have a single charge and one-thousandth of a gram of hydrogen ions per 100 grams dry soil gives a measure of one milliequivalent of hydrogen ion. Calcium, with an atomic weight 40 times that of hydrogen and with a valence of two, converts to (40/2) x 1 milliequivalent = 20 milliequivalents of hydrogen ion per 100 grams of dry soil or 20 meq/100 g. The modern measure of CEC is expressed as centimoles of positive charge per kilogram (cmol/kg) of oven-dry soil.

Most of the soil’s CEC occurs on clay and humus colloids, and the lack of those in hot, humid, wet climates, due to leaching and decomposition, respectively, explains the apparent sterility of tropical soils.[171] Live plant roots also have some CEC, linked to their specific surface area.[172]

Cation exchange capacity for soils; soil textures; soil colloids
Soil State CEC meq/100 g
Charlotte fine sand Florida 1.0
Ruston fine sandy loam Texas 1.9
Glouchester loam New Jersey 11.9
Grundy silt loam Illinois 26.3
Gleason clay loam California 31.6
Susquehanna clay loam Alabama 34.3
Davie mucky fine sand Florida 100.8
Sands —— 1–5
Fine sandy loams —— 5–10
Loams and silt loams —– 5–15
Clay loams —– 15–30
Clays —– over 30
Sesquioxides —– 0–3
Kaolinite —– 3–15
Illite —– 25–40
Montmorillonite —– 60–100
Vermiculite (similar to illite) —– 80–150
Humus —– 100–300

Anion exchange capacity (AEC)[[[[edit]

Anion exchange capacity should be thought of as the soil’s ability to remove anions (e.g. nitrate, phosphate) from the soil water solution and sequester those for later exchange as the plant roots release carbonate anions to the soil water solution. Those colloids which have low CEC tend to have some AEC. Amorphous and sesquioxide clays have the highest AEC,[174] followed by the iron oxides. Levels of AEC are much lower than for CEC, because of the generally higher rate of positively (versus negatively) charged surfaces on soil colloids, to the exception of variable-charge soils.[175] Phosphates tend to be held at anion exchange sites.[176]

Iron and aluminum hydroxide clays are able to exchange their hydroxide anions (OH) for other anions.[177] The order reflecting the strength of anion adhesion is as follows:

H2PO4 replaces SO42− replaces NO3 replaces Cl

The amount of exchangeable anions is of a magnitude of tenths to a few milliequivalents per 100 g dry soil. As pH rises, there are relatively more hydroxyls, which will displace anions from the colloids and force them into solution and out of storage; hence AEC decreases with increasing pH (alkalinity).[178]

Reactivity (pH)[[[[edit]

Soil reactivity is expressed in terms of pH and is a measure of the acidity or alkalinity of the soil. More precisely, it is a measure of hydrogen ion concentration in an aqueous solution and ranges in values from 0 to 14 (acidic to basic) but practically speaking for soils, pH ranges from 3.5 to 9.5, as pH values beyond those extremes are toxic to life forms.[179]

At 25 °C an aqueous solution that has a pH of 3.5 has 10−3.5moles H+ (hydrogen ions) per litre of solution (and also 10−10.5 mole/litre OH). A pH of 7, defined as neutral, has 10−7 moles of hydrogen ions per litre of solution and also 10−7 moles of OH per litre; since the two concentrations are equal, they are said to neutralise each other. A pH of 9.5 has 10−9.5 moles hydrogen ions per litre of solution (and also 10−2.5 mole per litre OH). A pH of 3.5 has one million times more hydrogen ions per litre than a solution with pH of 9.5 (9.5–3.5 = 6 or 106) and is more acidic.[180]

The effect of pH on a soil is to remove from the soil or to make available certain ions. Soils with high acidity tend to have toxic amounts of aluminium and manganese.[181] As a result of a trade-off between toxicity and requirement most nutrients are better available to plants at moderate pH,[182] although most minerals are more soluble in acid soils. Soil organisms are hindered by high acidity, and most agricultural crops do best with mineral soils of pH 6.5 and organic soils of pH 5.5. Given that at low pH toxic metals (e.g. cadmium, zinc, lead) are positively charged as cations and organic pollutants are in non-ionic form, thus both made more available to organisms,[184][185] it has been suggested that plants, animals and microbes commonly living in acid soils are pre-adapted to every kind of pollution, whether of natural or human origin.[186]

In high rainfall areas, soils tend to acidify as the basic cations are forced off the soil colloids by the mass action of hydrogen ions from the rain against those attached to the colloids. High rainfall rates can then wash the nutrients out, leaving the soil inhabited only by those organisms which are particularly efficient to uptake nutrients in very acid conditions, like in tropical rainforests.[187] Once the colloids are saturated with H+, the addition of any more hydrogen ions or aluminum hydroxyl cations drives the pH even lower (more acidic) as the soil has been left with no buffering capacity.[188] In areas of extreme rainfall and high temperatures, the clay and humus may be washed out, further reducing the buffering capacity of the soil.[189] In low rainfall areas, unleached calcium pushes pH to 8.5 and with the addition of exchangeable sodium, soils may reach pH 10.[190] Beyond a pH of 9, plant growth is reduced.[191] High pH results in low micro-nutrient mobility, but water-soluble chelates of those nutrients can correct the deficit.[192] Sodium can be reduced by the addition of gypsum (calcium sulphate) as calcium adheres to clay more tightly than does sodium causing sodium to be pushed into the soil water solution where it can be washed out by an abundance of water.[194]

Base saturation percentage[[[[edit]

There are acid-forming cations (e.g. hydrogen, aluminium, iron) and there are base-forming cations (e.g. calcium, magnesium, sodium). The fraction of the negatively-charged soil colloid exchange sites (CEC) that are occupied by base-forming cations is called base saturation. If a soil has a CEC of 20 meq and 5 meq are aluminium and hydrogen cations (acid-forming), the remainder of positions on the colloids (20-5 = 15 meq) are assumed occupied by base-forming cations, so that the base saturation is 15/20 x 100% = 75% (the compliment 25% is assumed acid-forming cations or protons). Base saturation is almost in direct proportion to pH (it increases with increasing pH).[195] It is of use in calculating the amount of lime needed to neutralise an acid soil (lime requirement). The amount of lime needed to neutralize a soil must take account of the amount of acid forming ions on the colloids (exchangeable acidity), not just those in the soil water solution (free acidity).[196] The addition of enough lime to neutralize the soil water solution will be insufficient to change the pH, as the acid forming cations stored on the soil colloids will tend to restore the original pH condition as they are pushed off those colloids by the calcium of the added lime.

Buffering[[[[edit]

The resistance of soil to change in pH, as a result of the addition of acid or basic material, is a measure of the buffering capacity of a soil and (for a particular soil type) increases as the CEC increases. Hence, pure sand has almost no buffering ability, while soils high in colloids (whether mineral or organic) have high buffering capacity.[198] Buffering occurs by cation exchange and neutralisation. However, colloids are not the only regulators of soil pH. The role of carbonates should be underlined, too.[199] More generally, according to pH levels, several buffer systems take precedence over each other, from calcium carbonate buffer range to iron buffer range.[200]

The addition of a small amount of highly basic aqueous ammonia to a soil will cause the ammonium to displace hydrogen ions from the colloids, and the end product is water and colloidally fixed ammonium, but little permanent change overall in soil pH.

The addition of a small amount of lime, Ca(OH)2, will displace hydrogen ions from the soil colloids, causing the fixation of calcium to colloids and the evolution of CO2 and water, with little permanent change in soil pH.

The above are examples of the buffering of soil pH. The general principal is that an increase in a particular cation in the soil water solution will cause that cation to be fixed to colloids (buffered) and a decrease in solution of that cation will cause it to be withdrawn from the colloid and moved into solution (buffered). The degree of buffering is often related to the CEC of the soil; the greater the CEC, the greater the buffering capacity of the soil.

Nutrients[[[[edit]

Plant nutrients, their chemical symbols, and the ionic forms common in soils and available for plant uptake
Element Symbol Ion or molecule
Carbon C CO2 (mostly through leaves)
Hydrogen H H+, HOH (water)
Oxygen O O2−, OH, CO32−, SO42−, CO2
Phosphorus P H2PO4, HPO42− (phosphates)
Potassium K K+
Nitrogen N NH4+, NO3 (ammonium, nitrate)
Sulfur S SO42−
Calcium Ca Ca2+
Iron Fe Fe2+, Fe3+ (ferrous, ferric)
Magnesium Mg Mg2+
Boron B H3BO3, H2BO3, B(OH)4
Manganese Mn Mn2+
Copper Cu Cu2+
Zinc Zn Zn2+
Molybdenum Mo MoO42− (molybdate)
Chlorine Cl Cl (chloride)

Seventeen elements or nutrients are essential for plant growth and reproduction. They are carbon (C), hydrogen (H), oxygen (O), nitrogen (N), phosphorus (P), potassium (K), sulfur (S), calcium (Ca), magnesium (Mg), iron (Fe), boron (B), manganese (Mn), copper (Cu), zinc (Zn), molybdenum (Mo), nickel (Ni) and chlorine (Cl).[205] Nutrients required for plants to complete their life cycle are considered essential nutrients. Nutrients that enhance the growth of plants but are not necessary to complete the plant’s life cycle are considered non-essential. With the exception of carbon, hydrogen and oxygen, which are supplied by carbon dioxide and water, and nitrogen, provided through nitrogen fixation,[205] the nutrients derive originally from the mineral component of the soil. The Law of the Minimum expresses that when the available form of a nutrient is not in enough proportion in the soil solution, then other nutrients cannot be taken up at an optimum rate by a plant.[206] A particular nutrient ratio of the soil solution is thus mandatory for optimizing plant growth, a value which might differ from nutrient ratios calculated from plant composition.[207]

Plant uptake of nutrients can only proceed when they are present in a plant-available form. In most situations, nutrients are absorbed in an ionic form from (or together with) soil water. Although minerals are the origin of most nutrients, and the bulk of most nutrient elements in the soil is held in crystalline form within primary and secondary minerals, they weather too slowly to support rapid plant growth. For example, the application of finely ground minerals, feldspar and apatite, to soil seldom provides the necessary amounts of potassium and phosphorus at a rate sufficient for good plant growth, as most of the nutrients remain bound in the crystals of those minerals.

The nutrients adsorbed onto the surfaces of clay colloids and soil organic matter provide a more accessible reservoir of many plant nutrients (e.g. K, Ca, Mg, P, Zn). As plants absorb the nutrients from the soil water, the soluble pool is replenished from the surface-bound pool. The decomposition of soil organic matter by microorganisms is another mechanism whereby the soluble pool of nutrients is replenished – this is important for the supply of plant-available N, S, P, and B from soil.[209]

Gram for gram, the capacity of humus to hold nutrients and water is far greater than that of clay minerals, most of the soil cation exchange capacity arising from charged carboxylic groups on organic matter.[210] However, despite the great capacity of humus to retain water once water-soaked, its high hydrophobicity decreases its wettability.[211] All in all, small amounts of humus may remarkably increase the soil’s capacity to promote plant growth.[209]

Soil organic matter[[[[edit]

Soil organic matter is made up of organic compounds and includes plant, animal and microbial material, both living and dead. A typical soil has a biomass composition of 70% microorganisms, 22% macrofauna, and 8% roots. The living component of an acre of soil may include 900 lb of earthworms, 2400 lb of fungi, 1500 lb of bacteria, 133 lb of protozoa and 890 lb of arthropods and algae.[213]

A few percent of the soil organic matter, with small residence time, consists of the microbial biomass and metabolites of bacteria, molds, and actinomycetes that work to break down the dead organic matter.[214][215] Were it not for the action of these micro-organisms, the entire carbon dioxide part of the atmosphere would be sequestered as organic matter in the soil. However, in the same time soil microbes contribute to carbon sequestration in the topsoil through the formation of stable humus.[216] In the aim to sequester more carbon in the soil for alleviating the greenhouse effect it would be more efficient in the long-term to stimulate humification than to decrease litter decomposition.[217]

The main part of soil organic matter is a complex assemblage of small organic molecules, collectively called humus or humic substances. The use of these terms, which do not rely on a clear chemical classification, has been considered as obsolete.[218] Other studies showed that the classical notion of molecule is not convenient for humus, which escaped most attempts done over two centuries to resolve it in unit components, but still is chemically distinct from polysaccharides, lignins and proteins.[219]

Most living things in soils, including plants, animals, bacteria, and fungi, are dependent on organic matter for nutrients and/or energy. Soils have organic compounds in varying degrees of decomposition which rate is dependent on the temperature, soil moisture, and aeration. Bacteria and fungi feed on the raw organic matter, which are fed upon by protozoa, which in turn are fed upon by nematodes, annelids and arthropods, themselves able to consume and transform raw or humified organic matter. This has been called the soil food web, through which all organic matter is processed as in a digestive system.[220] Organic matter holds soils open, allowing the infiltration of air and water, and may hold as much as twice its weight in water. Many soils, including desert and rocky-gravel soils, have little or no organic matter. Soils that are all organic matter, such as peat (histosols), are infertile.[221] In its earliest stage of decomposition, the original organic material is often called raw organic matter. The final stage of decomposition is called humus.

In grassland, much of the organic matter added to the soil is from the deep, fibrous, grass root systems. By contrast, tree leaves falling on the forest floor are the principal source of soil organic matter in the forest. Another difference is the frequent occurrence in the grasslands of fires that destroy large amounts of aboveground material but stimulate even greater contributions from roots. Also, the much greater acidity under any forests inhibits the action of certain soil organisms that otherwise would mix much of the surface litter into the mineral soil. As a result, the soils under grasslands generally develop a thicker A horizon with a deeper distribution of organic matter than in comparable soils under forests, which characteristically store most of their organic matter in the forest floor (O horizon) and thin A horizon.[222]

Humus[[[[edit]

Humus refers to organic matter that has been decomposed by soil microflora and fauna to the point where it is resistant to further breakdown. Humus usually constitutes only five percent of the soil or less by volume, but it is an essential source of nutrients and adds important textural qualities crucial to soil health and plant growth.[223] Humus also feeds arthropods, termites and earthworms which further improve the soil.[224] The end product, humus, is suspended in colloidal form in the soil solution and forms a weak acid that can attack silicate minerals.[225] Humus has a high cation and anion exchange capacity that on a dry weight basis is many times greater than that of clay colloids. It also acts as a buffer, like clay, against changes in pH and soil moisture.[226]

Humic acids and fulvic acids, which begin as raw organic matter, are important constituents of humus. After the death of plants, animals, and microbes, microbes begin to feed on the residues through their production of extra-cellular soil enzymes, resulting finally in the formation of humus.[227] As the residues break down, only molecules made of aliphatic and aromatic hydrocarbons, assembled and stabilized by oxygen and hydrogen bonds, remain in the form of complex molecular assemblages collectively called humus.[219] Humus is never pure in the soil, because it reacts with metals and clays to form complexes which further contribute to its stability and to soil structure.[226] While the structure of humus has in itself few nutrients (with the exception of constitutive metals such as calcium, iron and aluminum) it is able to attract and link, by weak bonds, cation and anion nutrients that can further be released into the soil solution in response to selective root uptake and changes in soil pH, a process of paramount importance for the maintenance of fertility in tropical soils.[228]

Lignin is resistant to breakdown and accumulates within the soil. It also reacts with proteins,[229] which further increases its resistance to decomposition, including enzymatic decomposition by microbes.[230]Fats and waxes from plant matter have still more resistance to decomposition and persist in soils for thousand years, hence their use as tracers of past vegetation in buried soil layers.[231] Clay soils often have higher organic contents that persist longer than soils without clay as the organic molecules adhere to and are stabilised by the clay.[232]Proteins normally decompose readily, to the exception of scleroproteins, but when bound to clay particles they become more resistant to decomposition.[233] As for other proteins clay particles absorb the enzymes exuded by microbes, decreasing enzyme activity while protecting extracellular enzymes from degradation.[234] The addition of organic matter to clay soils can render that organic matter and any added nutrients inaccessible to plants and microbes for many years,[[[[citation needed] while a study showed increased soil fertility following the addition of mature compost to a clay soil.[235] High soil tannin content can cause nitrogen to be sequestered as resistant tannin-protein complexes.[236][237]

Humus formation is a process dependent on the amount of plant material added each year and the type of base soil. Both are affected by climate and the type of organisms present.[238] Soils with humus can vary in nitrogen content but typically have 3 to 6 percent nitrogen. Raw organic matter, as a reserve of nitrogen and phosphorus, is a vital component affecting soil fertility.[221] Humus also absorbs water, and expands and shrinks between dry and wet states to a higher extent than clay, increasing soil porosity.[239] Humus is less stable than the soil’s mineral constituents, as it is reduced by microbial decomposition, and over time its concentration diminishes without the addition of new organic matter. However, humus in its most stable forms may persist over centuries if not millennia.[240]Charcoal is a source of highly stable humus, called black carbon,[241] which had been used traditionally to improve the fertility of nutrient-poor tropical soils. This very ancient practice, as ascertained in the genesis of Amazonian dark earths, has been renewed and became popular under the name of biochar. It has been suggested that biochar could be used to sequester more carbon in the fight against the greenhouse effect.[242]

Climatological influence[[[[edit]

The production, accumulation and degradation of organic matter are greatly dependent on climate. Temperature, soil moisture and topography are the major factors affecting the accumulation of organic matter in soils. Organic matter tends to accumulate under wet or cold conditions where decomposer activity is impeded by low temperature[243] or excess moisture which results in anaerobic conditions.[244] Conversely, excessive rain and high temperatures of tropical climates enables rapid decomposition of organic matter and leaching of plant nutrients. Forest ecosystems on these soils rely on efficient recycling of nutrients and plant matter by the living plant and microbial biomass to maintain their productivity, a process which is disturbed by human activities.[245] Excessive slope, in particular in the presence of cultivation for the sake of agriculture, may encourage the erosion of the top layer of soil which holds most of the raw organic material that would otherwise eventually become humus.[246]

Plant residue[[[[edit]

Typical types and percentages of plant residue components

Cellulose (45%)

Lignin (20%)

Hemicellulose (18%)

Protein (8%)

Sugars and starches (5%)

Fats and waxes (2%)

Cellulose and hemicellulose undergo fast decomposition by fungi and bacteria, with a half-life of 12–18 days in a temperate climate.[247]Brown rot fungi can decompose the cellulose and hemicellulose, leaving the lignin and phenolic compounds behind. Starch, which is an energy storage system for plants, undergoes fast decomposition by bacteria and fungi. Lignin consists of polymers composed of 500 to 600 units with a highly branched, amorphous structure, linked to cellulose, hemicellulose and pectin in plant cell walls. Lignin undergoes very slow decomposition, mainly by white rot fungi and actinomycetes; its half-life under temperate conditions is about six months.[247]

Horizons[[[[edit]

A horizontal layer of the soil, whose physical features, composition and age are distinct from those above and beneath, is referred to as a soil horizon. The naming of a horizon is based on the type of material of which it is composed. Those materials reflect the duration of specific processes of soil formation. They are labelled using a shorthand notation of letters and numbers which describe the horizon in terms of its colour, size, texture, structure, consistency, root quantity, pH, voids, boundary characteristics and presence of nodules or concretions.[248] No soil profile has all the major horizons. Some, called entisols, may have only one horizon or are currently considered as having no horizon, in particular incipient soils from unreclaimed mining waste deposits,[249]moraines,[250]volcanic cones[251]sand dunes or alluvial terraces.[252] Upper soil horizons may be lacking in truncated soils following wind or water ablation, with concomitant downslope burying of soil horizons, a natural process aggravated by agricultural practices such as tillage.[253] The growth of trees is another source of disturbance, creating a micro-scale heterogeneity which is still visible in soil horizons once trees have died.[254] By passing from a horizon to another, from the top to the bottom of the soil profile, one goes back in time, with past events registered in soil horizons like in sediment layers. Sampling pollen, testate amoebae and plant remains in soil horizons may help to reveal environmental changes (e.g. climate change, land use change) which occurred in the course of soil formation.[255] Soil horizons can be dated by several methods such as radiocarbon, using pieces of charcoal provided they are of enough size to escape pedoturbation by earthworm activity and other mechanical disturbances.[256] Fossil soil horizons from paleosols can be found within sedimentary rock sequences, allowing the study of past environments.[257]

The exposure of parent material to favourable conditions produces mineral soils that are marginally suitable for plant growth, as is the case in eroded soils.[258] The growth of vegetation results in the production of organic residues which fall on the ground as litter for plant aerial parts (leaf litter) or are directly produced belowground for subterranean plant organs (root litter), and then release dissolved organic matter.[259] The remaining surficial organic layer, called the O horizon, produces a more active soil due to the effect of the organisms that live within it. Organisms colonise and break down organic materials, making available nutrients upon which other plants and animals can live.[260] After sufficient time, humus moves downward and is deposited in a distinctive organic-mineral surface layer called the A horizon, in which organic matter is mixed with mineral matter through the activity of burrowing animals, a process called pedoturbation. This natural process does not go to completion in the presence of conditions detrimental to soil life such as strong acidity, cold climate or pollution, stemming in the accumulation of undecomposed organic matter within a single organic horizon overlying the mineral soil[261] and in the juxtaposition of humified organic matter and mineral particles, without intimate mixing, in the underlying mineral horizons.[262]

Classification[[[[edit]

Soil is classified into categories in order to understand relationships between different soils and to determine the suitability of a soil in a particular region. One of the first classification systems was developed by the Russian scientist Vasily Dokuchaev around 1880.[263] It was modified a number of times by American and European researchers, and developed into the system commonly used until the 1960s. It was based on the idea that soils have a particular morphology based on the materials and factors that form them. In the 1960s, a different classification system began to emerge which focused on soil morphology instead of parental materials and soil-forming factors. Since then it has undergone further modifications. The World Reference Base for Soil Resources (WRB)[264] aims to establish an international reference base for soil classification.

Soil is used in agriculture, where it serves as the anchor and primary nutrient base for plants. The types of soil and available moisture determine the species of plants that can be cultivated. Agricultural soil science was the primeval domain of soil knowledge, long time before the advent of pedology in the 19th century. However, as demonstrated by aeroponics, aquaponics and hydroponics, soil material is not an absolute essential for agriculture, and soilless cropping systems have been claimed as the future of agriculture for an endless growing mankind.[265]

Soil material is also a critical component in the mining, construction and landscape development industries.[266] Soil serves as a foundation for most construction projects. The movement of massive volumes of soil can be involved in surface mining, road building and dam construction. Earth sheltering is the architectural practice of using soil for external thermal mass against building walls. Many building materials are soil based. Loss of soil through urbanization is growing at a high rate in many areas and can be critical for the maintenance of subsistence agriculture.[267]

Soil resources are critical to the environment, as well as to food and fibre production, producing 98.8% of food consumed by humans.[268] Soil provides minerals and water to plants according to several processes involved in plant nutrition. Soil absorbs rainwater and releases it later, thus preventing floods and drought, flood regulation being one of the major ecosystem services provided by soil.[269] Soil cleans water as it percolates through it.[270] Soil is the habitat for many organisms: the major part of known and unknown biodiversity is in the soil, in the form of invertebrates (earthworms, woodlice, millipedes, centipedes, snails, slugs, mites, springtails, enchytraeids, nematodes, protists), bacteria, archaea, fungi and algae; and most organisms living above ground have part of them (plants) or spend part of their life cycle (insects) below-ground.[271] Above-ground and below-ground biodiversities are tightly interconnected,[238][272] making soil protection of paramount importance for any restoration or conservation plan.

The biological component of soil is an extremely important carbon sink since about 57% of the biotic content is carbon. Even in deserts, cyanobacteria, lichens and mosses form biological soil crusts which capture and sequester a significant amount of carbon by photosynthesis. Poor farming and grazing methods have degraded soils and released much of this sequestered carbon to the atmosphere. Restoring the world’s soils could offset the effect of increases in greenhouse gas emissions and slow global warming, while improving crop yields and reducing water needs.[273][274][275]

Waste management often has a soil component. Septic drain fields treat septic tank effluent using aerobic soil processes. Land application of waste water relies on soil biology to aerobically treat BOD. Alternatively, Landfills use soil for daily cover, isolating waste deposits from the atmosphere and preventing unpleasant smells. Composting is now widely used to treat aerobically solid domestic waste and dried effluents of settling basins. Although compost is not soil, biological processes taking place during composting are similar to those occurring during decomposition and humification of soil organic matter.[276]

Organic soils, especially peat, serve as a significant fuel and horticultural resource. Peat soils are also commonly used for the sake of agriculture in nordic countries, because peatland sites, when drained, provide fertile soils for food production.[277] However, wide areas of peat production, such as rain-fed sphagnum bogs, also called blanket bogs or raised bogs, are now protected because of their patrimonial interest. As an example, Flow Country, covering 4,000 square kilometres of rolling expanse of blanket bogs in Scotland, is now candidate for being included in the World Heritage List. Under present-day global warming peat soils are thought to be involved in a self-reinforcing (positive feedback) process of increased emission of greenhouse gases (methane and carbon dioxide) and increased temperature,[278] a contention which is still under debate when replaced at field scale and including stimulated plant growth.[279]

Geophagy is the practice of eating soil-like substances. Both animals and humans occasionally consume soil for medicinal, recreational, or religious purposes.[280] It has been shown that some monkeys consume soil, together with their preferred food (tree foliage and fruits), in order to alleviate tannin toxicity.[281]

Soils filter and purify water and affect its chemistry. Rain water and pooled water from ponds, lakes and rivers percolate through the soil horizons and the upper rock strata, thus becoming groundwater. Pests (viruses) and pollutants, such as persistent organic pollutants (chlorinated pesticides, polychlorinated biphenyls), oils (hydrocarbons), heavy metals (lead, zinc, cadmium), and excess nutrients (nitrates, sulfates, phosphates) are filtered out by the soil.[282] Soil organisms metabolise them or immobilise them in their biomass and necromass,[283] thereby incorporating them into stable humus.[284] The physical integrity of soil is also a prerequisite for avoiding landslides in rugged landscapes.[285]

Degradation[[[[edit]

Land degradation refers to a human-induced or natural process which impairs the capacity of land to function.[286] Soil degradation involves acidification, contamination, desertification, erosion or salination.[287]

Soil acidification is beneficial in the case of alkaline soils, but it degrades land when it lowers crop productivity, soil biological activity and increases soil vulnerability to contamination and erosion. Soils are initially acid and remain such when their parent materials are low in basic cations (calcium, magnesium, potassium and sodium). On parent materials richer in weatherable minerals acidification occurs when basic cations are leached from the soil profile by rainfall or exported by the harvesting of forest or agricultural crops. Soil acidification is accelerated by the use of acid-forming nitrogenous fertilizers and by the effects of acid precipitation. Deforestation is another cause of soil acidification, mediated by increased leaching of soil nutrients in the absence of tree canopies.[288]

Soil contamination at low levels is often within a soil’s capacity to treat and assimilate waste material. Soil biota can treat waste by transforming it, mainly through microbial enzymatic activity.[289]Soil organic matter and soil minerals can adsorb the waste material and decrease its toxicity,[290] although when in colloidal form they may transport the adsorbed contaminants to subsurface environments.[291] Many waste treatment processes rely on this natural bioremediation capacity. Exceeding treatment capacity can damage soil biota and limit soil function. Derelict soils occur where industrial contamination or other development activity damages the soil to such a degree that the land cannot be used safely or productively. Remediation of derelict soil uses principles of geology, physics, chemistry and biology to degrade, attenuate, isolate or remove soil contaminants to restore soil functions and values. Techniques include leaching, air sparging, soil conditioners, phytoremediation, bioremediation and Monitored Natural Attenuation (MNA). An example of diffuse pollution with contaminants is copper accumulation in vineyards and orchards to which fungicides are repeatedly applied, even in organic farming.[292]

Desertification is an environmental process of ecosystem degradation in arid and semi-arid regions, often caused by badly adapted human activities such as overgrazing or excess harvesting of firewood. It is a common misconception that drought causes desertification.[293] Droughts are common in arid and semiarid lands. Well-managed lands can recover from drought when the rains return. Soil management tools include maintaining soil nutrient and organic matter levels, reduced tillage and increased cover.[294] These practices help to control erosion and maintain productivity during periods when moisture is available. Continued land abuse during droughts, however, increases land degradation. Increased population and livestock pressure on marginal lands accelerates desertification.[295] It is now questioned whether present-day climate warming will favour or disfavour desertification, with contradictory reports about predicted rainfall trends associated with increased temperature, and strong discrepancies among regions, even in the same country.[296]

Erosion of soil is caused by water, wind, ice, and movement in response to gravity. More than one kind of erosion can occur simultaneously. Erosion is distinguished from weathering, since erosion also transports eroded soil away from its place of origin (soil in transit may be described as sediment). Erosion is an intrinsic natural process, but in many places it is greatly increased by human activity, especially unsuitable land use practices.[297] These include agricultural activities which leave the soil bare during times of heavy rain or strong winds, overgrazing, deforestation, and improper construction activity. Improved management can limit erosion. Soil conservation techniques which are employed include changes of land use (such as replacing erosion-prone crops with grass or other soil-binding plants), changes to the timing or type of agricultural operations, terrace building, use of erosion-suppressing cover materials (including cover crops and other plants), limiting disturbance during construction, and avoiding construction during erosion-prone periods and in erosion-prone places such as steep slopes.[298] Historically, one of the best examples of large-scale soil erosion due to unsuitable land-use practices is wind erosion (the so-called dust bowl) which ruined American and Canadian prairies during the 1930s, when immigrant farmers, encouraged by the federal government of both countries, settled and converted the original shortgrass prairie to agricultural crops and cattle ranching.

A serious and long-running water erosion problem occurs in China, on the middle reaches of the Yellow River and the upper reaches of the Yangtze River. From the Yellow River, over 1.6 billion tons of sediment flow each year into the ocean. The sediment originates primarily from water erosion (gully erosion) in the Loess Plateau region of northwest China.[299]

Soil piping is a particular form of soil erosion that occurs below the soil surface.[300] It causes levee and dam failure, as well as sink hole formation. Turbulent flow removes soil starting at the mouth of the seep flow and the subsoil erosion advances up-gradient.[301] The term sand boil is used to describe the appearance of the discharging end of an active soil pipe.[302]

Soil salination is the accumulation of free salts to such an extent that it leads to degradation of the agricultural value of soils and vegetation. Consequences include corrosion damage, reduced plant growth, erosion due to loss of plant cover and soil structure, and water quality problems due to sedimentation. Salination occurs due to a combination of natural and human-caused processes. Arid conditions favour salt accumulation. This is especially apparent when soil parent material is saline. Irrigation of arid lands is especially problematic.[303] All irrigation water has some level of salinity. Irrigation, especially when it involves leakage from canals and overirrigation in the field, often raises the underlying water table. Rapid salination occurs when the land surface is within the capillary fringe of saline groundwater. Soil salinity control involves watertable control and flushing with higher levels of applied water in combination with tile drainage or another form of subsurface drainage.[304][305]

Reclamation[[[[edit]

Soils which contain high levels of particular clays with high swelling properties, such as smectites, are often very fertile. For example, the smectite-rich paddy soils of Thailand’s Central Plains are among the most productive in the world. However, the overuse of mineral nitrogen fertilizers and pesticides in irrigated intensive rice production has endangered these soils, forcing farmers to implement integrated practices based on Cost Reduction Operating Principles (CROP).[306]

Many farmers in tropical areas, however, struggle to retain organic matter and clay in the soils they work. In recent years, for example, productivity has declined and soil erosion has increased in the low-clay soils of northern Thailand, following the abandonment of shifting cultivation for a more permanent land use.[307] Farmers initially responded by adding organic matter and clay from termite mound material, but this was unsustainable in the long-term because of rarefaction of termite mounds. Scientists experimented with adding bentonite, one of the smectite family of clays, to the soil. In field trials, conducted by scientists from the International Water Management Institute in cooperation with Khon Kaen University and local farmers, this had the effect of helping retain water and nutrients. Supplementing the farmer’s usual practice with a single application of 200 kg bentonite per rai (6.26 rai = 1 hectare) resulted in an average yield increase of 73%.[308] Other studies showed that applying bentonite to degraded sandy soils reduced the risk of crop failure during drought years.[309]

In 2008, three years after the initial trials, IWMI scientists conducted a survey among 250 farmers in northeast Thailand, half of whom had applied bentonite to their fields. The average improvement for those using the clay addition was 18% higher than for non-clay users. Using the clay had enabled some farmers to switch to growing vegetables, which need more fertile soil. This helped to increase their income. The researchers estimated that 200 farmers in northeast Thailand and 400 in Cambodia had adopted the use of clays, and that a further 20,000 farmers were introduced to the new technique.[310]

If the soil is too high in clay or salts (e.g. saline sodic soil), adding gypsum, washed river sand and organic matter (e.g.municipal solid waste) will balance the composition.[311]

Adding organic matter, like ramial chipped wood or compost, to soil which is depleted in nutrients and too high in sand will boost its quality and improve production.[312][313]

Special mention must be made of the use of charcoal, and more generally biochar to improve nutrient-poor tropical soils, a process based on the higher fertility of anthropogenic pre-Colombian Amazonian Dark Earths, also called Terra Preta de Índio, due to interesting physical and chemical properties of soil black carbon as a source of stable humus.[314]

History of studies and research[[[[edit]

The history of the study of soil is intimately tied to humans’ urgent need to provide food for themselves and forage for their animals. Throughout history, civilizations have prospered or declined as a function of the availability and productivity of their soils.[315]

Studies of soil fertility[[[[edit]

The Greek historian Xenophon (450–355 BCE) is credited with being the first to expound upon the merits of green-manuring crops: « But then whatever weeds are upon the ground, being turned into earth, enrich the soil as much as dung. »

Columella’s Of husbandry, circa 60 CE, advocated the use of lime and that clover and alfalfa (green manure) should be turned under,[317] and was used by 15 generations (450 years) under the Roman Empire until its collapse. From the fall of Rome to the French Revolution, knowledge of soil and agriculture was passed on from parent to child and as a result, crop yields were low. During the European Middle Ages, Yahya Ibn al-‘Awwam’s handbook,[319] with its emphasis on irrigation, guided the people of North Africa, Spain and the Middle East; a translation of this work was finally carried to the southwest of the United States when under Spanish influence.[320]Olivier de Serres, considered as the father of French agronomy, was the first to suggest the abandonment of fallowing and its replacement by hay meadows within crop rotations, and he highlighted the importance of soil (the French terroir) in the management of vineyards. His famous book Le Théâtre d’Agriculture et mesnage des champs[321] contributed to the rise of modern, sustainable agriculture and to the collapse of old agricultural practices such as soil amendment for crops by the lifting of forest litter and assarting, which ruined the soils of western Europe during Middle Ages and even later on according to regions.[322]

Experiments into what made plants grow first led to the idea that the ash left behind when plant matter was burned was the essential element but overlooked the role of nitrogen, which is not left on the ground after combustion, a belief which prevailed until the 19th century.[323] In about 1635, the Flemish chemist Jan Baptist van Helmont thought he had proved water to be the essential element from his famous five years’ experiment with a willow tree grown with only the addition of rainwater. His conclusion came from the fact that the increase in the plant’s weight had apparently been produced only by the addition of water, with no reduction in the soil’s weight.[324][325]John Woodward (d. 1728) experimented with various types of water ranging from clean to muddy and found muddy water the best, and so he concluded that earthy matter was the essential element. Others concluded it was humus in the soil that passed some essence to the growing plant. Still others held that the vital growth principal was something passed from dead plants or animals to the new plants. At the start of the 18th century, Jethro Tull demonstrated that it was beneficial to cultivate (stir) the soil, but his opinion that the stirring made the fine parts of soil available for plant absorption was erroneous.[325]

As chemistry developed, it was applied to the investigation of soil fertility. The French chemist Antoine Lavoisier showed in about 1778 that plants and animals must [combust] oxygen internally to live and was able to deduce that most of the 165-pound weight of van Helmont’s willow tree derived from air.[328] It was the French agriculturalist Jean-Baptiste Boussingault who by means of experimentation obtained evidence showing that the main sources of carbon, hydrogen and oxygen for plants were air and water, while nitrogen was taken from soil.[329]Justus von Liebig in his book Organic chemistry in its applications to agriculture and physiology (published 1840), asserted that the chemicals in plants must have come from the soil and air and that to maintain soil fertility, the used minerals must be replaced.[330] Liebig nevertheless believed the nitrogen was supplied from the air. The enrichment of soil with guano by the Incas was rediscovered in 1802, by Alexander von Humboldt. This led to its mining and that of Chilean nitrate and to its application to soil in the United States and Europe after 1840.[331]

The work of Liebig was a revolution for agriculture, and so other investigators started experimentation based on it. In England John Bennet Lawes and Joseph Henry Gilbert worked in the Rothamsted Experimental Station, founded by the former, and (re)discovered that plants took nitrogen from the soil, and that salts needed to be in an available state to be absorbed by plants. Their investigations also produced the superphosphate, consisting in the acid treatment of phosphate rock. This led to the invention and use of salts of potassium (K) and nitrogen (N) as fertilizers. Ammonia generated by the production of coke was recovered and used as fertiliser.[333] Finally, the chemical basis of nutrients delivered to the soil in manure was understood and in the mid-19th century chemical fertilisers were applied. However, the dynamic interaction of soil and its life forms still awaited discovery.

In 1856 J. Thomas Way discovered that ammonia contained in fertilisers was transformed into nitrates,[334] and twenty years later Robert Warington proved that this transformation was done by living organisms.[335] In 1890 Sergei Winogradsky announced he had found the bacteria responsible for this transformation.[336]

It was known that certain legumes could take up nitrogen from the air and fix it to the soil but it took the development of bacteriology towards the end of the 19th century to lead to an understanding of the role played in nitrogen fixation by bacteria. The symbiosis of bacteria and leguminous roots, and the fixation of nitrogen by the bacteria, were simultaneously discovered by the German agronomist Hermann Hellriegel and the Dutch microbiologist Martinus Beijerinck.

Crop rotation, mechanisation, chemical and natural fertilisers led to a doubling of wheat yields in western Europe between 1800 and 1900.

Studies of soil formation[[[[edit]

The scientists who studied the soil in connection with agricultural practices had considered it mainly as a static substrate. However, soil is the result of evolution from more ancient geological materials, under the action of biotic and abiotic processes. After studies of the improvement of the soil commenced, other researchers began to study soil genesis and as a result also soil types and classifications.

In 1860, in Mississippi, Eugene W. Hilgard (1833-1916) studied the relationship between rock material, climate, vegetation, and the type of soils that were developed. He realised that the soils were dynamic, and considered the classification of soil types.[338] Unfortunately his work was not continued. At about the same time, Friedrich Albert Fallou was describing soil profiles and relating soil characteristics to their formation as part of his professional work evaluating forest and farm land for the principality of Saxony. His 1857 book, Anfangsgründe der Bodenkunde (First principles of soil science) established modern soil science.[339] Contemporary with Fallou’s work, and driven by the same need to accurately assess land for equitable taxation, Vasily Dokuchaev led a team of soil scientists in Russia who conducted an extensive survey of soils, observing that similar basic rocks, climate and vegetation types lead to similar soil layering and types, and established the concepts for soil classifications. Due to language barriers, the work of this team was not communicated to western Europe until 1914 through a publication in German by Konstantin Glinka, a member of the Russian team.[340]

Curtis F. Marbut, influenced by the work of the Russian team, translated Glinka’s publication into English,[341] and as he was placed in charge of the U.S. National Cooperative Soil Survey, applied it to a national soil classification system.[325]

See also[[[[edit]

References[[[[edit]

  1. ^ Chesworth, Ward, ed. (2008). Encyclopedia of soil science (PDF). Dordrecht, The Netherlands: Springer. ISBN 978-1-4020-3994-2. Archived from the original (PDF) on 5 September 2018.
  2. ^ Voroney, R. Paul; Heck, Richard J. (2007). « The soil habitat » (PDF). In Paul, Eldor A. (ed.). Soil microbiology, ecology and biochemistry (3rd ed.). Amsterdam, the Netherlands: Elsevier. pp. 25–49. doi:10.1016/B978-0-08-047514-1.50006-8. ISBN 978-0-12-546807-7. Archived from the original (PDF) on 10 July 2018.
  3. ^ Taylor, Sterling A.; Ashcroft, Gaylen L. (1972). Physical edaphology: the physics of irrigated and nonirrigated soils. San Francisco, California: W.H. Freeman. ISBN 978-0-7167-0818-6.
  4. ^ McCarthy, David F. (2006). Essentials of soil mechanics and foundations: basic geotechnics (PDF) (7th ed.). Upper Saddle River, New Jersey: Prentice Hall. ISBN 978-0-13-114560-3. Retrieved 17 January 2021.
  5. ^ Gilluly, James; Waters, Aaron Clement; Woodford, Alfred Oswald (1975). Principles of geology (4th ed.). San Francisco, California: W.H. Freeman. ISBN 978-0-7167-0269-6.
  6. ^ Ponge, Jean-François (2015). « The soil as an ecosystem ». Biology and Fertility of Soils. 51 (6): 645–48. doi:10.1007/s00374-015-1016-1. S2CID 18251180. Retrieved 24 January 2021.
  7. ^ Yu, Charley; Kamboj, Sunita; Wang, Cheng; Cheng, Jing-Jy (2015). « Data collection handbook to support modeling impacts of radioactive material in soil and building structures » (PDF). Argonne National Laboratory. pp. 13–21. Archived (PDF) from the original on 4 August 2018. Retrieved 24 January 2021.
  8. ^ une b Buol, Stanley W.; Southard, Randal J.; Graham, Robert C.; McDaniel, Paul A. (2011). Soil genesis and classification (7th ed.). Ames, Iowa: Wiley-Blackwell. ISBN 978-0-470-96060-8.
  9. ^ Retallack, Gregory J.; Krinsley, David H.; Fischer, Robert; Razink, Joshua J.; Langworthy, Kurt A. (2016). « Archean coastal-plain paleosols and life on land » (PDF). Gondwana Research. 40: 1–20. Bibcode:2016GondR..40….1R. doi:10.1016/j.gr.2016.08.003. Archived (PDF) from the original on 13 November 2018. Retrieved 24 January 2021.
  10. ^ « Glossary of terms in soil science ». Agriculture and Agri-Food Canada. Archived from the original on 27 October 2018. Retrieved 24 January 2021.
  11. ^ Amundson, Ronald. « Soil preservation and the future of pedology » (PDF). Faculty of Natural Resources. Songkhla, Thailand: Prince of Songkla University. Archived (PDF) from the original on 12 June 2018. Retrieved 24 January 2021.
  12. ^ Küppers, Michael; Vincent, Jean-Baptiste. « Impacts and formation of regolith ». Max Planck Institute for Solar System Research. Archived from the original on 4 August 2018. Retrieved 24 January 2021.
  13. ^ Amelung, Wulf; Bossio, Deborah; De Vries, Wim; Kögel-Knabner, Ingrid; Lehmann, Johannes; Amundson, Ronald; Bol, Roland; Collins, Chris; Lal, Rattan; Leifeld, Jens; Minasny, Buniman; Pan, Gen-Xing; Paustian, Keith; Rumpel, Cornelia; Sanderman, Jonathan; Van Groeningen, Jan Willem; Mooney, Siân; Van Wesemael, Bas; Wander, Michelle; Chabbi, Abad (27 October 2020). « Towards a global-scale soil climate mitigation strategy » (PDF). Nature Communications. 11 (1): 5427. Bibcode:2020NatCo..11.5427A. doi:10.1038/s41467-020-18887-7. ISSN 2041-1723. PMC 7591914. PMID 33110065. Retrieved 31 January 2021.
  14. ^ Pouyat, Richard; Groffman, Peter; Yesilonis, Ian; Hernandez, Luis (2002). « Soil carbon pools and fluxes in urban ecosystems ». Environmental Pollution. 116 (Supplement 1): S107–S118. doi:10.1016/S0269-7491(01)00263-9. PMID 11833898. Retrieved 7 February 2021. Our analysis of pedon data from several disturbed soil profiles suggests that physical disturbances and anthropogenic inputs of various materials (direct effects) can greatly alter the amount of C stored in these human « made » soils.
  15. ^ Davidson, Eric A.; Janssens, Ivan A. (2006). « Temperature sensitivity of soil carbon decomposition and feedbacks to climate change » (PDF). Nature. 440 (9 March 2006): 165‒73. Bibcode:2006Natur.440..165D. doi:10.1038/nature04514. PMID 16525463. S2CID 4404915. Retrieved 7 February 2021.
  16. ^ Powlson, David (2005). « Will soil amplify climate change? » (PDF). Nature. 433 (20 January 2005): 204‒05. Bibcode:2005Natur.433..204P. doi:10.1038/433204a. PMID 15662396. S2CID 35007042. Retrieved 7 February 2021.
  17. ^ Bradford, Mark A.; Wieder, William R.; Bonan, Gordon B.; Fierer, Noah; Raymond, Peter A.; Crowther, Thomas W. (2016). « Managing uncertainty in soil carbon feedbacks to climate change » (PDF). Nature Climate Change. 6 (27 July 2016): 751–58. Bibcode:2016NatCC…6..751B. doi:10.1038/nclimate3071. Retrieved 7 February 2021.
  18. ^ Dominati, Estelle; Patterson, Murray; Mackay, Alec (2010). « A framework for classifying and quantifying the natural capital and ecosystem services of soils ». Ecological Economics. 69 (9): 1858‒68. doi:10.1016/j.ecolecon.2010.05.002. Archived (PDF) from the original on 8 August 2017. Retrieved 14 February 2021.
  19. ^ Dykhuizen, Daniel E. (1998). « Santa Rosalia revisited: why are there so many species of bacteria? ». Antonie van Leeuwenhoek. 73 (1): 25‒33. doi:10.1023/A:1000665216662. PMID 9602276. S2CID 17779069. Retrieved 14 February 2021.
  20. ^ Torsvik, Vigdis; Øvreås, Lise (2002). « Microbial diversity and function in soil: from genes to ecosystems ». Current Opinion in Microbiology. 5 (3): 240‒45. doi:10.1016/S1369-5274(02)00324-7. PMID 12057676. Retrieved 14 February 2021.
  21. ^ Raynaud, Xavier; Nunan, Naoise (2014). « Spatial ecology of bacteria at the microscale in soil ». PLOS ONE. 9 (1): e87217. Bibcode:2014PLoSO…987217R. doi:10.1371/journal.pone.0087217. PMC 3905020. PMID 24489873.
  22. ^ Whitman, William B.; Coleman, David C.; Wiebe, William J. (1998). « Prokaryotes: the unseen majority ». Proceedings of the National Academy of Sciences of the USA. 95 (12): 6578‒83. Bibcode:1998PNAS…95.6578W. doi:10.1073/pnas.95.12.6578. PMC 33863. PMID 9618454.
  23. ^ Schlesinger, William H.; Andrews, Jeffrey A. (2000). « Soil respiration and the global carbon cycle ». Biogeochemistry. 48 (1): 7‒20. doi:10.1023/A:1006247623877. S2CID 94252768. Retrieved 14 February 2021.
  24. ^ Denmead, Owen Thomas; Shaw, Robert Harold (1962). « Availability of soil water to plants as affected by soil moisture content and meteorological conditions ». Agronomy Journal. 54 (5): 385‒90. doi:10.2134/agronj1962.00021962005400050005x. Retrieved 14 February 2021.
  25. ^ House, Christopher H.; Bergmann, Ben A.; Stomp, Anne-Marie; Frederick, Douglas J. (1999). « Combining constructed wetlands and aquatic and soil filters for reclamation and reuse of water ». Ecological Engineering. 12 (1–2): 27–38. doi:10.1016/S0925-8574(98)00052-4. Retrieved 14 February 2021.
  26. ^ Van Bruggen, Ariena H.C.; Semenov, Alexander M. (2000). « In search of biological indicators for soil health and disease suppression ». Applied Soil Ecology. 15 (1): 13–24. doi:10.1016/S0929-1393(00)00068-8. Retrieved 14 February 2021.
  27. ^ « A citizen’s guide to monitored natural attenuation » (PDF). Retrieved 14 February 2021.
  28. ^ Linn, Daniel Myron; Doran, John W. (1984). « Effect of water-filled pore space on carbon dioxide and nitrous oxide production in tilled and nontilled soils ». Soil Science Society of America Journal. 48 (6): 1267–72. Bibcode:1984SSASJ..48.1267L. doi:10.2136/sssaj1984.03615995004800060013x. Retrieved 14 February 2021.
  29. ^ Miller, Raymond W.; Donahue, Roy Luther (1990). Soils: an introduction to soils and plant growth. Upper Saddle River, New Jersey: Prentice Hall. ISBN 978-0-13-820226-2.
  30. ^ Bot, Alexandra; Benites, José (2005). The importance of soil organic matter: key to drought-resistant soil and sustained food and production (PDF). Rome: Food and Agriculture Organization of the United Nations. ISBN 978-92-5-105366-9. Retrieved 14 February 2021.
  31. ^ McClellan, Tai. « Soil composition ». University of Hawai‘i at Mānoa, College of Tropical Agriculture and Human Resources. Retrieved 21 February 2021.
  32. ^ « Arizona Master Gardener Manual ». Cooperative Extension, College of Agriculture, University of Arizona. 9 November 2017. Archived from the original on 29 May 2016. Retrieved 17 December 2017.
  33. ^ une b Vannier, Guy (1987). « The porosphere as an ecological medium emphasized in Professor Ghilarov’s work on soil animal adaptations » (PDF). Biology and Fertility of Soils. 3 (1): 39–44. doi:10.1007/BF00260577. S2CID 297400. Retrieved 21 February 2021.
  34. ^ Torbert, H. Allen; Wood, Wes (1992). « Effect of soil compaction and water-filled pore space on soil microbial activity and N losses ». Communications in Soil Science and Plant Analysis. 23 (11): 1321‒31. doi:10.1080/00103629209368668. Retrieved 21 February 2021.
  35. ^ Zanella, Augusto; Katzensteiner, Klaus; Ponge, Jean-François; Jabiol, Bernard; Sartori, Giacomo; Kolb, Eckart; Le Bayon, Renée-Claire; Aubert, Michaël; Ascher-Jenull, Judith; Englisch, Michael; Hager, Herbert (June 2019). « TerrHum: an iOS App for classifying terrestrial humipedons and some considerations about soil classification ». Soil Science Society of America Journal. 83 (S1): S42–S48. doi:10.2136/sssaj2018.07.0279. Retrieved 28 February 2021.
  36. ^ Bronick, Carol J.; Lal, Ratan (January 2005). « Soil structure and management: a review » (PDF). Geoderma. 124 (1/2): 3–22. Bibcode:2005Geode.124….3B. doi:10.1016/j.geoderma.2004.03.005. Retrieved 21 February 2021.
  37. ^ « Soil and water ». Food and Agriculture Organization of the United Nations. Retrieved 21 February 2021.
  38. ^ Valentin, Christian; d’Herbès, Jean-Marc; Poesen, Jean (1999). « Soil and water components of banded vegetation patterns ». Catena. 37 (1): 1‒24. doi:10.1016/S0341-8162(99)00053-3. Retrieved 21 February 2021.
  39. ^ Brady, Nyle C.; Weil, Ray R. (2007). « The colloidal fraction: seat of soil chemical and physical activity ». In Brady, Nyle C.; Weil, Ray R. (eds.). The nature and properties of soils (14th ed.). London, United Kingdom: Pearson. pp. 310–57. ISBN 978-0132279383. Retrieved 21 February 2021.
  40. ^ « Soil colloids: properties, nature, types and significance » (PDF). Tamil Nadu Agricultural University. Retrieved 7 March 2021.
  41. ^ une b « Cation exchange capacity in soils, simplified ». Retrieved 7 March 2021.
  42. ^ Miller, Jarrod O. « Soil pH affects nutrient availability » (PDF). University of Maryland. Retrieved 7 March 2021.
  43. ^ Goulding, Keith W.T.; Bailey, Neal J.; Bradbury, Nicola J.; Hargreaves, Patrick; Howe, M.T.; Murphy, Daniel V.; Poulton, Paul R.; Willison, Toby W. (1998). « Nitrogen deposition and its contribution to nitrogen cycling and associated soil processes ». New Phytologist. 139 (1): 49‒58. doi:10.1046/j.1469-8137.1998.00182.x.
  44. ^ Kononova, M.M. (2013). Soil organic matter: its nature, its role in soil formation and in soil fertility (2nd ed.). Amsterdam, The Netherlands: Elsevier. ISBN 978-1-4831-8568-2.
  45. ^ Burns, Richards G.; DeForest, Jared L.; Marxsen, Jürgen; Sinsabaugh, Robert L.; Stromberger, Mary E.; Wallenstein, Matthew D.; Weintraub, Michael N.; Zoppini, Annamaria (2013). « Soil enzymes in a changing environment: current knowledge and future directions ». Soil Biology and Biochemistry. 58: 216‒34. doi:10.1016/j.soilbio.2012.11.009.
  46. ^ Sengupta, Aditi; Kushwaha, Priyanka; Jim, Antonia; Troch, Peter A.; Maier, Raina (2020). « New soil, old plants, and ubiquitous microbes: evaluating the potential of incipient basaltic soil to support native plant growth and influence belowground soil microbial community composition ». Sustainability. 12 (10): 4209. doi:10.3390/su12104209.
  47. ^ Bishop, Janice L.; Murchie, Scott L.; Pieters, Carlé L.; Zent, Aaron P. (2002). « A model for formation of dust, soil, and rock coatings on Mars: physical and chemical processes on the Martian surface ». Journal of Geophysical Research. 107 (E11): 7-1–7-17. Bibcode:2002JGRE..107.5097B. doi:10.1029/2001JE001581.
  48. ^ Navarro-González, Rafael; Rainey, Fred A.; Molina, Paola; Bagaley, Danielle R.; Hollen, Becky J.; de la Rosa, José; Small, Alanna M.; Quinn, Richard C.; Grunthaner, Frank J.; Cáceres, Luis; Gomez-Silva, Benito; McKay, Christopher P. (2003). « Mars-like soils in the Atacama desert, Chile, and the dry limit of microbial life ». Science. 302 (5647): 1018–21. Bibcode:2003Sci…302.1018N. doi:10.1126/science.1089143. PMID 14605363. S2CID 18220447. Retrieved 14 March 2021.
  49. ^ Guo, Yong; Fujimura, Reiko; Sato, Yoshinori; Suda, Wataru; Kim, Seok-won; Oshima, Kenshiro; Hattori, Masahira; Kamijo, Takashi; Narisawa, Kazuhiko; Ohta, Hiroyuki (2014). « Characterization of early microbial communities on volcanic deposits along a vegetation gradient on the island of Miyake, Japan ». Microbes and Environments. 29 (1): 38–49. doi:10.1264/jsme2.ME13142.
  50. ^ Van Schöll, Laura; Smits, Mark M.; Hoffland, Ellis (2006). « Ectomycorrhizal weathering of the soil minerals muscovite and hornblende ». New Phytologist. 171 (4): 805–14. doi:10.1111/j.1469-8137.2006.01790.x. PMID 16918551.
  51. ^ Stretch, Rachelle C.; Viles, Heather A. (2002). « The nature and rate of weathering by lichens on lava flows on Lanzarote ». Geomorphology. 47 (1): 87–94. doi:10.1016/S0169-555X(02)00143-5. Retrieved 21 March 2021.
  52. ^ Dojani, Stephanie; Lakatos, Michael; Rascher, Uwe; Waneck, Wolfgang; Luettge, Ulrich; Büdel, Burkhard (2007). « Nitrogen input by cyanobacterial biofilms of an inselberg into a tropical rainforest in French Guiana ». Flora. 202 (7): 521–29. doi:10.1016/j.flora.2006.12.001. Retrieved 21 March 2021.
  53. ^ Kabala, Cesary; Kubicz, Justyna (2012). « Initial soil development and carbon accumulation on moraines of the rapidly retreating Werenskiold Glacier, SW Spitsbergen, Svalbard archipelago ». Geoderma. 175/176: 9–20. Bibcode:2012Geode.175….9K. doi:10.1016/j.geoderma.2012.01.025. Retrieved 26 May 2019.
  54. ^ Jenny, Hans (1941). Factors of soil formation: a system of qunatitative pedology (PDF). New York: McGraw-Hill. Archived (PDF) from the original on 8 August 2017. Retrieved 21 March 2021.
  55. ^ Ritter, Michael E. « The physical environment: an introduction to physical geography » (PDF). Retrieved 21 March 2021.
  56. ^ Gardner, Catriona M.K.; Laryea, Kofi Buna; Unger, Paul W. (1999). Soil physical constraints to plant growth and crop production (PDF) (1st ed.). Rome: Food and Agriculture Organization of the United Nations. Archived from the original (PDF) on 8 August 2017. Retrieved 24 December 2017.
  57. ^ Six, Johan; Paustian, Keith; Elliott, Edward T.; Combrink, Clay (2000). « Soil structure and organic matter. I. Distribution of aggregate-size classes and aggregate-associated carbon ». Soil Science Society of America Journal. 64 (2): 681–89. Bibcode:2000SSASJ..64..681S. doi:10.2136/sssaj2000.642681x. Retrieved 28 March 2021.
  58. ^ Håkansson, Inge; Lipiec, Jerzy (2000). « A review of the usefulness of relative bulk density values in studies of soil structure and compaction » (PDF). Soil and Tillage Research. 53 (2): 71–85. doi:10.1016/S0167-1987(99)00095-1. S2CID 30045538. Retrieved 28 March 2021.
  59. ^ Schwerdtfeger, W.J. (1965). « Soil resistivity as related to underground corrosion and cathodic protection ». Journal of Research of the National Bureau of Standards. 69C (1): 71–77. doi:10.6028/jres.069c.012.
  60. ^ Tamboli, Prabhakar Mahadeo (1961). The influence of bulk density and aggregate size on soil moisture retention. Ames, Iowa: Iowa State University. Retrieved 28 March 2021.
  61. ^ Wallace, James S.; Batchelor, Charles H. (1997). « Managing water resources for crop production » (PDF). Philosophical Transactions of the Royal Society B: Biological Sciences. 352 (1356): 937–47. doi:10.1098/rstb.1997.0073. PMC 1691982. Retrieved 4 April 2021.
  62. ^ Veihmeyer, Frank J.; Hendrickson, Arthur H. (1927). « Soil-moisture conditions in relation to plant growth ». Plant Physiology. 2 (1): 71–82. doi:10.1104/pp.2.1.71. PMC 439946. PMID 16652508.
  63. ^ Ratliff, Larry F.; Ritchie, Jerry T.; Cassel, D. Keith (1983). « Field-measured limits of soil water availability as related to laboratory-measured properties ». Soil Science Society of America Journal. 47 (4): 770–75. doi:10.2136/sssaj1983.03615995004700040032x. Retrieved 4 April 2021.
  64. ^ Snyman, Henny A.; Venter, W.D.; Van Rensburg, W.L.J.; Opperman, D.P.J. (1987). « Ranking of grass species according to visible wilting order and rate of recovery in the Central Orange Free State ». Journal of the Grassland Society of Southern Africa. 4 (2): 78–81. doi:10.1080/02566702.1987.9648075. Retrieved 5 April 2021.
  65. ^ « Water movement in soils ». Oklahoma State University, Department of Plant and Soil Sciences. Stillwater, Oklahoma. Retrieved 5 April 2021.
  66. ^ Le Bissonnais, Yves (2016). « Aggregate stability and assessment of soil crustability and erodibility. I. Theory and methodology ». European Journal of Soil Science. 67 (1): 11–21. doi:10.1111/ejss.4_12311. Retrieved 5 April 2021.
  67. ^ une b Easton, Zachary M.; Bock, Emily. « Soil and soil water relationships » (PDF). Virginia Tech. hdl:10919/75545. Retrieved 11 April 2021.
  68. ^ Sims, J. Thomas; Simard, Régis R.; Joern, Brad Christopher (1998). « Phosphorus loss in agricultural drainage: historical perspective and current research ». Journal of Environmental Quality. 27 (2): 277–93. doi:10.2134/jeq1998.00472425002700020006x. Retrieved 11 April 2021.
  69. ^ Brooks, R.H.; Corey, Arthur T. (1964). Hydraulic properties of porous media (PDF). Fort Collins, Colorado: Colorado State University. Retrieved 11 April 2021.
  70. ^ McElrone, Andrew J.; Choat, Brendan; Gambetta, Greg A.; Brodersen, Craig R. « Water uptake and transport in vascular plants ». The Nature Education Knowledge Project. Retrieved 6 May 2018.
  71. ^ Steudle, Ernst (2000). « Water uptake by plant roots: an integration of views » (PDF). Plant and Soil. 226 (1): 45–56. doi:10.1023/A:1026439226716. S2CID 3338727. Retrieved 6 May 2018.
  72. ^ Wilcox, Carolyn S.; Ferguson, Joseph W.; Fernandez, George C.J.; Nowak, Robert S. (2004). « Fine root growth dynamics of four Mojave Desert shrubs as related to soil moisture and microsite » (PDF). Journal of Arid Environments. 56 (1): 129–48. Bibcode:2004JArEn..56..129W. doi:10.1016/S0140-1963(02)00324-5. Retrieved 6 May 2018.
  73. ^ Hunter, Albert S.; Kelley, Omer J. (1946). « The extension of plant roots into dry soil ». Plant Physiology. 21 (4): 445–51. doi:10.1104/pp.21.4.445. PMC 437296. PMID 16654059.
  74. ^ Zhang, Yongqiang; Kendy, Eloise; Qiang, Yu; Liu, Changming; Shen, Yanjun; Sun, Hongyong (2004). « Effect of soil water deficit on evapotranspiration, crop yield, and water use efficiency in the North China Plain » (PDF). Agricultural Water Management. 64 (2): 107–22. doi:10.1016/S0378-3774(03)00201-4. Retrieved 6 May 2018.
  75. ^ Oyewole, Olusegun Ayodeji; Inselsbacher, Erich; Näsholm, Torgny (2014). « Direct estimation of mass flow and diffusion of nitrogen compounds in solution and soil » (PDF). New Phytologist. 201 (3): 1056–64. doi:10.1111/nph.12553. PMID 24134319. Retrieved 10 May 2018.
  76. ^ « the GCOS Essential Climate Variables ». GCOS. 2013. Retrieved 5 November 2013.
  77. ^ Brocca, L.; Hasenauer, S.; Lacava, T.; oramarco, T.; Wagner, W.; Dorigo, W.; Matgen, P.; Martínez-Fernández, J.; Llorens, P.; Latron, C.; Martin, C.; Bittelli, M. (2011). « Soil moisture estimation through ASCAT and AMSR-E sensors: An intercomparison and validation study across Europe ». Remote Sensing of Environment. 115 (12): 3390–3408. Bibcode:2011RSEnv.115.3390B. doi:10.1016/j.rse.2011.08.003.
  78. ^ « Soil and water ». Food and Agriculture Organization of the United Nations. Retrieved 10 May 2018.
  79. ^ Petersen, Lis Wollesen; Møldrup, Per; Jacobsen, Ole H.; Rolston, Dennis E. (1996). « Relations between specific surface area and soil physical and chemical properties » (PDF). Soil Science. 161 (1): 9–21. Bibcode:1996SoilS.161….9P. doi:10.1097/00010694-199601000-00003. Retrieved 10 May 2018.
  80. ^ une b Gupta, Satish C.; Larson, William E. (1979). « Estimating soil water retention characteristics from particle size distribution, organic matter percent, and bulk density ». Water Resources Research. 15 (6): 1633–35. Bibcode:1979WRR….15.1633G. CiteSeerX 10.1.1.475.497. doi:10.1029/WR015i006p01633.
  81. ^ « Soil Water Potential ». AgriInfo.in. Archived from the original on 17 August 2017. Retrieved 15 March 2019.
  82. ^ Savage, Michael J.; Ritchie, Joe T.; Bland, William L.; Dugas, William A. (1996). « Lower limit of soil water availability » (PDF). Agronomy Journal. 88 (4): 644–51. doi:10.2134/agronj1996.00021962008800040024x. Retrieved 12 May 2018.
  83. ^ Al-Ani, Tariq; Bierhuizen, Johan Frederik (1971). « Stomatal resistance, transpiration, and relative water content as influenced by soil moisture stress » (PDF). Acta Botanica Neerlandica. 20 (3): 318–26. doi:10.1111/j.1438-8677.1971.tb00715.x. Retrieved 12 May 2018.
  84. ^ Rawls, W. J.; Brakensiek, D. L.; Saxtonn, K. E. (1982). « Estimation of Soil Water Properties » (PDF). Transactions of the ASAE. 25 (5): 1316–1320. doi:10.13031/2013.33720. Retrieved 17 March 2019.
  85. ^ « Soil water movement: saturated and unsaturated flow and vapour movement, soil moisture constants and their importance in irrigation » (PDF). Tamil Nadu Agricultural University. Retrieved 19 May 2018.
  86. ^ Cueto-Felgueroso, Luis; Juanes, Ruben (2008). « Nonlocal interface dynamics and pattern formation in gravity-driven unsaturated flow through porous media » (PDF). Physical Review Letters. 101 (24): 244504. Bibcode:2008PhRvL.101x4504C. doi:10.1103/PhysRevLett.101.244504. PMID 19113626. S2CID 21874968. Retrieved 21 May 2018.
  87. ^ « Finger flow in coarse soils ». Cornell University. Retrieved 21 May 2018.
  88. ^ Ghestem, Murielle; Sidle, Roy C.; Stokes, Alexia (2011). « The influence of plant root systems on subsurface flow: implications for slope stability ». BioScience. 61 (11): 869–79. doi:10.1525/bio.2011.61.11.6.
  89. ^ Bartens, Julia; Day, Susan D.; Harris, J. Roger; Dove, Joseph E.; Wynn, Theresa M. (2008). « Can urban tree roots improve infiltration through compacted subsoils for stormwater management? » (PDF). Journal of Environmental Quality. 37 (6): 2048–57. doi:10.2134/jeq2008.0117. PMID 18948457. Retrieved 21 May 2018.
  90. ^ Zhang, Guohua; Feng, Gary; Li, Xinhu; Xie, Congbao; P, Xiaoyu (2017). « Flood effect on groundwater recharge on a typical silt loam soil ». L’eau. 9 (7): 523. doi:10.3390/w9070523.
  91. ^ Nielsen, Donald R.; Biggar, James W.; Erh, Koon T. (1973). « Spatial variability of field-measured soil-water properties ». Hilgardia. 42 (7): 215–59. doi:10.3733/hilg.v42n07p215.
  92. ^ Rimon, Yaara; Dahan, Ofer; Nativ, Ronit; Geyer, Stefan (2007). « Water percolation through the deep vadose zone and groundwater recharge: preliminary results based on a new vadose zone monitoring system ». Water Resources Research. 43 (5): W05402. Bibcode:2007WRR….43.5402R. doi:10.1029/2006WR004855.
  93. ^ Weiss, Peter T.; LeFevre, Greg; Gulliver, John S. (2008). « Contamination of soil and groundwater due to stormwater infiltration practices: a literature review ». CiteSeerX 10.1.1.410.5113.
  94. ^ Hagedorn, Charles; Hansen, Debra T.; Simonson, Gerald H. (1978). « Survival and movement of fecal indicator bacteria in soil under conditions of saturated flow » (PDF). Journal of Environmental Quality. 7 (1): 55–59. doi:10.2134/jeq1978.00472425000700010011x. S2CID 774611. Retrieved 24 June 2018.
  95. ^ Ng, Charles W.W.; Pang, Wenyan (2000). « Influence of stress state on soil-water characteristics and slope stability » (PDF). Journal of Geotechnical and Geoenvironmental Engineering. 126 (2): 157–66. doi:10.1061/(ASCE)1090-0241(2000)126:2(157). Retrieved 1 July 2018.
  96. ^ Richards, L.A. (1931). « Capillary conduction of liquids through porous mediums ». Physics. 1 (5): 318–333. Bibcode:1931Physi…1..318R. doi:10.1063/1.1745010.
  97. ^ Richardson, Lewis Fry (1922). Weather prediction by numerical process. Cambridge, The University press. p. 262.
  98. ^ Ogden, Fred L.; Allen, Myron B.; Lai, Wencong; Zhu, Julian; Douglas, Craig C.; Seo, Mookwon; Talbot, Cary A. (2017). « The Soil Moisture Velocity Equation ». J. Adv. Modeling Earth Syst. 9 (2): 1473–1487. Bibcode:2017JAMES…9.1473O. doi:10.1002/2017MS000931.
  99. ^ Talbot, Cary A.; Ogden, Fred L. (2008). « A method for computing infiltration and redistribution in a discretized moisture content domain ». Water Resour. Res. 44 (8): 8. Bibcode:2008WRR….44.8453T. doi:10.1029/2008WR006815.
  100. ^ Ogden, Fred L.; Lai, Wencong; Steinke, Robert C.; Zhu, Julian; Talbot, Cary A.; Wilson, John L. (2015). « A new general 1-D vadose zone solution method ». Water Resour. Res. 51 (6): 4282–4300. Bibcode:2015WRR….51.4282O. doi:10.1002/2015WR017126.
  101. ^ Šimůnek, J.; Saito, H.; Sakai, M.; van Genuchten, M. Th. (2013). « The HYDRUS-1D Software Package for Simulating the One-Dimensional Movement of Water, Heat, and Multiple Solutes in Variably-Saturated Media ». Retrieved 15 March 2019.
  102. ^ Bouma, Johan (1981). « Soil morphology and preferential flow along macropores » (PDF). Geoderma. 3 (4): 235–50. doi:10.1016/0378-3774(81)90009-3. Retrieved 1 July 2018.
  103. ^ Luo, Lifang; Lin, Henry; Halleck, Phil (2008). « Quantifying soil structure and preferential flow in intact soil Using X-ray computed tomography ». Soil Science Society of America Journal. 72 (4): 1058–69. Bibcode:2008SSASJ..72.1058L. CiteSeerX 10.1.1.455.2567. doi:10.2136/sssaj2007.0179.
  104. ^ Beven, Keith; Germann, Peter (2013). « Macropores and water flow in soils revisited » (PDF). Water Resources Research. 49 (6): 3071–92. Bibcode:2013WRR….49.3071B. doi:10.1002/wrcr.20156.
  105. ^ Aston, M.J.; Lawlor, David W. (1979). « The relationship between transpiration, root water uptake, and leaf water potential » (PDF). Journal of Experimental Botany. 30 (1): 169–81. doi:10.1093/jxb/30.1.169. Retrieved 8 July 2018.
  106. ^ Powell, D.B.B. (1978). « Regulation of plant water potential by membranes of the endodermis in young roots » (PDF). Plant, Cell and Environment. 1 (1): 69–76. doi:10.1111/j.1365-3040.1978.tb00749.x. Retrieved 7 July 2018.
  107. ^ Irvine, James; Perks, Michael P.; Magnani, Federico; Grace, John (1998). « The response of Pinus sylvestris to drought: stomatal control of transpiration and hydraulic conductance ». Tree Physiology. 18 (6): 393–402. doi:10.1093/treephys/18.6.393. PMID 12651364.
  108. ^ Jackson, Robert B.; Sperry, John S.; Dawson, Todd E. (2000). « Root water uptake and transport: using physiological processes in global predictions » (PDF). Trends in Plant Science. 5 (11): 482–88. doi:10.1016/S1360-1385(00)01766-0. PMID 11077257. S2CID 8311441. Retrieved 8 July 2018.
  109. ^ Steudle, Ernst (2000). « Water uptake by plant roots: an integration of views » (PDF). Plant and Soil. 226 (1): 45–56. doi:10.1023/A:1026439226716. S2CID 3338727. Retrieved 8 July 2018.
  110. ^ Kaufmann, Merrill R.; Eckard, Alan N. (1971). « Evaluation of water stress control with polyethylene glycols by analysis of guttation ». Plant Physiology. 47 (4): 453–6. doi:10.1104/pp.47.4.453. PMC 396708. PMID 16657642.
  111. ^ Kramer, Paul J.; Coile, Theodore S. (1940). « An estimation of the volume of water made available by root extension ». Plant Physiology. 15 (4): 743–47. doi:10.1104/pp.15.4.743. PMC 437871. PMID 16653671.
  112. ^ Lynch, Jonathan (1995). « Root architecture and plant productivity ». Plant Physiology. 109 (1): 7–13. doi:10.1104/pp.109.1.7. PMC 157559. PMID 12228579.
  113. ^ Comas, Louise H.; Eissenstat, David M.; Lakso, Alan N. (2000). « Assessing root death and root system dynamics in a study of grape canopy pruning ». New Phytologist. 147 (1): 171–78. doi:10.1046/j.1469-8137.2000.00679.x.
  114. ^ Schlesinger, William H.; Jasechko, Scott (2014). « Transpiration in the global water cycle » (PDF). Agricultural and Forest Meteorology. 189/190: 115–17. Bibcode:2014AgFM..189..115S. doi:10.1016/j.agrformet.2014.01.011. Retrieved 22 July 2018.
  115. ^ Erie, Leonard J.; French, Orrin F.; Harris, Karl (1968). Consumptive use of water by crops in Arizona (PDF). Tucson, Arizona: The University of Arizona. Retrieved 15 July 2018.
  116. ^ Tolk, Judy A.; Howell, Terry A.; Evett, Steve R. (1999). « Effect of mulch, irrigation, and soil type on water use and yield of maize » (PDF). Soil and Tillage Research. 50 (2): 137–47. doi:10.1016/S0167-1987(99)00011-2. Retrieved 15 July 2018.
  117. ^ Qi, Jingen; Marshall, John D.; Mattson, Kim G. (1994). « High soil carbon dioxide concentrations inhibit root respiration of Douglas fir ». New Phytologist. 128 (3): 435–42. doi:10.1111/j.1469-8137.1994.tb02989.x.
  118. ^ Karberg, Noah J.; Pregitzer, Kurt S.; King, John S.; Friend, Aaron L.; Wood, James R. (2005). « Soil carbon dioxide partial pressure and dissolved inorganic carbonate chemistry under elevated carbon dioxide and ozone » (PDF). Oecologia. 142 (2): 296–306. Bibcode:2005Oecol.142..296K. doi:10.1007/s00442-004-1665-5. PMID 15378342. S2CID 6161016. Retrieved 26 August 2018.
  119. ^ Chang, H.T.; Loomis, W.E. (1945). « Effect of carbon dioxide on absorption of water and nutrients by roots ». Plant Physiology. 20 (2): 221–32. doi:10.1104/pp.20.2.221. PMC 437214. PMID 16653979.
  120. ^ McDowell, Nate J.; Marshall, John D.; Qi, Jingen; Mattson, Kim (1999). « Direct inhibition of maintenance respiration in western hemlock roots exposed to ambient soil carbon dioxide concentrations » (PDF). Tree Physiology. 19 (9): 599–605. doi:10.1093/treephys/19.9.599. PMID 12651534. Retrieved 22 July 2018.
  121. ^ Xu, Xia; Nieber, John L.; Gupta, Satish C. (1992). « Compaction effect on the gas diffusion coefficient in soils » (PDF). Soil Science Society of America Journal. 56 (6): 1743–50. Bibcode:1992SSASJ..56.1743X. doi:10.2136/sssaj1992.03615995005600060014x. Retrieved 29 July 2018.
  122. ^ une b Smith, Keith A.; Ball, Tom; Conen, Franz; Dobbie, Karen E.; Massheder, Jonathan; Rey, Ana (2003). « Exchange of greenhouse gases between soil and atmosphere: interactions of soil physical factors and biological processes » (PDF). European Journal of Soil Science. 54 (4): 779–91. doi:10.1046/j.1351-0754.2003.0567.x. S2CID 18442559. Retrieved 5 August 2018.
  123. ^ Ruser, Reiner; Flessa, Heiner; Russow, Rolf; Schmidt, G.; Buegger, Franz; Munch, J.C. (2006). « Emission of N2O, N2 and CO2 from soil fertilized with nitrate: effect of compaction, soil moisture and rewetting » (PDF). Soil Biology and Biochemistry. 38 (2): 263–74. doi:10.1016/j.soilbio.2005.05.005. Retrieved 5 August 2018.
  124. ^ Hartmann, Adrian A.; Buchmann, Nina; Niklaus, Pascal A. (2011). « A study of soil methane sink regulation in two grasslands exposed to drought and N fertilization » (PDF). Plant and Soil. 342 (1/2): 265–75. doi:10.1007/s11104-010-0690-x. hdl:20.500.11850/34759. S2CID 25691034. Retrieved 12 August 2018.
  125. ^ Moore, Tim R.; Dalva, Moshe (1993). « The influence of temperature and water table position on carbon dioxide and methane emissions from laboratory columns of peatland soils » (PDF). Journal of Soil Science. 44 (4): 651–64. doi:10.1111/j.1365-2389.1993.tb02330.x. Retrieved 12 August 2018.
  126. ^ Hiltpold, Ivan; Toepfer, Stefan; Kuhlmann, Ulrich; Turlings, Ted C.J. (2010). « How maize root volatiles affect the efficacy of entomopathogenic nematodes in controlling the western corn rootworm? » (PDF). Chemoecology. 20 (2): 155–62. doi:10.1007/s00049-009-0034-6. S2CID 30214059. Retrieved 12 August 2018.
  127. ^ Ryu, Choong-Min; Farag, Mohamed A.; Hu, Chia-Hui; Reddy, Munagala S.; Wei, Han-Xun; Paré, Paul W.; Kloepper, Joseph W. (2003). « Bacterial volatiles promote growth in Arabidopsis » (PDF). Proceedings of the National Academy of Sciences of the United States of America. 100 (8): 4927–32. Bibcode:2003PNAS..100.4927R. doi:10.1073/pnas.0730845100. PMC 153657. PMID 12684534. Retrieved 12 August 2018.
  128. ^ Hung, Richard; Lee, Samantha; Bennett, Joan W. (2015). « Fungal volatile organic compounds and their role in ecosystems » (PDF). Applied Microbiology and Biotechnology. 99 (8): 3395–405. doi:10.1007/s00253-015-6494-4. PMID 25773975. S2CID 14509047. Retrieved 12 August 2018.
  129. ^ Purrington, Foster Forbes; Kendall, Paricia A.; Bater, John E.; Stinner, Benjamin R. (1991). « Alarm pheromone in a gregarious poduromorph collembolan (Collembola: Hypogastruridae) » (PDF). Great Lakes Entomologist. 24 (2): 75–78. Retrieved 12 August 2018.
  130. ^ Badri, Dayakar V.; Weir, Tiffany L.; Van der Lelie, Daniel; Vivanco, Jorge M (2009). « Rhizosphere chemical dialogues: plant–microbe interactions ». Current Opinion in Biotechnology. 20 (6): 642–50. doi:10.1016/j.copbio.2009.09.014. PMID 19875278.
  131. ^ Salmon, Sandrine; Ponge, Jean-François (2001). « Earthworm excreta attract soil springtails: laboratory experiments on Heteromurus nitidus (Collembola: Entomobryidae) » (PDF). Soil Biology and Biochemistry. 33 (14): 1959–69. doi:10.1016/S0038-0717(01)00129-8. Retrieved 19 August 2018.
  132. ^ Lambers, Hans; Mougel, Christophe; Jaillard, Benoît; Hinsinger, Philipe (2009). « Plant-microbe-soil interactions in the rhizosphere: an evolutionary perspective » (PDF). Plant and Soil. 321 (1/2): 83–115. doi:10.1007/s11104-009-0042-x. S2CID 6840457. Retrieved 19 August 2018.
  133. ^ Peñuelas, Josep; Asensio, Dolores; Tholl, Dorothea; Wenke, Katrin; Rosenkranz, Maaria; Piechulla, Birgit; Schnitzler, Jörg-Petter (2014). « Biogenic volatile emissions from the soil ». Plant, Cell and Environment. 37 (8): 1866–91. doi:10.1111/pce.12340. PMID 24689847.
  134. ^ Buzuleciu, Samuel A.; Crane, Derek P.; Parker, Scott L. (2016). « Scent of disinterred soil as an olfactory cue used by raccoons to locate nests of diamond-backed terrapins (Malaclemys terrapin) » (PDF). Herpetological Conservation and Biology. 11 (3): 539–51. Retrieved 19 August 2018.
  135. ^ Saxton, Keith E.; Rawls, Walter J. (2006). « Soil water characteristic estimates by texture and organic matter for hydrologic solutions » (PDF). Soil Science Society of America Journal. 70 (5): 1569–78. Bibcode:2006SSASJ..70.1569S. CiteSeerX 10.1.1.452.9733. doi:10.2136/sssaj2005.0117. S2CID 16826314. Retrieved 2 September 2018.
  136. ^ College of Tropical Agriculture and Human Resources. « Soil Mineralogy ». cms.ctahr.hawaii.edu/. University of Hawai‘i at Mānoa. Retrieved 2 September 2018.
  137. ^ Sposito, Garrison (1984). The surface chemistry of soils (PDF). New York, New York: Oxford University Press. Retrieved 21 April 2019.
  138. ^ Wynot, Christopher. « Theory of diffusion in colloidal suspensions » (PDF). Retrieved 21 April 2019.
  139. ^ Sposito, Garrison; Skipper, Neal T.; Sutton, Rebecca; Park, Sung-Ho; Soper, Alan K.; Greathouse, Jeffery A. (1999). « Surface geochemistry of the clay minerals ». Proceedings of the National Academy of Sciences of the United States of America. 96 (7): 3358–64. Bibcode:1999PNAS…96.3358S. doi:10.1073/pnas.96.7.3358. PMC 34275. PMID 10097044.
  140. ^ Bickmore, Barry R.; Rosso, Kevin M.; Nagy, Kathryn L.; Cygan, Randall T.; Tadanier, Christopher J. (2003). « Ab initio determination of edge surface structures for dioctahedral 2:1 phyllosilicates: implications for acid-base reactivity » (PDF). Clays and Clay Minerals. 51 (4): 359–71. Bibcode:2003CCM….51..359B. doi:10.1346/CCMN.2003.0510401. S2CID 97428106. Retrieved 21 April 2019.
  141. ^ Rajamathi, Michael; Thomas, Grace S.; Kamath, P. Vishnu (2001). « The many ways of making anionic clays » (PDF). Journal of Chemical Sciences. 113 (5–6): 671–80. doi:10.1007/BF02708799. S2CID 97507578. Retrieved 27 April 2019.
  142. ^ Moayedi, Hossein; Kazemian, Sina (2012). « Zeta potentials of suspended humus in multivalent cationic saline solution and its effect on electro-osomosis behavior » (PDF). Journal of Dispersion Science and Technology. 34 (2): 283–94. doi:10.1080/01932691.2011.646601. S2CID 94333872. Retrieved 27 April 2019.
  143. ^ Pettit, Robert E. « Organic matter, humus, humate, humic acid, fulvic acid and humin: their importance in soil fertility and plant health » (PDF). Retrieved 27 April 2019.
  144. ^ Diamond, Sidney; Kinter, Earl B. (1965). « Mechanisms of soil-lime stabilization: an interpretative review » (PDF). Highway Research Record. 92: 83–102. Retrieved 27 April 2019.
  145. ^ Woodruff, Clarence M. (1955). « The energies of replacement of calcium by potassium in soils » (PDF). Soil Science Society of America Journal. 19 (2): 167–71. Bibcode:1955SSASJ..19..167W. doi:10.2136/sssaj1955.03615995001900020014x. Retrieved 28 April 2019.
  146. ^ Fronæus, Sture (1953). « On the application of the mass action law to cation exchange equilibria » (PDF). Acta Chemica Scandinavica. 7: 469–80. doi:10.3891/acta.chem.scand.07-0469. Retrieved 4 May 2019.
  147. ^ Bolland, Mike D. A.; Posner, Alan M.; Quirk, James P. (1980). « pH-independent and pH-dependent surface charges on kaolinite ». Clays and Clay Minerals. 28 (6): 412–18. Bibcode:1980CCM….28..412B. doi:10.1346/CCMN.1980.0280602.
  148. ^ Silber, Avner; Levkovitch, Irit; Graber, Ellen R. (2010). « pH-dependent mineral release and surface properties of cornstraw biochar: agronomic implications » (PDF). Environmental Science and Technology. 44 (24): 9318–23. Bibcode:2010EnST…44.9318S. doi:10.1021/es101283d. PMID 21090742. Retrieved 4 May 2019.
  149. ^ Dakora, Felix D.; Phillips, Donald D. (2002). « Root exudates as mediators of mineral acquisition in low-nutrient environments » (PDF). Plant and Soil. 245: 35–47. doi:10.1023/A:1020809400075. S2CID 3330737. Archived from the original (PDF) on 19 August 2019. Retrieved 25 July 2019.
  150. ^ Singh, Jamuna Sharan; Raghubanshi, Akhilesh Singh; Singh, Raj S.; Srivastava, S. C. (1989). « Microbial biomass acts as a source of plant nutrient in dry tropical forest and savanna » (PDF). Nature. 338 (6215): 499–500. Bibcode:1989Natur.338..499S. doi:10.1038/338499a0. S2CID 4301023. Retrieved 12 May 2019.
  151. ^ Szatanik-Kloc, Alicja; Szerement, Justyna; Józefaciuk, Grzegorz (2017). « The role of cell walls and pectins in cation exchange and surface area of plant roots » (PDF). Journal of Plant Physiology. 215: 85–90. doi:10.1016/j.jplph.2017.05.017. PMID 28600926. Retrieved 25 July 2019.
  152. ^ Gu, Baohua; Schulz, Robert K. (1991). « Anion retention in soil: possible application to reduce migration of buried technetium and iodine, a review » (PDF). doi:10.2172/5980032. Retrieved 19 May 2019.
  153. ^ Sollins, Phillip; Robertson, G. Philip; Uehara, Goro (1988). « Nutrient mobility in variable- and permanent-charge soils » (PDF). Biogeochemistry. 6 (3): 181–99. doi:10.1007/BF02182995. S2CID 4505438. Retrieved 19 May 2019.
  154. ^ Sanders, W. M. H. (1964). « Extraction of soil phosphate by anion-exchange membrane ». New Zealand Journal of Agricultural Research. 7 (3): 427–31. doi:10.1080/00288233.1964.10416423.
  155. ^ Hinsinger, Philippe (2001). « Bioavailability of soil inorganic P in the rhizosphere as affected by root-induced chemical changes: a review » (PDF). Plant and Soil. 237 (2): 173–95. doi:10.1023/A:1013351617532. S2CID 8562338. Retrieved 19 May 2019.
  156. ^ Chichester, Fredrick Wesley; Harward, Moyle E.; Youngberg, Chester T. (1970). « pH dependent ion exchange properties of soils and clays from Mazama pumice ». Clays and Clay Minerals. 18 (2): 81–90. Bibcode:1970CCM….18…81C. doi:10.1346/CCMN.1970.0180203.
  157. ^ Robertson, Bryan. « pH requirements of freshwater aquatic life » (PDF). Retrieved 26 May 2019.
  158. ^ Chang, Raymond, ed. (2010). Chemistry (PDF) (10th ed.). New York, New York: McGraw-Hill. p. 663. ISBN 9780073511092. Retrieved 26 May 2019.
  159. ^ Singleton, Peter L.; Edmeades, Doug C.; Smart, R. E.; Wheeler, David M. (2001). « The many ways of making anionic clays » (PDF). Journal of Chemical Sciences. 113 (5–6): 671–80. doi:10.1007/BF02708799. S2CID 97507578. Retrieved 27 April 2019.
  160. ^ Läuchli, André; Grattan, Steve R. (2012). « Soil pH extremes » (PDF). In Shabala, Sergey (ed.). Plant stress physiology (1st ed.). Wallingford, United Kingdom: CAB International. pp. 194–209. ISBN 978-1845939953. Retrieved 2 June 2019.
  161. ^ Calmano, Wolfgang; Hong, Jihua; Förstner, Ulrich (1993). « Binding and mobilization of heavy metals in contaminated sediments affected by pH and redox potential » (PDF). Water Science and Technology. 28 (8–9): 223–35. doi:10.2166/wst.1993.0622. Retrieved 9 June 2019.
  162. ^ Ren, Xiaoya; Zeng, Guangming; Tang, Lin; Wang, Jingjing; Wan, Jia; Liu, Yani; Yu, Jiangfang; Yi, Huan; Ye, Shujing; Deng, Rui (2018). « Sorption, transport and biodegradation: an insight into bioavailability of persistent organic pollutants in soil » (PDF). Science of the Total Environment. 610–611: 1154–63. Bibcode:2018ScTEn.610.1154R. doi:10.1016/j.scitotenv.2017.08.089. PMID 28847136. Retrieved 9 June 2019.
  163. ^ Ponge, Jean-François (2003). « Humus forms in terrestrial ecosystems: a framework to biodiversity » (PDF). Soil Biology and Biochemistry. 35 (7): 935–45. CiteSeerX 10.1.1.467.4937. doi:10.1016/S0038-0717(03)00149-4. Retrieved 9 June 2019.
  164. ^ Fujii, Kazumichi (2003). « Soil acidification and adaptations of plants and microorganisms in Bornean tropical forests ». Ecological Research. 29 (3): 371–81. doi:10.1007/s11284-014-1144-3.
  165. ^ Kauppi, Pekka; Kämäri, Juha; Posch, Maximilian; Kauppi, Lea (1986). « Acidification of forest soils: model development and application for analyzing impacts of acidic deposition in Europe » (PDF). Ecological Modelling. 33 (2–4): 231–53. doi:10.1016/0304-3800(86)90042-6. Retrieved 10 June 2019.
  166. ^ Andriesse, J. P. (1969). « A study of the environment and characteristics of tropical podzols in Sarawak (East-Malaysia) » (PDF). Geoderma. 2 (3): 201–27. Bibcode:1969Geode…2..201A. doi:10.1016/0016-7061(69)90038-X. Retrieved 10 June 2019.
  167. ^ Rengasamy, Pichu (2006). « World salinization with emphasis on Australia » (PDF). Journal of Experimental Botany. 57 (5): 1017–23. doi:10.1093/jxb/erj108. PMID 16510516. Retrieved 16 June 2019.
  168. ^ Arnon, Daniel I.; Johnson, Clarence M. (1942). « Influence of hydrogen ion concentration on the growth of higher plants under controlled conditions ». Plant Physiology. 17 (4): 525–39. doi:10.1104/pp.17.4.525. PMC 438054. PMID 16653803.
  169. ^ Chaney, Rufus L.; Brown, John C.; Tiffin, Lee O. (1972). « Obligatory reduction of ferric chelates in iron uptake by soybeans ». Plant Physiology. 50 (2): 208–13. doi:10.1104/pp.50.2.208. PMC 366111. PMID 16658143.
  170. ^ Ahmad, Sagheer; Ghafoor, Abdul; Qadir, Manzoor; Aziz, M. Abbas (2006). « Amelioration of a calcareous saline-sodic soil by gypsum application and different crop rotations » (PDF). International Journal of Agriculture and Biology. 8 (2): 142–46. Retrieved 16 June 2019.
  171. ^ McFee, William W.; Kelly, J. M.; Beck, R. H. (1976). « Acid precipitation effects on soil pH and base saturation of exchange sites » (PDF). USDA Forest Service, Northeastern Research Station, General Technical Reports. NE-23 (3): 725–35. Bibcode:1977WASP….7..401M. CiteSeerX 10.1.1.549.37. doi:10.1007/BF00284134. S2CID 95001535. Retrieved 23 June 2019.
  172. ^ Farina, Martin Patrick W.; Sumner, Malcolm E.; Plank, C. O.; Letzsch, W. Stephen (1980). « Exchangeable aluminum and pH as indicators of lime requirement for corn » (PDF). Soil Science Society of America Journal. 44 (5): 1036–41. Bibcode:1980SSASJ..44.1036F. doi:10.2136/sssaj1980.03615995004400050033x. Retrieved 30 June 2019.
  173. ^ Sposito, Garrison; Skipper, Neal T.; Sutton, Rebecca; Park, Sun-Ho; Soper, Alan K.; Greathouse, Jeffery A. (1999). « Surface geochemistry of the clay minerals » (PDF). Proceedings of the National Academy of Sciences of the United States of America. 96 (7): 3358–64. Bibcode:1999PNAS…96.3358S. doi:10.1073/pnas.96.7.3358. PMC 34275. PMID 10097044. Retrieved 7 July 2019.
  174. ^ Sparks, Donald L. « Soil buffering and acidic and basic soils » (PDF). University of California, Department of Land, Air, and Water Resources. Retrieved 7 July 2019.
  175. ^ Ulrich, Bernhard (1983). « Soil acidity and its relations to acid deposition ». In Ulrich, Bernhard; Pankrath, Jürgen (eds.). Effects of accumulation of air pollutants in forest ecosystems (1st ed.). Dordrecht, The Netherlands: D. Reidel Publishing Company. pp. 127–46. ISBN 978-94-009-6985-8.
  176. ^ une b Brady, Nyle C.; Weil, Ray R. (2008). The nature and properties of soils (14th ed.). Upper Saddle River: Pearson.
  177. ^ Van der Ploeg, Rienk R.; Böhm, Wolfgang; Kirkham, Mary Beth (1999). « On the origin of the theory of mineral nutrition of plants and the Law of the Minimum ». Soil Science Society of America Journal. 63 (5): 1055–62. Bibcode:1999SSASJ..63.1055V. CiteSeerX 10.1.1.475.7392. doi:10.2136/sssaj1999.6351055x.
  178. ^ Knecht, Magnus F.; Göransson, Anders (2004). « Terrestrial plants require nutrients in similar proportions ». Tree Physiology. 24 (4): 447–60. doi:10.1093/treephys/24.4.447. PMID 14757584.
  179. ^ une b Roy, R. N.; Finck, Arnold; Blair, Graeme J.; Tandon, Hari Lal Singh (2006). « Chapter 4: Soil fertility and crop production » (PDF). Plant nutrition for food security: a guide for integrated nutrient management. Rome, Italy: Food and Agriculture Organization of the United Nations. pp. 43–90. ISBN 978-92-5-105490-1. Retrieved 21 July 2019.
  180. ^ Parfitt, Roger L.; Giltrap, Donna J.; Whitton, Joe S. (1995). « Contribution of organic matter and clay minerals to the cation exchange capacity of soil » (PDF). Communications in Soil Science and Plant Analysis. 26 (9–10): 1343–55. doi:10.1080/00103629509369376. Retrieved 28 July 2019.
  181. ^ Hajnos, Mieczyslaw; Jozefaciuk, Grzegorz; Sokołowska, Zofia; Greiffenhagen, Andreas; Wessolek, Gerd (2003). « Water storage, surface, and structural properties of sandy forest humus horizons » (PDF). Journal of Plant Nutrition and Soil Science. 166 (5): 625–34. doi:10.1002/jpln.200321161. Retrieved 28 July 2019.
  182. ^ Pimentel, David; Harvey, Celia; Resosudarmo, Resosudarmo; Sinclair, K.; Kurz, D.; McNair, M.; Crist, S.; Shpritz, L.; Fitton, L.; Saffouri, R.; Blair, R. (1995). « Environmental and economic costs of soil erosion and conservation benefits » (PDF). Science. 267 (5201): 1117–23. Bibcode:1995Sci…267.1117P. doi:10.1126/science.267.5201.1117. PMID 17789193. S2CID 11936877. Archived from the original (PDF) on 13 December 2016. Retrieved 23 February 2020.
  183. ^ Schnürer, Johan; Clarholm, Marianne; Rosswall, Thomas (1985). « Microbial biomass and activity in an agricultural soil with different organic matter contents » (PDF). Soil Biology and Biochemistry. 17 (5): 611–18. doi:10.1016/0038-0717(85)90036-7. Retrieved 1 March 2020.
  184. ^ Sparling, Graham P. (1992). « Ratio of microbial biomass carbon to soil organic carbon as a sensitive indicator of changes in soil organic matter » (PDF). Australian Journal of Soil Research. 30 (2): 195–207. doi:10.1071/SR9920195. Retrieved 1 March 2020.
  185. ^ Varadachari, Chandrika; Ghosh, Kunal (1984). « On humus formation » (PDF). Plant and Soil. 77 (2): 305–13. doi:10.1007/BF02182933. S2CID 45102095. Retrieved 8 March 2020.
  186. ^ Prescott, Cindy E. (2010). « Litter decomposition: what controls it and how can we alter it to sequester more carbon in forest soils? » (PDF). Biogeochemistry. 101 (1): 133–49. doi:10.1007/s10533-010-9439-0. S2CID 93834812. Retrieved 1 March 2020.
  187. ^ Lehmann, Johannes; Kleber, Markus (2015). « The contentious nature of soil organic matter ». Nature. 528 (7580): 60–68. Bibcode:2015Natur.528…60L. doi:10.1038/nature16069. PMID 26595271. S2CID 205246638.
  188. ^ une b Piccolo, Alessandro (2002). « The supramolecular structure of humic substances: a novel understanding of humus chemistry and implications in soil science » (PDF). Advances in Agronomy. 75: 57–134. doi:10.1016/S0065-2113(02)75003-7. ISBN 9780120007936. Retrieved 15 March 2020.
  189. ^ Scheu, Stefan (2002). « The soil food web: structure and perspectives » (PDF). European Journal of Soil Biology. 38 (1): 11–20. doi:10.1016/S1164-5563(01)01117-7. Retrieved 8 March 2020.
  190. ^ une b Foth, Henry D. (1984). Fundamentals of soil science (7th ed.). New York, New York: Wiley. p. 151. ISBN 978-0-471-88926-7.
  191. ^ Ponge, Jean-François (2003). « Humus forms in terrestrial ecosystems: a framework to biodiversity ». Soil Biology and Biochemistry. 35 (7): 935–45. CiteSeerX 10.1.1.467.4937. doi:10.1016/s0038-0717(03)00149-4. Retrieved 22 March 2020.
  192. ^ Pettit, Robert E. « Organic matter, humus, humate, humic acid, fulvic acid and humin: their importance in soil fertility and plant health » (PDF). Retrieved 5 April 2020.
  193. ^ Ji, Rong; Kappler, Andreas; Brune, Andreas (2000). « Transformation and mineralization of synthetic 14C-labeled humic model compounds by soil-feeding termites ». Soil Biology and Biochemistry. 32 (8–9): 1281–91. CiteSeerX 10.1.1.476.9400. doi:10.1016/S0038-0717(00)00046-8.
  194. ^ Gilluly, James; Waters, Aaron Clement; Woodford, Alfred Oswald (1975). Principles of geology (4th ed.). San Francisco, California: W.H. Freeman. p. 216. ISBN 978-0-7167-0269-6.
  195. ^ une b Piccolo, Alessandro (1996). « Humus and soil conservation ». In Piccolo, Alessandro (ed.). Humic substances in terrestrial ecosystems. Amsterdam, The Netherlands: Elsevier. pp. 225–64. doi:10.1016/B978-044481516-3/50006-2. ISBN 978-0-444-81516-3. Retrieved 22 March 2020.
  196. ^ Varadachari, Chandrika; Ghosh, Kunal (1984). « On humus formation ». Plant and Soil. 77 (2): 305–13. doi:10.1007/BF02182933. S2CID 45102095. Retrieved 5 April 2020.
  197. ^ Mendonça, Eduardo S.; Rowell, David L. (1996). « Mineral and organic fractions of two oxisols and their influence on effective cation-exchange capacity ». Soil Science Society of America Journal. 60 (6): 1888–92. Bibcode:1996SSASJ..60.1888M. doi:10.2136/sssaj1996.03615995006000060038x. Retrieved 12 April 2020.
  198. ^ Heck, Tobias; Faccio, Greta; Richter, Michael; Thöny-Meyer, Linda (2013). « Enzyme-catalyzed protein crosslinking ». Applied Microbiology and Biotechnology. 97 (2): 461–75. doi:10.1007/s00253-012-4569-z. PMC 3546294. PMID 23179622.
  199. ^ Lynch, D. L.; Lynch, C. C. (1958). « Resistance of protein–lignin complexes, lignins and humic acids to microbial attack ». Nature. 181 (4621): 1478–79. Bibcode:1958Natur.181.1478L. doi:10.1038/1811478a0. PMID 13552710. S2CID 4193782.
  200. ^ Dawson, Lorna A.; Hillier, Stephen (2010). « Measurement of soil characteristics for forensic applications » (PDF). Surface and Interface Analysis. 42 (5): 363–77. doi:10.1002/sia.3315. Retrieved 19 April 2020.
  201. ^ Manjaiah, K.M.; Kumar, Sarvendra; Sachdev, M. S.; Sachdev, P.; Datta, S. C. (2010). « Study of clay–organic complexes ». Current Science. 98 (7): 915–21. Retrieved 19 April 2020.
  202. ^ Theng, Benny K.G. (1982). « Clay-polymer interactions: summary and perspectives ». Clays and Clay Minerals. 30 (1): 1–10. Bibcode:1982CCM….30….1T. CiteSeerX 10.1.1.608.2942. doi:10.1346/CCMN.1982.0300101. S2CID 98176725.
  203. ^ Tietjen, Todd; Wetzel, Robert G. (2003). « Extracellular enzyme-clay mineral complexes: enzyme adsorption, alteration of enzyme activity, and protection from photodegradation » (PDF). Aquatic Ecology. 37 (4): 331–39. doi:10.1023/B:AECO.0000007044.52801.6b. S2CID 6930871. Retrieved 26 April 2020.
  204. ^ Melero, Sebastiana; Madejón, Engracia; Ruiz, Juan Carlos; Herencia, Juan Francisco (2007). « Chemical and biochemical properties of a clay soil under dryland agriculture system as affected by organic fertilization ». European Journal of Agronomy. 26 (3): 327–34. doi:10.1016/j.eja.2006.11.004.
  205. ^ Joanisse, Gilles D.; Bradley, Robert L.; Preston, Caroline M.; Bending, Gary D. (2009). « Sequestration of soil nitrogen as tannin–protein complexes may improve the competitive ability of sheep laurel (Kalmia angustifolia) relative to black spruce (Picea mariana) ». New Phytologist. 181 (1): 187–98. doi:10.1111/j.1469-8137.2008.02622.x. PMID 18811620.
  206. ^ Fierer, Noah; Schimel, Joshua P.; Cates, Rex G.; Zou, Jiping (2001). « Influence of balsam poplar tannin fractions on carbon and nitrogen dynamics in Alaskan taiga floodplain soils ». Soil Biology and Biochemistry. 33 (12–13): 1827–39. doi:10.1016/S0038-0717(01)00111-0. Retrieved 3 May 2020.
  207. ^ une b Ponge, Jean-François (2003). « Humus forms in terrestrial ecosystems: a framework to biodiversity » (PDF). Soil Biology and Biochemistry. 35 (7): 935–45. CiteSeerX 10.1.1.467.4937. doi:10.1016/S0038-0717(03)00149-4. Archived from the original on 29 January 2016.
  208. ^ Peng, Xinhua; Horn, Rainer (2007). « Anisotropic shrinkage and swelling of some organic and inorganic soils ». European Journal of Soil Science. 58 (1): 98–107. doi:10.1111/j.1365-2389.2006.00808.x.
  209. ^ Wang, Yang; Amundson, Ronald; Trumbmore, Susan (1996). « Radiocarbon dating of soil organic matter » (PDF). Quaternary Research. 45 (3): 282–88. Bibcode:1996QuRes..45..282W. doi:10.1006/qres.1996.0029. Retrieved 10 May 2020.
  210. ^ Brodowski, Sonja; Amelung, Wulf; Haumaier, Ludwig; Zech, Wolfgang (2007). « Black carbon contribution to stable humus in German arable soils ». Geoderma. 139 (1–2): 220–28. Bibcode:2007Geode.139..220B. doi:10.1016/j.geoderma.2007.02.004. Retrieved 17 May 2020.
  211. ^ Criscuoli, Irene; Alberti, Giorgio; Baronti, Silvia; Favilli, Filippo; Martinez, Cristina; Calzolari, Costanza; Pusceddu, Emanuela; Rumpel, Cornelia; Viola, Roberto; Miglietta, Franco (2014). « Carbon sequestration and fertility after centennial time scale incorporation of charcoal into soil ». PLOS ONE. 9 (3): e91114. Bibcode:2014PLoSO…991114C. doi:10.1371/journal.pone.0091114. PMC 3948733. PMID 24614647.
  212. ^ Wagai, Rota; Mayer, Lawrence M.; Kitayama, Kanehiro; Knicker, Heike (2008). « Climate and parent material controls on organic matter storage in surface soils: a three-pool, density-separation approach ». Geoderma. 147 (1–2): 23–33. Bibcode:2008Geode.147…23W. doi:10.1016/j.geoderma.2008.07.010. hdl:10261/82461. Retrieved 24 May 2020.
  213. ^ Minayeva, Tatiana Y.; Trofimov, Sergey Ya.; Chichagova, Olga A.; Dorofeyeva, E. I.; Sirin, Andrey A.; Glushkov, Igor V.; Mikhailov, N. D.; Kromer, Bernd (2008). « Carbon accumulation in soils of forest and bog ecosystems of southern Valdai in the Holocene ». Biology Bulletin. 35 (5): 524–32. doi:10.1134/S1062359008050142. S2CID 40927739. Retrieved 24 May 2020.
  214. ^ Vitousek, Peter M.; Sanford, Robert L. (1986). « Nutrient cycling in moist tropical forest ». Annual Review of Ecology and Systematics. 17: 137–67. doi:10.1146/annurev.es.17.110186.001033.
  215. ^ Rumpel, Cornelia; Chaplot, Vincent; Planchon, Olivier; Bernadou, J.; Valentin, Christian; Mariotti, André (2006). « Preferential erosion of black carbon on steep slopes with slash and burn agriculture ». Catena. 65 (1): 30–40. doi:10.1016/j.catena.2005.09.005. Retrieved 31 May 2020.
  216. ^ une b Paul, Eldor A.; Paustian, Keith H.; Elliott, E. T.; Cole, C. Vernon (1997). Soil organic matter in temperate agroecosystems: long-term experiments in North America. Boca Raton, Florida: CRC Press. p. 80. ISBN 978-0-8493-2802-2.
  217. ^ « Horizons ». Soils of Canada. Archived from the original on 22 September 2019. Retrieved 7 June 2020.
  218. ^ Frouz, Jan; Prach, Karel; Pizl, Václav; Háněl, Ladislav; Starý, Josef; Tajovský, Karel; Materna, Jan; Balík, Vladimír; Kalčík, Jiří; Řehounková, Klára (2008). « Interactions between soil development, vegetation and soil fauna during spontaneous succession in post mining sites » (PDF). European Journal of Soil Biology. 44 (1): 109–21. doi:10.1016/j.ejsobi.2007.09.002. Retrieved 21 June 2020.
  219. ^ Kabala, Cezary; Zapart, Justyna (2012). « Initial soil development and carbon accumulation on moraines of the rapidly retreating Werenskiold Glacier, SW Spitsbergen, Svalbard archipelago » (PDF). Geoderma. 175–176: 9–20. Bibcode:2012Geode.175….9K. doi:10.1016/j.geoderma.2012.01.025. Retrieved 21 June 2020.
  220. ^ Ugolini, Fiorenzo C.; Dahlgren, Randy A. (2002). « Soil development in volcanic ash » (PDF). Global Environmental Research. 6 (2): 69–81. Retrieved 28 June 2020.
  221. ^ Huggett, Richard J. (1998). « Soil chronosequences, soil development, and soil evolution: a critical review » (PDF). Catena. 32 (3): 155–72. doi:10.1016/S0341-8162(98)00053-8. Retrieved 28 June 2020.
  222. ^ De Alba, Saturnio; Lindstrom, Michael; Schumacher, Thomas E.; Malo, Douglas D. (2004). « Soil landscape evolution due to soil redistribution by tillage: a new conceptual model of soil catena evolution in agricultural landscapes » (PDF). Catena. 58 (1): 77–100. doi:10.1016/j.catena.2003.12.004. Retrieved 28 June 2020.
  223. ^ Phillips, Jonathan D.; Marion, Daniel A. (2004). « Pedological memory in forest soil development » (PDF). Forest Ecology and Management. 188 (1): 363–80. doi:10.1016/j.foreco.2003.08.007. Retrieved 5 July 2020.
  224. ^ Mitchell, Edward A.D.; Van der Knaap, Willem O.; Van Leeuwen, Jacqueline F.N.; Buttler, Alexandre; Warner, Barry G.; Gobat, Jean-Michel (2001). « The palaeoecological history of the Praz-Rodet bog (Swiss Jura) based on pollen, plant macrofossils and testate amoebae(Protozoa) » (PDF). The Holocene. 11 (1): 65–80. Bibcode:2001Holoc..11…65M. doi:10.1191/095968301671777798. S2CID 131032169. Retrieved 5 July 2020.
  225. ^ Carcaillet, Christopher (2001). « Soil particles reworking evidences by AMS 14C dating of charcoal » (PDF). Comptes Rendus de l’Académie des Sciences, Série IIA. 332 (1): 21–28. doi:10.1016/S1251-8050(00)01485-3. Retrieved 14 June 2020.
  226. ^ Retallack, Gregory J. (1991). « Untangling the effects of burial alteration and ancient soil formation ». Annual Review of Earth and Planetary Sciences. 19 (1): 183–206. Bibcode:1991AREPS..19..183R. doi:10.1146/annurev.ea.19.050191.001151.
  227. ^ Bakker, Martha M.; Govers, Gerard; Jones, Robert A.; Rounsevell, Mark D.A. (2007). « The effect of soil erosion on Europe’s crop yields » (PDF). Ecosystems. 10 (7): 1209–19. doi:10.1007/s10021-007-9090-3. Retrieved 19 July 2020.
  228. ^ Uselman, Shauna M.; Qualls, Robert G.; Lilienfein, Juliane (2007). « Contribution of root vs. leaf litter to dissolved organic carbon leaching through soil » (PDF). Soil Science Society of America Journal. 71 (5): 1555–63. Bibcode:2007SSASJ..71.1555U. doi:10.2136/sssaj2006.0386. Retrieved 19 July 2020.
  229. ^ Schulz, Stefanie; Brankatschk, Robert; Dümig, Alexander; Kögel-Knabner, Ingrid; Schloter, Michae; Zeyer, Josef (2013). « The role of microorganisms at different stages of ecosystem development for soil formation » (PDF). Biogeosciences. 10 (6): 3983–96. Bibcode:2013BGeo…10.3983S. doi:10.5194/bg-10-3983-2013. Archived from the original (PDF) on 19 July 2020. Retrieved 19 July 2020.
  230. ^ Gillet, Servane; Ponge, Jean-François (2002). « Humus forms and metal pollution in soil » (PDF). European Journal of Soil Science. 53 (4): 529–39. doi:10.1046/j.1365-2389.2002.00479.x. Retrieved 12 July 2020.
  231. ^ Bardy, Marion; Fritsch, Emmanuel; Derenne, Sylvie; Allard, Thierry; do Nascimento, Nadia Régina; Bueno, Guilherme (2008). « Micromorphology and spectroscopic characteristics of organic matter in waterlogged podzols of the upper Amazon basin ». Geoderma. 145 (3): 222–30. Bibcode:2008Geode.145..222B. CiteSeerX 10.1.1.455.4179. doi:10.1016/j.geoderma.2008.03.008.
  232. ^ Dokuchaev, Vasily Vasilyevich (1967). « Russian Chernozem » (PDF). Jerusalem, Israel: Israel Program for Scientific Translations. Retrieved 26 July 2020.
  233. ^ IUSS Working Group WRB (2015). World Reference Base for Soil Resources 2014: international soil classification system for naming soils and creating legends for soil maps, update 2015 (PDF). Rome, Italy: Food and Agriculture Organization. ISBN 978-92-5-108370-3. Retrieved 26 July 2020.
  234. ^ AlShrouf, Ali (2017). « Hydroponics, aeroponic and aquaponic as compared with conventional farming » (PDF). American Scientific Research Journal for Engineering, Technology, and Sciences. 27 (1): 247–55. Retrieved 2 August 2020.
  235. ^ Leake, Simon; Haege, Elke (2014). Soils for landscape development: selection, specification and validation. Clayton, Victoria, Australia: CSIRO Publishing. ISBN 978-0643109650.
  236. ^ Pan, Xian-Zhang; Zhao, Qi-Guo (2007). « Measurement of urbanization process and the paddy soil loss in Yixing city, China between 1949 and 2000 » (PDF). Catena. 69 (1): 65–73. doi:10.1016/j.catena.2006.04.016. Retrieved 2 August 2020.
  237. ^ Kopittke, Peter M.; Menzies, Neal W.; Wang, Peng; McKenna, Brigid A.; Lombi, Enzo (2019). « Soil and the intensification of agriculture for global food security » (PDF). Environment International. 132: 105078. doi:10.1016/j.envint.2019.105078. ISSN 0160-4120. PMID 31400601. Retrieved 9 August 2020.
  238. ^ Stürck, Julia; Poortinga, Ate; Verburg, Peter H. (2014). « Mapping ecosystem services: the supply and demand of flood regulation services in Europe » (PDF). Ecological Indicators. 38: 198–211. doi:10.1016/j.ecolind.2013.11.010. Retrieved 16 August 2020.
  239. ^ Van Cuyk, Sheila; Siegrist, Robert; Logan, Andrew; Masson, Sarah; Fischer, Elizabeth; Figueroa, Linda (2001). « Hydraulic and purification behaviors and their interactions during wastewater treatment in soil infiltration systems » (PDF). Water Research. 35 (4): 953–64. doi:10.1016/S0043-1354(00)00349-3. PMID 11235891. Retrieved 9 August 2020.
  240. ^ Jeffery, Simon; Gardi, Ciro; Arwyn, Jones (2010). European atlas of soil biodiversity (PDF). Luxembourg, Luxembourg: Publications Office of the European Union. doi:10.2788/94222. ISBN 978-92-79-15806-3. Retrieved 9 August 2020.
  241. ^ De Deyn, Gerlinde B.; Van der Putten, Wim H. (2005). « Linking aboveground and belowground diversity » (PDF). Trends in Ecology and Evolution. 20 (11): 625–33. doi:10.1016/j.tree.2005.08.009. PMID 16701446. Retrieved 16 August 2020.
  242. ^ Hansen, James; Sato, Makiko; Kharecha, Pushker; Beerling, David; Berner, Robert; Masson-Delmotte, Valerie; Pagani, Mark; Raymo, Maureen; Royer, Dana L.; Zachos, James C. (2008). « Target atmospheric CO2: where should humanity aim? » (PDF). Open Atmospheric Science Journal. 2 (1): 217–31. arXiv:0804.1126. Bibcode:2008OASJ….2..217H. doi:10.2174/1874282300802010217. S2CID 14890013. Retrieved 23 August 2020.
  243. ^ Lal, Rattan (11 June 2004). « Soil carbon sequestration impacts on global climate change and food security » (PDF). Science. 304 (5677): 1623–27. Bibcode:2004Sci…304.1623L. doi:10.1126/science.1097396. PMID 15192216. S2CID 8574723. Retrieved 23 August 2020.
  244. ^ Blakeslee, Thomas (24 February 2010). « Greening deserts for carbon credits ». Orlando, Florida, USA: Renewable Energy World. Archived from the original on 1 November 2012. Retrieved 30 August 2020.
  245. ^ Mondini, Claudio; Contin, Marco; Leita, Liviana; De Nobili, Maria (2002). « Response of microbial biomass to air-drying and rewetting in soils and compost » (PDF). Geoderma. 105 (1–2): 111–24. Bibcode:2002Geode.105..111M. doi:10.1016/S0016-7061(01)00095-7. Retrieved 30 August 2020.
  246. ^ « Peatlands and farming ». Stoneleigh, United Kingdom: National Farmers’ Union of England and Wales. 6 July 2020. Retrieved 6 September 2020.
  247. ^ van Winden, Julia F.; Reichart, Gert-Jan; McNamara, Niall P.; Benthien, Albert; Sinninghe Damste, Jaap S. (2012). « Temperature-induced increase in methane release from peat bogs: a mesocosm experiment ». PLoS ONE. 7 (6): e39614. Bibcode:2012PLoSO…739614V. doi:10.1371/journal.pone.0039614. PMC 3387254. PMID 22768100.
  248. ^ Davidson, Eric A.; Janssens, Ivan A. (2006). « Temperature sensitivity of soil carbon decomposition and feedbacks to climate change » (PDF). Nature. 440 (7081): 165–73. Bibcode:2006Natur.440..165D. doi:10.1038/nature04514. PMID 16525463. S2CID 4404915. Retrieved 13 September 2020.
  249. ^ Abrahams, Pter W. (1997). « Geophagy (soil consumption) and iron supplementation in Uganda » (PDF). Tropical Medicine and International Health. 2 (7): 617–23. doi:10.1046/j.1365-3156.1997.d01-348.x. PMID 9270729. S2CID 19647911. Retrieved 20 September 2020.
  250. ^ Setz, Eleonore Zulnara Freire; Enzweiler, Jacinta; Solferini, Vera Nisaka; Amêndola, Monica Pimenta; Berton, Ronaldo Severiano (1999). « Geophagy in the golden-faced saki monkey (Pithecia pithecia chrysocephala) in the Central Amazon » (PDF). Journal of Zoology. 247 (1): 91–103. doi:10.1111/j.1469-7998.1999.tb00196.x. Retrieved 27 September 2020.
  251. ^ Kohne, John Maximilian; Koehne, Sigrid; Simunek, Jirka (2009). « A review of model applications for structured soils: a) Water flow and tracer transport » (PDF). Journal of Contaminant Hydrology. 104 (1–4): 4–35. Bibcode:2009JCHyd.104….4K. CiteSeerX 10.1.1.468.9149. doi:10.1016/j.jconhyd.2008.10.002. PMID 19012994. Archived from the original (PDF) on 7 November 2017. Retrieved 1 November 2017.
  252. ^ Diplock, Elizabeth E.; Mardlin, Dave P.; Killham, Kenneth S.; Paton, Graeme Iain (2009). « Predicting bioremediation of hydrocarbons: laboratory to field scale » (PDF). Environmental Pollution. 157 (6): 1831–40. doi:10.1016/j.envpol.2009.01.022. PMID 19232804. Retrieved 27 September 2020.
  253. ^ Moeckel, Claudia; Nizzetto, Luca; Di Guardo, Antonio; Steinnes, Eiliv; Freppaz, Michele; Filippa, Gianluca; Camporini, Paolo; Benner, Jessica; Jones, Kevin C. (2008). « Persistent organic pollutants in boreal and montane soil profiles: distribution, evidence of processes and implications for global cycling » (PDF). Environmental Science and Technology. 42 (22): 8374–80. Bibcode:2008EnST…42.8374M. doi:10.1021/es801703k. PMID 19068820. Retrieved 27 September 2020.
  254. ^ Rezaei, Khalil; Guest, Bernard; Friedrich, Anke; Fayazi, Farajollah; Nakhaei, Mohamad; Aghda, Seyed Mahmoud Fatemi; Beitollahi, Ali (2009). « Soil and sediment quality and composition as factors in the distribution of damage at the December 26, 2003, Bam area earthquake in SE Iran (M (s)=6.6) » (PDF). Journal of Soils and Sediments. 9: 23–32. doi:10.1007/s11368-008-0046-9. S2CID 129416733. Retrieved 27 September 2020.
  255. ^ Johnson, Dan L.; Ambrose, Stanley H.; Bassett, Thomas J.; Bowen, Merle L.; Crummey, Donald E.; Isaacson, John S.; Johnson, David N.; Lamb, Peter; Saul, Mahir; Winter-Nelson, Alex E. (1997). « Meanings of environmental terms » (PDF). Journal of Environmental Quality. 26 (3): 581–89. doi:10.2134/jeq1997.00472425002600030002x. Retrieved 4 October 2020.
  256. ^ Oldeman, L. Roel (1993). « Global extent of soil degradation ». ISRIC Bi-Annual Report 1991-1992 (PDF). Wageningen, The Netherlands: International Soil Reference and Information Centre(ISRIC). pp. 19–36. Retrieved 4 October 2020.
  257. ^ Sumner, Malcolm E.; Noble, Andrew D. (2003). « Soil acidification: the world story ». In Rengel, Zdenko (ed.). Handbook of soil acidity (PDF). New York, NY, USA: Marcel Dekker. pp. 1–28. Retrieved 11 October 2020.
  258. ^ Karam, Jean; Nicell, James A. (1997). « Potential applications of enzymes in waste treatment ». Journal of Chemical Technology & Biotechnology. 69 (2): 141–53. doi:10.1002/(SICI)1097-4660(199706)69:2<141::AID-JCTB694>3.0.CO;2-U. Retrieved 25 October 2020.
  259. ^ Sheng, Guangyao; Johnston, Cliff T.; Teppen, Brian J.; Boyd, Stephen A. (2001). « Potential contributions of smectite clays and organic matter to pesticide retention in soils ». Journal of Agricultural and Food Chemistry. 49 (6): 2899–2907. doi:10.1021/jf001485d. PMID 11409985. Retrieved 25 October 2020.
  260. ^ Sprague, Lori A.; Herman, Janet S.; Hornberger, George M.; Mills, Aaron L. (2000). « Atrazine adsorption and colloid‐facilitated transport through the unsaturated zone » (PDF). Journal of Environmental Quality. 29 (5): 1632–41. doi:10.2134/jeq2000.00472425002900050034x. Retrieved 18 October 2020.
  261. ^ Ballabio, Cristiano; Panagos, Panos; Lugato, Emanuele; Huang, Jen-How; Orgiazzi, Alberto; Jones, Arwyn; Fernández-Ugalde, Oihane; Borrelli, Pasquale; Montanarella, Luca (15 September 2018). « Copper distribution in European topsoils: an assessment based on LUCAS soil survey ». Science of the Total Environment. 636: 282–98. Bibcode:2018ScTEn.636..282B. doi:10.1016/j.scitotenv.2018.04.268. ISSN 0048-9697. PMID 29709848.
  262. ^ Le Houérou, Henry N. (1996). « Climate change, drought and desertification » (PDF). Journal of Arid Environments. 34 (2): 133–85. Bibcode:1996JArEn..34..133L. doi:10.1006/jare.1996.0099.
  263. ^ Lyu, Yanli; Shi, Peijun; Han, Guoyi; Liu, Lianyou; Guo, Lanlan; Hu, Xia; Zhang, Guoming (2020). « Desertification control practices in China ». Sustainability. 12 (8): 3258. doi:10.3390/su12083258. ISSN 2071-1050.
  264. ^ Kéfi, Sonia; Rietkerk, Max; Alados, Concepción L.; Pueyo, Yolanda; Papanastasis, Vasilios P.; El Aich, Ahmed; de Ruiter, Peter C. (2007). « Spatial vegetation patterns and imminent desertification in Mediterranean arid ecosystems ». Nature. 449 (7159): 213–217. Bibcode:2007Natur.449..213K. doi:10.1038/nature06111. hdl:1874/25682. PMID 17851524. S2CID 4411922.
  265. ^ Wang, Xunming; Yang, Yi; Dong, Zhibao; Zhang, Caixia (2009). « Responses of dune activity and desertification in China to global warming in the twenty-first century ». Global and Planetary Change. 67 (3–4): 167–85. Bibcode:2009GPC….67..167W. doi:10.1016/j.gloplacha.2009.02.004.
  266. ^ Yang, Dawen; Kanae, Shinjiro; Oki, Taikan; Koike, Toshio; Musiake, Katumi (2003). « Global potential soil erosion with reference to land use and climate changes » (PDF). Hydrological Processes. 17 (14): 2913–28. Bibcode:2003HyPr…17.2913Y. doi:10.1002/hyp.1441.
  267. ^ Sheng, Jian-an; Liao, An-zhong (1997). « Erosion control in South China ». Catena. 29 (2): 211–21. doi:10.1016/S0341-8162(96)00057-4. ISSN 0341-8162.
  268. ^ Ran, Lishan; Lu, Xi Xi; Xin, Zhongbao (2014). « Erosion-induced massive organic carbon burial and carbon emission in the Yellow River basin, China » (PDF). Biogeosciences. 11 (4): 945–59. Bibcode:2014BGeo…11..945R. doi:10.5194/bg-11-945-2014.
  269. ^ Verachtert, Els; Van den Eeckhaut, Miet; Poesen, Jean; Deckers, Jozef (2010). « Factors controlling the spatial distribution of soil piping erosion on loess-derived soils: a case study from central Belgium ». Geomorphology. 118 (3): 339–48. Bibcode:2010Geomo.118..339V. doi:10.1016/j.geomorph.2010.02.001.
  270. ^ Jones, Anthony (1976). « Soil piping and stream channel initiation ». Water Resources Research. 7 (3): 602–10. Bibcode:1971WRR…..7..602J. doi:10.1029/WR007i003p00602.
  271. ^ Dooley, Alan (June 2006). « Sandboils 101: Corps has experience dealing with common flood danger ». Engineer Update. US Army Corps of Engineers. Archived from the original on 18 April 2008. Retrieved 14 May 2008.
  272. ^ Oosterbaan, Roland J. (1988). « Effectiveness and social/environmental impacts of irrigation projects: a critical review » (PDF). Annual Reports of the International Institute for Land Reclamation and Improvement (ILRI). Wageningen, The Netherlands. pp. 18–34. Archived (PDF) from the original on 19 February 2009.
  273. ^ Drainage manual: a guide to integrating plant, soil, and water relationships for drainage of irrigated lands (PDF). Washington, D.C.: United States Department of the Interior, Bureau of Reclamation. 1993. ISBN 978-0-16-061623-5.
  274. ^ « Free articles and software on drainage of waterlogged land and soil salinity control ». Archived from the original on 16 August 2010. Retrieved 28 July 2010.
  275. ^ Stuart, Alexander M.; Pame, Anny Ruth P.; Vithoonjit, Duangporn; Viriyangkura, Ladda; Pithuncharurnlap, Julmanee; Meesang, Nisa; Suksiri, Prarthana; Singleton, Grant R.; Lampayan, Rubenito M. (2018). « The application of best management practices increases the profitability and sustainability of rice farming in the central plains of Thailand ». Field Crops Research. 220: 78–87. doi:10.1016/j.fcr.2017.02.005. Retrieved 6 December 2020.
  276. ^ Turkelboom, Francis; Poesen, Jean; Ohler, Ilse; Van Keer, Koen; Ongprasert, Somchai; Vlassak, Karel (1997). « Assessment of tillage erosion rates on steep slopes in northern Thailand ». Catena. 29 (1): 29–44. doi:10.1016/S0341-8162(96)00063-X. Retrieved 13 December 2020.
  277. ^ Saleth, Rathinasamy Maria; Inocencio, Arlene; Noble, Andrew; Ruaysoongnern, Sawaeng (2009). « Economic gains of improving soil fertility and water holding capacity with clay application: the impact of soil remediation research in Northeast Thailand » (PDF). Journal of Development Effectiveness. 1 (3): 336–352. doi:10.1080/19439340903105022. S2CID 18049595. Retrieved 13 December 2020.
  278. ^ Semalulu, Onesmus; Magunda, Matthias; Mubiru, Drake N. (2015). « Amelioration of sandy soils in drought stricken areas through use of Ca-bentonite ». Uganda Journal of Agricultural Sciences. 16 (2): 195–205. doi:10.4314/ujas.v16i2.5. Retrieved 20 December 2020.
  279. ^ International Water Management Institute (2010). « Improving soils and boosting yields in Thailand » (PDF). Success Stories (2). doi:10.5337/2011.0031. Archived (PDF) from the original on 7 June 2012.
  280. ^ Prapagar, Komathy; Indraratne, Srimathie P.; Premanandharajah, Punitha (2012). « Effect of soil amendments on reclamation of saline-sodic soil ». Tropical Agricultural Research. 23 (2): 168–76. doi:10.4038/tar.v23i2.4648. Retrieved 27 December 2020.
  281. ^ Lemieux, Gilles; Germain, Diane (December 2000). « Ramial chipped wood: the clue to a sustainable fertile soil » (PDF). Université Laval, Département des Sciences du Bois et de la Forêt, Québec, Canada. Retrieved 27 December 2020.
  282. ^ Arthur, Emmanuel; Cornelis, Wim; Razzaghi, Fatemeh (2012). « Compost amendment of sandy soil affects soil properties and greenhouse tomato productivity ». Compost Science and Utilization. 20 (4): 215–21. doi:10.1080/1065657X.2012.10737051. S2CID 96896374. Retrieved 3 January 2021.
  283. ^ Glaser, Bruno; Haumaier, Ludwig; Guggenberger, Georg; Zech, Wolfgang (2001). « The ‘Terra Preta’ phenomenon: a model for sustainable agriculture in the humid tropics » (PDF). Naturwissenschaften. 88 (1): 37–41. Bibcode:2001NW…..88…37G. doi:10.1007/s001140000193. PMID 11302125. Retrieved 3 January 2021.
  284. ^ Hillel, Daniel (1992). Out of the Earth: civilization and the life of the soil. Berkeley, California: University of California Press. ISBN 978-0-520-08080-5.
  285. ^ Columella, Lucius Junius Moderatus (1745). Of husbandry, in twelve books, and his book concerning trees, with several illustrations from Pliny, Cato, Varro, Palladius, and other antient and modern authors, translated into English. London, United Kingdom: Andrew Millar. Retrieved 10 January 2021.
  286. ^ Ibn al-‘Awwam (1864). Le livre de l’agriculture, traduit de l’arabe par Jean Jacques Clément-Mullet (in French). Paris, France: Librairie A. Franck. Retrieved 10 January 2021.
  287. ^ Jelinek, Lawrence J. (1982). Harvest empire: a history of California agriculture. San Francisco, California: Boyd and Fraser. ISBN 978-0-87835-131-2.
  288. ^ de Serres, Olivier (1600). Le Théâtre d’Agriculture et mesnage des champs (in French). Paris, France: Jamet Métayer. Retrieved 10 January 2020.
  289. ^ Virto, Iñigo; Imaz, María José; Fernández-Ugalde, Oihane; Gartzia-Bengoetxea, Nahia; Enrique, Alberto; Bescansa, Paloma (2015). « Soil degradation and soil quality in western Europe: current situation and future perspectives ». Sustainability. 7 (1): 313–65. doi:10.3390/su7010313.
  290. ^ Van der Ploeg, Rienk R.; Schweigert, Peter; Bachmann, Joerg (2001). « Use and misuse of nitrogen in agriculture: the German story ». Scientific World Journal. 1 (S2): 737–44. doi:10.1100/tsw.2001.263. PMC 6084271. PMID 12805882.
  291. ^ « Van Helmont’s experiments on plant growth ». BBC World Service. Retrieved 17 January 2021.
  292. ^ une b c Brady, Nyle C. (1984). The nature and properties of soils (9th ed.). New York, New York: Collier Macmillan. ISBN 978-0-02-313340-4. Retrieved 17 January 2021.
  293. ^ de Lavoisier, Antoine-Laurent (1777). « Mémoire sur la combustion en général » (PDF). Mémoires de l’Académie Royale des Sciences (in French). Retrieved 17 January 2021.
  294. ^ Boussingault, Jean-Baptiste (1860–1874). Agronomie, chimie agricole et physiologie, volumes 1–5 (in French). Paris, France: Mallet-Bachelier. Retrieved 17 January 2021.
  295. ^ von Liebig, Justus (1840). Organic chemistry in its applications to agriculture and physiology. London: Taylor and Walton. Retrieved 17 January 2021.
  296. ^ Way, J. Thomas (1849). « On the composition and money value of the different varieties of guano ». Journal of the Royal Agricultural Society of England. 10: 196–230. Retrieved 17 January 2021.
  297. ^ Tandon, Hari L.S. « A short history of fertilisers ». Fertiliser Development and Consultation Organisation. Archived from the original on 23 January 2017. Retrieved 17 December 2017.
  298. ^ Way, J. Thomas (1852). « On the power of soils to absorb manure ». Journal of the Royal Agricultural Society of England. 13: 123–43. Retrieved 17 January 2021.
  299. ^ Warington, Robert (1878). Note on the appearance of nitrous acid during the evaporation of water: a report of experiments made in the Rothamsted laboratory. London, United Kingdom: Harrison and Sons.
  300. ^ Winogradsky, Sergei (1890). « Sur les organismes de la nitrification ». Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences (in French). 110 (1): 1013–16. Retrieved 17 January 2021.
  301. ^ Hilgard, Eugene W. (1921). Soils: their formation, properties, composition, and relations to climate and plant growth in the humid and arid regions. London, United Kingdom: The Macmillan Company. Retrieved 17 January 2021.
  302. ^ Fallou, Friedrich Albert (1857). Anfangsgründe der Bodenkunde (PDF) (in German). Dresden, Germany: G. Schönfeld’s Buchhandlung. Archived from the original (PDF) on 15 December 2018. Retrieved 15 December 2018.
  303. ^ Glinka, Konstantin Dmitrievich (1914). Die Typen der Bodenbildung: ihre Klassifikation und geographische Verbreitung (in German). Berlin, Germany: Borntraeger.
  304. ^ Glinka, Konstantin Dmitrievich (1927). The great soil groups of the world and their development. Ann Arbor, Michigan: Edwards Brothers. Retrieved 17 January 2021.

Bibliographie[[[[edit]

Further reading[[[[edit]

  • Soil-Net.com A free schools-age educational site teaching about soil and its importance.
  • Adams, J.A. 1986. Dirt. College Station, Texas: Texas A&M University Press ISBN 0-89096-301-0
  • Certini, G., Scalenghe, R. 2006. Soils: Basic concepts and future challenges. Cambridge Univ Press, Cambridge.
  • David R. Montgomery, Dirt: The Erosion of Civilizations, ISBN 978-0-520-25806-8
  • Faulkner, Edward H. 1943. Plowman’s Folly. New York, Grosset & Dunlap. ISBN 0-933280-51-3
  • LandIS Free Soilscapes Viewer Free interactive viewer for the Soils of England and Wales
  • Jenny, Hans. 1941. Factors of Soil Formation: A System of Quantitative Pedology
  • Logan, W.B. 1995. Dirt: The ecstatic skin of the earth. ISBN 1-57322-004-3
  • Mann, Charles C. September 2008.  » Our good earth » National Geographic Magazine
  • « 97 Flood ». USGS. Archived from the original on 24 June 2008. Retrieved 8 July 2008. Photographs of sand boils.
  • Soil Survey Division Staff. 1999. Soil survey manual. Soil Conservation Service. U.S. Department of Agriculture Handbook 18.
  • Soil Survey Staff. 1975. Soil Taxonomy: A basic system of soil classification for making and interpreting soil surveys. USDA-SCS Agric. Handb. 436. United States Government Printing Office, Washington, DC.
  • Soils (Matching suitable forage species to soil type), Oregon State University
  • Gardiner, Duane T. « Lecture 1 Chapter 1 Why Study Soils? ». ENV320: Soil Science Lecture Notes. Texas A&M University-Kingsville. Archived from the original on 9 February 2018. Retrieved 7 January 2019.
  • Janick, Jules. 2002. Soil notes, Purdue University
  • LandIS Soils Data for England and Wales a pay source for GIS data on the soils of England and Wales and soils data source; they charge a handling fee to researchers.

External links[[[[edit]

480 ms 14.3%
? 460 ms 13.7%
Scribunto_LuaSandboxCallback::gsub 420 ms 12.5%
Scribunto_LuaSandboxCallback::getExpandedArgument 200 ms 6.0%
Scribunto_LuaSandboxCallback::callParserFunction 180 ms 5.4%
Scribunto_LuaSandboxCallback::match 140 ms 4.2%
Scribunto_LuaSandboxCallback::getAllExpandedArguments 120 ms 3.6%
recursiveClone 120 ms 3.6%
Scribunto_LuaSandboxCallback::plain 100 ms 3.0%
80 ms 2.4%
[others] 1060 ms 31.5%
Number of Wikibase entities loaded: 1/400
->


Source de l’article

A découvrir