Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-26T15:44:53.919Z Has data issue: false hasContentIssue false

A Review of the Reactivity of Phosphatase Controlled by Clays and Clay Minerals: Implications for Understanding Phosphorus Mineralization in Soils

Published online by Cambridge University Press:  01 January 2024

Ai Chen
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
Yuji Arai*
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
Rights & Permissions [Opens in a new window]

Abstract

Mineralizable macronutrients (e.g. C, N, P, and S) are sorbed readily (i.e. adsorption and precipitation) in clays and clay minerals. Phosphorus (P) is one of the limiting macronutrients in soils because both phosphate and organic P undergo chemisorption in soil minerals. Furthermore, phosphatases that mineralize the organic P species tend to partition into soil minerals, suppressing the interactions between organic P and phosphatase. Adsorbed phosphatase on the mineral surfaces can regulate the enzyme activity and influence the biochemical properties of the enzyme (e.g. kinetics, conformation, and stability), affecting the P cycle in the terrestrial environment. Phosphatase–mineral interactions are widely reported to decrease the enzyme activity while enhancing the enzyme stability (e.g. thermal and proteolysis stability). Contradictory findings have also been reported. Specific enzymes, mineral characteristics, and reaction conditions are probably responsible for various reactivity (e.g. mineralization). The purpose of the present review was to summarize current and past investigations of acid and alkaline phosphatase sorption in clays and clay minerals and to examine phosphatase chemical properties (e.g. kinetic activity, thermal and proteolysis stability) and factors (e.g. pH, saturating cations of the mineral, enzyme structure, and mineral surface polarity) influencing the phosphatase-mineral interaction. Lastly, also reviewed is the application of phosphatase–mineral interactions with some expansion to other enzymes as an indication of potential future application for phosphatase and future research needs.

Type
Review
Copyright
Copyright © The Author(s), under exclusive licence to The Clay Minerals Society 2023

Introduction

Globally, around half of the topsoils contain 15–30% clay and the clay content in ~ 20% of the topsoils is > 30% (Wei et al., Reference Wei, Dai, Duan, Liu and Yuan2014). Depending on factors such as parent material and weathering, the dominant clays and clay minerals vary in different soils. For example, in weathered soils, Al and Fe (oxyhydr)oxides are the dominant clay minerals (Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011). In young soils developed from volcanic parent materials, allophanes and ferrihydrite are found commonly (Allison, Reference Allison2006). Clays from the illite and mica groups are commonly found in arid and high-latitude regions, while humid tropical soils contain predominantly 1:1 kaolinite groups (Ito & Wagai, Reference Ito and Wagai2017). Fibrous silicates such as palygorskite and sepiolite are found in calcareous and gypsiferous soils in arid regions (Shirvani et al., Reference Shirvani, Khalili, Kalbasi, Shariatmadari and Nourbakhsh2020). Smectite-group minerals are ubiquitous in many soils, especially in the semi-arid and arid regions, in Vertisols, and in continental and oceanic sediments (Ito & Wagai, Reference Ito and Wagai2017; Odom, Reference Odom1984).

The extracellular enzymes are common in soils. They are excreted by soil microbes such as fungi and actinomycetes, and by plants. The extracellular enzymes carry out many critical functions including mineralizing soil nutrients. They are also involved in the formation and decomposition of soil organic matter, the bioremediation of polluted soils, and the nutrient cycling processes such as mineralization and immobilization (Bollag et al., Reference Bollag, Mertz, Otjen, Anderson and Coats1994; Burns et al., Reference Burns, DeForest, Marxsen, Sinsabaugh, Stromberger, Wallenstein, Weintraub and Zoppini2013; Naidja et al., Reference Naidja, Huang and Bollag2000; Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011). Furthermore, the activity of some enzymes can be used as a soil quality indicator to assess soil health, soil productivity, and ecological functionality (Alkorta et al., Reference Alkorta, Aizpurua, Riga, Albizu, Amézaga and Garbisu2003; Gil-Sotres et al., Reference Gil-Sotres, Trasar-Cepeda, Leirós and Seoane2005).

The secreted, extracellular enzymes in soils are known to be susceptible to adsorption and immobilization by soil minerals. The interaction between soil enzymes and minerals has been discussed for decades, covering a wide range of studies from the investigation of pure layered silicate–enzyme reactions to models mimicking the realistic soil mineral colloid–enzyme interaction (Naidja et al., Reference Naidja, Huang and Bollag2000). Datta et al. (Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017) also summarized some key results on the adsorption process (e.g. driving forces, adsorption isotherm) of enzymes on minerals and on adsorption kinetics. The general effects of some soil parameters on enzyme adsorption and some key factors that affect enzyme activity were also reviewed in the same work.

Phosphatases, as a group of important enzymes in soils, along with their properties and activities, have been studied extensively but only brief discussion about the interaction between phosphatases and clay minerals has been included in reviews with a general topic of enzyme–clay mineral interaction (Nannipieri et al., Reference Nannipieri, Giagnoni, Landi, Renella, Bünemann, Oberson and Frossard2011). More comprehensive reviews focusing on phosphatase behavior and properties during adsorption are still lacking. The objective of the present review, therefore, was to focus specifically on the adsorption of phosphatases on clay minerals and how it alters several important properties of phosphatases.

In recognition of the potential application of immobilized (e.g. adsorbed) enzymes, many reviews discuss potential industrial applications. For instance, Jesionowski et al. (Reference Jesionowski, Zdarta and Krajewska2014) focused on selecting an appropriate carrier as the adsorbent for an enzyme and compared the advantages of different kinds of commonly used carriers (e.g. silicas, metals, clay minerals, and natural and synthetic polymers). Many recent reviews discussed different industrial applications of commonly used clay mineral-immobilized enzymes (e.g. An et al., Reference An, Zhou, Zhuang, Tong and Yu2015; Basso & Serban, Reference Basso and Serban2019; Maghraby et al., Reference Maghraby, El-Shabasy, Ibrahim and Azzazy2023). Application of the interaction between phosphatase and minerals is also discussed in the present review, therefore. As current application of such interactions is still limited, this review also expands the discussion into enzymes that can mineralize other macronutrients as indications of future potential application of phosphatases.

Mineralizable Macronutrient Retention in Soils by Clays and Clay Minerals

Soil minerals play an important role in numerous soil reactions and one of the most important roles is the retention of mineralizable macronutrients (C, N, P, and S). With small particle size and, therefore, large specific surface area, the minerals carry abundant positive and negative charges on their internal and external surfaces, making them effective in interacting with different forms of mineralizable macronutrients, including inorganic anions (e.g. phosphate, nitrate, and sulfate), cations (e.g. NH4 +), and organic compounds (e.g. enzymes, organic acids).

The distribution of charges on the variable-charge mineral surface is not homogeneous, and the two sources of charges are permanent charge, which is an inherent characteristic of the mineral, and pH-dependent charge. For some clay minerals, such as allophane, kaolinite, and Al/Fe (oxyhydr)oxides, the charges are predominantly pH-dependent while a large percentage of permanent charge is often observed in chlorite and 2:1 clays such as smectite and vermiculite (McBride, Reference McBride, Dixon and Weed1989; Sumner, Reference Sumner2000).

Dissolved ions attracted and retained by clays can exchange with free ions in the surrounding soil solution and the cation- and anion-exchange capacities of a specific clay mineral are pH-dependent. pH-dependent charges arise through the adsorption of ions and protonation and deprotonation of the exposed hydroxyl groups on the edges of the mineral surfaces where the mineral unit layer is disrupted (Cross & Yariv, Reference Cross, Yariv, Cross and Yariv1979).

The complexation between ions and the charged mineral surfaces can be divided into two mechanisms: inner-sphere and outer-sphere complexation. In the case of outer-sphere complexation, water molecules are involved and form a bridge between the adsorbed ion and the clay-mineral surface. Outer-sphere complexation is, therefore, rather weak, rapid, and reversible, as it mainly involves electrostatic interactions and occurs only between surfaces and ions that bear the opposite charges (Sparks, Reference Sparks and Sparks2003).

In contrast, the adsorbed ion is bonded directly to the mineral surface in inner-sphere complexation, which provides a relatively strong bonding environment and is often irreversible compared to outer-sphere complexation (Sparks, Reference Sparks and Sparks2003). Depending on the characteristics of the adsorbate and adsorbent, inner-sphere adsorbed ions are less susceptible to being exchanged easily by ions other than the outer-sphere adsorbed ones (Sposito, Reference Sposito1984).

Through ion-exchange reactions, the minerals can store and supply the adsorbed ions, largely affecting the fate and bioavailability of nutrients in soils. For example, a large amount of nitrate has been found to be retained in Oxisols and Ultisols (Kome et al., Reference Kome, Enang, Tabi and Yerima2019; Lehmann et al., Reference Lehmann, Lehmann, Lilienfein, Rebel, do Carmo Lima and Wilcke2004). Ammonium is commonly fixed in the interlayers of 2:1 phyllosilicates (e.g. vermiculite, montmorillonite) (Mamo et al., Reference Mamo, Taylor and Shuford1993). Clays such as Fe and Al (oxyhydr)oxides (e.g. ferrihydrite, gibbsite), and clay minerals such as allophane and imogolite are effective at retaining phosphate and organic phosphorus compounds through adsorption and precipitation (Redel et al., Reference Redel, Cartes, Velásquez, Poblete-Grant, Poblete-Grant, Bol and Mora2016; Sims & Sharpley, Reference Sims and Sharpley2005).

The retention of mineralizable macronutrients may also be a result of the interaction between the minerals and enzymes involved in the mineralization. The enzyme–clay mineral interaction can regulate enzyme activities and alter their kinetic properties (Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017). As the free enzymes are probably prone to rapid denaturation, the bound forms may serve as a reservoir of potential enzyme activity (Burns, Reference Burns1982; Rejsek et al., Reference Rejsek, Vranova, Pavelka and Formanek2012; Shindo et al., Reference Shindo, Watanabe, Onaga, Urakawa, Nakahara and Huang2002). In some cases, the association between enzymes and minerals stabilized the structure of the enzymes, and the lifetime of the enzyme was prolonged while high activity was maintained (Bollag et al., Reference Bollag, Mertz, Otjen, Anderson and Coats1994; Burns et al., Reference Burns, DeForest, Marxsen, Sinsabaugh, Stromberger, Wallenstein, Weintraub and Zoppini2013). The stabilization effect can also protect the enzymes against proteolysis and abiotic processes such as thermal denaturation and dehydration (Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011). Decreased enzyme activity upon enzyme adsorption, however, has also been documented widely when the adsorption made the active site of the enzyme less accessible or when enzyme deformation was induced (Dick & Tabatabai, Reference Dick and Tabatabai1987; Hughes & Simpson, Reference Hughes and Simpson1978; Makboul & Ottow, Reference Makboul and Ottow1979; Tietjen & Wetzel, Reference Tietjen and Wetzel2003; Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011). Other properties related to the enzyme activity (e.g. optimal temperature, pH, or substrate affinity) may be largely influenced by adsorption to minerals, which would indirectly affect the mineralizable macronutrient cycles, such as enzymatic degradation of organic C, N, P, and S (Rao et al., Reference Rao, Violante and Gianfreda2000).

Importance of Organic P Mineralization

Of all the macronutrients, P is acknowledged to be rather immobile and unavailable due to precipitation with various metals and the partitioning reaction (e.g. adsorption) with minerals and organic matter (Abel et al., Reference Abel, Ticconi and Delatorre2002; Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005; Vance et al., Reference Vance, Uhde-Stone and Allan2003). Adsorbed phosphates are difficult to exchange due to transformation into more recalcitrant precipitates. Furthermore, the clay-sized colloidal phosphate has also been reported to have enhanced mobility and is prone to loss via surface runoff and subsurface routes such as preferential flow, leading to nutrient loss (Chen & Arai, Reference Chen, Arai and Sparks2020). Phosphorus deficiency, therefore, has been observed in many soils, such as tropical forests and weathered soils, and biomass accumulation and productivity can be severely constrained in P-deficient soils (Oliverio et al., Reference Oliverio, Bissett, McGuire, Saltonstall, Turner and Fierer2020). Furthermore, due to the strong fixation of phosphate in soils, the inefficiency of phosphate fertilizer has been a problem, so an excessive amount is usually applied (Vance et al., Reference Vance, Uhde-Stone and Allan2003; Zhang et al., Reference Zhang, Tang, Feng, Wang, Li, Shen and Zhang2019). A consequence of excessive P fertilization is the loss of P to waters, resulting in eutrophication. By estimation, the inexpensive yet finite and non-renewable source of the widely used rock phosphate fertilizer is likely to be depleted within decades, especially given the increasing requirement for P fertilization and the political instability of some areas where rock phosphate is found (Abelson, Reference Abelson1999; Garske & Ekardt, Reference Garske and Ekardt2021; Richardson et al., Reference Richardson, Lynch, Ryan, Delhaize, Smith, Smith, Harvey, Ryan, Veneklaas, Lambers, Oberson, Culvenor and Simpson2011). As a result, the price of phosphate fertilizer has increased to >$700/ton (with an increase of $100–400/ton since 2009) recently in 2021 following the sharp price increase in 2008 (Beghin & Nogueira, Reference Beghin and Nogueira2021; Cordell & White, Reference Cordell and White2014).

Soils can provide phosphate through weathering of mineral phosphate and mineralization of organic P compounds (Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005). Organic P accounts for 20–80% of total P in soils and it can represent a particularly important potential pool to provide phosphate, especially in soils with no fertilizer input and/or poor phosphate solubility (Tiessen et al., Reference Tiessen, Stewart and Cole1984). Other strategies such as promoted mineralization of the organic P, therefore, could be another option to increase the P availability in soils. Phosphohydrolases are a group of enzymes that can catalyze the hydrolysis of organic P and some inorganic P such as linear polyphosphate and pyrophosphate (Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005). Another group of enzymes that can release P from organic P compounds in soils is lyases involved in the P cycle (Rodríguez et al., Reference Rodríguez, Fraga, Gonzalez and Bashan2006).

Classification of Soil Phosphohydrolases and Phosphate-Releasing Lyases

Phosphohydrolases can catalyze the hydrolysis of different bonds and they can be classified depending on the bond type according to the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology (NC-IUBMB) (Fig. 1). The Enzyme Commission (EC) number is used as a numerical classification to categorize enzymes based on the reactions catalyzed. For example, many phosphohydrolases target the ester bonds. Phosphoric monoester hydrolases (EC 3.1.3) are enzymes that can cleave the ester bonds on a phosphomonoester. Of all phosphoric monoester hydrolases, phosphatases have been studied widely, and depending on their optimal pH (below or above pH 7.0), phosphatases can be categorized into acid phosphatases (EC 3.1.3.2) and alkaline phosphatases (EC 3.1.3.1) (Vincent et al., Reference Vincent, Crowder and Averill1992). Depending on substrate specificity, phosphatases can also be grouped into, for example, protein serine/threonine phosphatase (PSTP or PSP), protein tyrosine phosphatase (PTP), histidine phosphatase, lipid phosphatase, sugar phosphatases, and glycerophosphatases (Alef & Nannipieri, Reference Alef, Nannipieri, Alef and Nannipieri1995; Fahs et al., Reference Fahs, Lujan and Kohn2016). Other examples of phosphoric monoester hydrolases include phytases and nucleotidases.

Fig. 1 Classification of enzymes that can release phosphate from organic P compounds, based on the enzyme database from the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology (NC-IUBMB). Alkaline (EC 3.1.3.1) and acid (EC 3.1.3.2) phosphatases are the main focus of the subsequent sections

Other ester bond-targeting phosphohydrolases consist of phosphoric diester hydrolases (EC 3.1.4) such as phospholipases, triphosphoric monoester hydrolases (EC 3.1.5), diphosphoric monoester hydrolases (EC 3.1.7), and phosphoric triester hydrolases (EC 3.1.8).

Phosphohydrolases also include enzymes acting on guanosine triphosphate (GTP) (EC 3.6.5), enzymes targeting phosphoryl-containing anhydrides (EC 3.6.1) such as pyrophosphotases (or diphosphate phosphohydrolase) (EC 3.6.1.1) which catalyze the hydrolysis of pyrophosphate, phosphoamidases which act on the P–N bonds (EC 3.9.1), and phosphonatases (or phosphonate hydrolases) (EC 3.11) (e.g. phosphonoacetaldehyde hydrolase, phosphonopyruvate hydrolase, and phosphonoacetate hydrolase) which target the C–P bond and hydrolyze C-phosphono groups (Alef & Nannipieri, Reference Alef, Nannipieri, Alef and Nannipieri1995; Chin et al., Reference Chin, McGrath and Quinn2016; Nannipieri et al., Reference Nannipieri, Giagnoni, Landi, Renella, Bünemann, Oberson and Frossard2011; White & Metcalf, Reference White and Metcalf2007).

Lyases are enzymes that can catalyze the cleavage of bonds by means other than hydrolysis or oxidation and they can also be classified according to the bonds they attack. For example, a group of P-O lyases (EC 4.6), the ribonucleases, are usually secreted under phosphate starvation conditions, and they can catalyze the cleavage of P–O bonds in RNAs (Abel et al., Reference Abel, Ticconi and Delatorre2002; Bariola et al., Reference Bariola, Howard, Crispin, Verburg, Jaglan and Green1994). C-P lyases (EC 4.7) have also been found in phosphate-limited environments and they target the C–P bonds in organophosphonates (Oliverio et al., Reference Oliverio, Bissett, McGuire, Saltonstall, Turner and Fierer2020; Tapia-Torres et al., Reference Tapia-Torres, Rodríguez-torres, Elser, Islas, Souza, García-Oliva and Olmedo-Álvarez2016).

Reaction Mechanisms of Phosphohydrolases and Phosphate-Releasing Lyases

The majority of P-containing organic compounds contain the ester bond, and the activity of the phosphomonoesterases is usually greater than that of phosphodi- or phosphotri-esterases in soils (Alef & Nannipieri, Reference Alef, Nannipieri, Alef and Nannipieri1995). Among all phosphomonoesterases in soils, alkaline and acid phosphatases are those studied most widely (Nannipieri et al., Reference Nannipieri, Giagnoni, Landi, Renella, Bünemann, Oberson and Frossard2011).

Alkaline phosphatases have a common conserved core structure containing a serine residue that binds with phosphate on the phosphomonoester during the catalysis, resulting in the formation of a phosphoserine intermediate. The fully active site for each alkaline phosphatase monomer also contains two Zn(II) which are bridged by the phosphate group, and one Mg(II) coordinated with a water molecule that is hydrogen-bonded to the phosphate (Coleman, Reference Coleman1992; Kim & Wyckoff, Reference Kim and Wyckoff1991; Stec et al., Reference Stec, Holtz and Kantrowitz2000). Unlike alkaline phosphatases, a histidine residue acts as the nucleophilic acceptor in the catalytic process of acid phosphatases, and phosphohistidine intermediates are formed. Acid phosphatases also have conserved arginine residues and many of them carry a characteristic RHGXRXP (R: arginine; H: histidine; G:glycine; P: proline; X: any amino acid) motif (some acid phosphatases such as Aspergillus niger acid phosphatase only have the RHG motif) (Ostanin et al., Reference Ostanin, Harms, Stevis, Kuciel, Zhou and Van Ettene1992; Ullah et al., Reference Ullah, Cummins and Dischinger1991). For both alkaline and acid phosphatase, the final hydrolysis products are phosphate and corresponding hydrocarbons such as alcohol or phenol (Fig. 2). Phosphodiesters are converted to corresponding phosphomonoesters through phosphodiesterases, which then go through subsequent hydrolysis via phosphatases and release phosphate (Blake et al., Reference Blake, O’Neil and Surkov2005). Phosphotriesters are hydrolyzed to phosphodiesters which then experience subsequent hydrolysis while releasing phosphate. myo-Inositol hexakisphosphate, or phytic acid, is the predominant source organic P in many soils, accounting for more than half of the organic P pool (Anderson, Reference Anderson, Khasawneh chairman, Sample and Kamprath1980; Liu et al., Reference Liu, Cai, Chen, Mo, Ding, Liang, Liu and Tian2018). The enzymes that hydrolyze phytic acid are referred to as phytases and, depending on catalytic mechanisms, they are generally classified into four categories. The histidine acid phytases (HAPhys) share a similar catalysis mechanism, such as the conserved RHGXRXP motif and the formation of phosphohistidine intermediate (Oh et al., Reference Oh, Choi, Park, Kim and Oh2004; Ullah et al., Reference Ullah, Cummins and Dischinger1991). β-Propeller phytases (BPPhys) have an optimum alkaline pH, and metal ions, especially Ca(II), are found in the active site and contribute to the catalytic activity of BPPhy (Kim et al., Reference Kim, Kim, Bae, Yu and Oh1998). Shin et al. (Reference Shin, Ha, Oh, Oh and Oh2001) classified the active sites on the enzyme and categorized them into “cleavage site” and “affinity site.” Two adjacent phosphate groups on a phytic acid molecule bind simultaneously with Ca(II) in the cleavage site and affinity site, respectively. The phosphate bound to the cleavage site would later be removed, while the other one increases the affinity of enzymes for the remaining myo-inositol polyphosphate intermediate. Cysteine phytases (Cphys) are members of the cysteine phosphatases (CP). The Cphys share some substantial properties with the protein tyrosine phosphatases, another group of enzymes from the CP family. For example, they have a conserved nucleophilic cysteine residue and the catalytic hydrolysis involves the formation of a cysteinyl phosphate intermediate. Furthermore, a characteristic active-site motif of these enzymes is HCXXGXXR(T/S) (C: cysteine; G: glycine; S: serine; T: threonine). The formation of a “cysteinyl-phosphate trigonal–bipyramidal” pentavalent intermediate has been postulated and sequential hydrolysis is facilitated with the reorientation of the intermediate (Chu et al., Reference Chu, Guo, Lin, Chou, Shr, Lai, Tang, Cheng, Selinger and Wang2004; Turner et al., Reference Turner, Richardson and Mullaney2007; Yanke et al., Reference Yanke, Selinger and Cheng1999). Purple acid phytases (PAPhys) are named for to their distinct purple color and they have binuclear metal bridging at the active sites, and most PAPhys are coordinated with five conserved consensus motifs (Dionisio et al., Reference Dionisio, Madsen, Holm, Welinder, Jorgensen, Stoger, Arcalis and Brinch-pedersen2011; Rodríguez, Reference Rodríguez2018). The metal bridge is usually in the form of Fe(III)-divalent metal where Fe(III) is on the chromophoric site and the divalent metal can be Fe(II), Zn(II), or Mn(II) (Faba-Rodriguez et al., Reference Faba-Rodriguez, Gu, Salmon, Dionisio, Brinch-Pedersen, Brearley and Hemmings2022; Hegeman & Grabau, Reference Hegeman and Grabau2001; Nasrabadi et al., Reference Nasrabadi, Greiner, Yamchi and Nourzadeh Roshan2018). The scissile phosphate group binds with the metal ions on the active site through hydrogen bonds. The substrate is then rearranged so that the phosphate group is coordinated directly with metal ions, forming a catalytic complex that facilitates the nucleophilic attack, initiating the hydrolysis (Rodríguez, Reference Rodríguez2018). Depending on the species of phytases, hydrolysis products are phosphate and different myo-inositol polyphosphates.

Fig. 2 Phosphomonoester hydrolysis reaction catalyzed by alkaline and acid phosphatases

As well as the compounds containing P in the highest oxidation state (+ 5) such as phosphates and phosphate esters, reduced P forms are also found in soils, such as the organophosphonate, glyphosate, which involves a directly connected C–P bond (Kehler et al., Reference Kehler, Haygarth, Tamburini and Blackwell2021; Ternan et al., Reference Ternan, Mc Grath, Mc Mullan and Quinn1998).

Organophosphonates can be catabolized by phosphonatases or C-P lyases. The characteristic property of the substrate of the phosphonatases is the presence of an electron-withdrawing β-carbonyl group which facilitates the heterolytic cleavage of the C–P bond (Kamat & Raushel, Reference Kamat and Raushel2013). The C-P lyases have a redox-active [4Fe-4S]-cluster and the catalysis process involves a radical-based homolytic cleavage of the C–P bond. The products of the C-P lyase degradation are inorganic phosphate and corresponding hydrocarbon, e.g. alkanes, alkenes, and benzenes from alkyl-, vinyl-, or phenyl-phosphonate degradation (Daughton et al., Reference Daughton, Cook and Alexander1979; Stosiek et al., Reference Stosiek, Talma and Klimek-ochab2020; Wackett et al., Reference Wackett, Shames, Venditti and Walsh1987). More detailed mechanisms of phosphonatases and C-P lyases have been reviewed in the literature (Kamat & Raushel, Reference Kamat and Raushel2013; Stosiek et al., Reference Stosiek, Talma and Klimek-ochab2020).

Enzyme Activity and other Properties Affected by the Formation of the Enzyme–Clay Mineral Complex

Enzyme Adsorption Mechanisms in Clays and Clay Minerals

Protein enzymes bear a variety of functional groups which can interact with the clay mineral surface and the adsorption of enzymes onto the mineral surface can be the result of various forces and interactions including salt linkage, ligand exchange, hydrogen bonding, hydrophobic interaction, conformational entropy, Coulombic force (or electrostatic interaction), and van der Waals forces (Boyd & Mortland, Reference Boyd, Mortland, Bollag and Stotzky1990; Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017; Roth & Lenhoff, Reference Roth and Lenhoff1995; Theng, Reference Theng2012).

Among all of the forces mentioned, electrostatic attraction, as a long-range force, has been proposed by some authors to be the one factor responsible for the first contact between the enzyme and clay (Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017). Positively charged amino acids (e.g. arginine, histidine, and lysine above pH 4) can be attracted electrostatically to a negatively charged mineral surface. Furthermore, ligand exchange between the carboxylic groups of the enzyme and hydroxyl groups of the mineral has been proposed as the mechanism for enzyme adsorption on metal oxide surfaces (Sepelyak et al., Reference Sepelyak, Feldkamp, Moody, White and Hem1984).

Hydrophobic interaction also contributes significantly to the enzyme–mineral reaction. Phyllosilicates such as montmorillonite have hydrophobic siloxane layers and hydrophilic exchangeable ions on the mineral surfaces (Staunton & Quiquampoix, Reference Staunton and Quiquampoix1994). Some amino acids have a hydrophobic side chain (e.g. tryptophan, tyrosine), which would probably engage in hydrophobic interaction with the siloxane layer after the enzyme exchanges with the surface exchangeable ions, and the dehydration from such hydrophobic interaction facilitates the adsorption of most enzymes (Norde, Reference Norde2008). For a hydrophobic mineral surface, even when both the enzyme and surface have charges of the same sign, hydrophobic force may, therefore, overcome the electrostatic repulsion and bring the surface and the enzyme together (Norde, Reference Norde2008; Quiquampoix et al., Reference Quiquampoix, Abadie, Baron, Leprince, Matumoto-Pintro, Ratcliffe, Staunton, Horbett and Brash1995). The adsorption process may be further augmented by enzyme conformational change (Fig. 3), van der Waals forces, and hydrogen bonding.

Fig. 3 a The enzyme is in its active conformation and the substrate enters the active site. b Substrate binds to the active site, forming an enzyme–substrate complex. c Conformational change of the enzyme alters the shape of the active site, preventing substrate binding

The adsorption of the enzyme on the mineral surface can induce a conformational change in the enzyme, leading to changes in the enzyme’s conformational entropy and Gibbs free energy of the system. For example, a decrease in the ordered secondary structure of the enzyme due to adsorption would lead to increased conformational entropy, thus reducing the Gibbs free energy of the system. Subsequently, the modification of enzyme structure and adsorption becomes more irreversible (Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005).

Van der Waals interaction consists of three components, permanent dipole-permanent dipole (Keesom force), permanent dipole-induced dipole (Debye force), and induced dipole-induced dipole (London force). Van der Waals forces increase as the distance between the enzyme and mineral surface decreases and as the molecular size increases (Roth & Lenhoff, Reference Roth and Lenhoff1995; Stotzky, Reference Stotzky, Huang and Schnitzer1986). Although the attractive force is rather weak, it can work on all molecules despite charges and can be strengthened through the formation of a salt bridge between enzyme-bound salt ions and the mineral surface or between clay-complexed cations and the enzyme, providing a polarizable layer, or through adjacent van der Waals forces acting together, creating a stronger interaction (Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005; Roth & Lenhoff, Reference Roth and Lenhoff1995; Stotzky, Reference Stotzky, Huang and Schnitzer1986).

Hydrogen bonds can be formed between OH groups or O on the mineral surface, or more importantly, clay-associated water, and amino acid functional groups (e.g. carbonyl, amine) (Stotzky, Reference Stotzky, Huang and Schnitzer1986). Like the van der Waals forces, hydrogen bonding becomes a strong binding force when numerous hydrogen bonds are functioning together (Stotzky, Reference Stotzky, Huang and Schnitzer1986). Hydrogen bonds are generally formed between the enzyme and a polar surface while a nonpolar surface doesn’t interact with the enzyme via hydrogen bonds (Norde, Reference Norde2008).

Factors Affecting the Adsorption of Phosphatase on the Mineral Surface

The adsorption of enzymes on a mineral surface is affected by many factors in various aspects. The general effects of temperature, pH, soil moisture, and dissolved ion concentration on enzyme adsorption have been reviewed by Datta et al. (Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017). The interaction between clay minerals and phosphatases is the most widely studied and the question of how the adsorption onto the clay mineral affects other P-related enzymes (e.g. phytase) has been investigated much less commonly, or not at all. This review, therefore, focused on the influences of several widely studied factors on alkaline and acid phosphatase adsorption and, subsequently, the effects of adsorption on phosphatase activity and other properties.

The electrostatic interaction between an enzyme and the mineral surface depends on the pH, ionic strength of the bathing solution, and background cation charge (Roth & Lenhoff, Reference Roth and Lenhoff1995). At a pH above the point of zero charge of the mineral and below the isoelectric point of the enzyme, i.e. when the enzymes are positively charged, greater adsorption generally occurs with a negatively charged mineral surface (Carrasco et al., Reference Carrasco, Rad and González-Carcedo1995; Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017). pH not only affects the charges on the mineral surface and the enzyme but enzyme conformation is also altered upon pH changes, influencing the adsorption by minerals. For example, a decrease in pH, i.e. an increase in the positive charge on the enzyme, could induce the enzyme to unfold on an electronegative surface (Quiquampoix et al., Reference Quiquampoix, Abadie, Baron, Leprince, Matumoto-Pintro, Ratcliffe, Staunton, Horbett and Brash1995). Furthermore, strongly acidic or alkaline pH could inhibit adsorption by denaturing the enzymes and a near-neutral pH is likely more appropriate for enzyme adsorption (Sarkar & Leonowicz, Reference Sarkar and Leonowicz1989). The effects of ionic strength are less commonly studied, and Leprince and Quiquampoix (Reference Leprince and Quiquampoix1996) found decreased adsorption of acid phosphatases on montmorillonite with increasing ionic strength and they attributed it to the competition between phosphatase and Na+ for negative sites.

The adsorption also depends on the structure of enzymes and the polarity of the mineral surface. Norde (Reference Norde2008) adopted two notions, “hard” and “soft”, to describe proteins and, therefore, enzymes, depending on their behavior upon adsorption. “Hard” enzymes (e.g. rigid or tightly coiled) undergo limited conformational change when adsorbed on the polar mineral surface and the adsorption is only favored in the case of electrostatic attraction. On the other hand, “soft” enzymes undergo structural change upon adsorption and can gain conformational entropy, as discussed in the section Enzyme adsorption mechanisms in clays and clay minerals, via decreased ordered structure, thus facilitating adsorption even on an electrostatically unfavorable surface.

The types of clay minerals may influence their ability to adsorb enzymes and many factors such as swelling, interlayer space, and charge density contribute to their adsorption capacity. Montmorillonite has been found to have a larger adsorption capacity for acid phosphatase than kaolinite, possibly due to large surface area, high charge density, and interlattice fixation of the enzyme in montmorillonite (Gianfreda & Bollag, Reference Gianfreda and Bollag1994; Makboul & Ottow, Reference Makboul and Ottow1979). Greater adsorption of alkaline phosphatase by goethite than by montmorillonite was discovered by Tan et al. (Reference Tan, He, Wang, Li, Kong, Tian, Shen, Megharaj and He2018), while the opposite trend was reported by Zhu et al. (Reference Zhu, Wu, Feng, Liu and Giesy2016). Such discrepancy was attributed to different pH used in the adsorption experiment and, therefore, different charge properties of the phosphatases and minerals (Tan et al., Reference Tan, He, Wang, Li, Kong, Tian, Shen, Megharaj and He2018).

According to Shindo et al. (Reference Shindo, Watanabe, Onaga, Urakawa, Nakahara and Huang2002), the adsorption of an acid phosphatase followed the order: montmorillonite > > kaolinite > Mn oxide > Fe oxide > Al oxide > > allophane. In their study, a positive correlation between the amount of adsorption and mineral-specific surface area could not be established. The charges of the enzyme and mineral surfaces, and thus electrostatic interaction and charge density, mineral microstructure, and ligand exchange capacity of the oxides likely played a larger role in affecting the extent of adsorption. More adsorption of acid phosphatase was observed (Huang et al., Reference Huang, Liang and Cai2005) on goethite than on fine clays; this, in turn, was greater than the adsorption on coarse clays from an Ultisol; and, finally, this, in turn, was also greater than the adsorption on kaolinite. Those authors concluded that large surface area and ion exchange capacity were responsible for greater adsorption.

Even with the same mineral, different saturating cations may influence the extent of adsorption, which was probably due to the difference in how easily the cations can be displaced by the enzyme (Boyd & Mortland, Reference Boyd, Mortland, Bollag and Stotzky1990). For example, Carrasco et al. (Reference Carrasco, Rad and González-Carcedo1995) found that the amount of alkaline phosphatase adsorption followed the order: Ca-sepiolite > Na-sepiolite > H-sepiolite and that H-sepiolite-adsorbed enzyme was desorbed easily under high salinity and pH.

On the contrary, little effect from varying clay minerals has been reported. For instance, Sedaghat et al. (Reference Sedaghat, Ghiaci, Aghaei and Soleimanian-zad2009) found a similar amount of adsorption of an alkaline phosphatase on sepiolite and bentonite despite a much larger specific surface area of sepiolite. They concluded that the enzyme was preferentially immobilized on the external surface of the mineral, instead of penetrating the interlayer. Discrepancies among different literature might be due to different reaction conditions such as pH, the specific microstructure of the mineral, and properties of different enzymes and, therefore, varied electrostatic attraction or repulsion between the mineral surface and the enzyme.

The interaction between clays and organic matter could also affect the adsorption of enzymes because the organic matter may serve as a protective layer on the clay surface, decreasing enzyme adsorption (Quiquampoix et al., Reference Quiquampoix, Abadie, Baron, Leprince, Matumoto-Pintro, Ratcliffe, Staunton, Horbett and Brash1995). Kelleher et al. (Reference Kelleher, Willeford, Simpson, Simpson, Stout, Rafferty and Kingery2004) found less acid phosphatase immobilized on organic molecule-intercalated montmorillonite (i.e. the organic molecules were immobilized between layers of the montmorillonite) than on pure montmorillonite. The opposite trend was also observed. For example, Tang et al. (Reference Tang, Shen, Suib, Coughlin and Vinopal1993) reported more adsorption of alkaline phosphatase by protamine-intercalated bentonite at pH 10.4. Both the phosphatase and mineral surface were negatively charged while the protamine was positively charged, acting as a bridge to connect the enzyme and mineral. Excess protamine decreased phosphatase adsorption due to steric blockage of active sites, however. Huang et al. (Reference Huang, Liang and Cai2005) found that within the same size fraction of clays, natural clays that contained organic matter expressed more enzyme adsorption than inorganic clays (natural clays with organic matter removed by H2O2). According to those authors, the enhanced adsorption was attributed to the additional adsorption of phosphatase by humic substances as well as their ability to trap the phosphatase within the “macromolecular net” of humic acids.

Enzyme Activity and Kinetics Parameters Affected by its Adsorption

Upon adsorption by clay minerals, phosphatase activity was mostly inhibited, while enhanced catalytic activity has seldom been observed (Table 1). Interestingly, enhanced phosphatase activity was more commonly observed when adsorbed on poorly crystalline minerals such as allophane and ferrihydrite. The mesoporosity of allophanes was considered to be responsible for making the minerals more suitable supports in biocatalytic processes (Calabi-Floody et al., Reference Calabi-Floody, Velásquez, Gianfreda, Saggar, Bolan, Rumpel and Mora2012; Wang, Reference Wang2006). From the literature, possible contributors of enhanced enzyme activity upon adsorption could be: greater concentrations of the enzyme and the substrate on the clay mineral surface, conformational change, or stabilization of the enzyme structure induced by adsorption (Allison, Reference Allison2006; An et al., Reference An, Zhou, Zhuang, Tong and Yu2015; Tietjen & Wetzel, Reference Tietjen and Wetzel2003). Inhibited enzyme activity is mostly attributed to conformational change or even deformation of the enzyme, the incorporation of enzymes into the internal structure of the clay, steric hindrance, the interaction between amino acids on the active site and the mineral surface, or diffusional limitation, leading to restricted access of the substrate to the enzyme (Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017; Dick & Tabatabai, Reference Dick and Tabatabai1987; Olagoke et al., Reference Olagoke, Kalbitz and Vogel2019; Tietjen & Wetzel, Reference Tietjen and Wetzel2003; Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011).

Table 1 Experimental conditions and changes in phosphatase properties in phosphatase-clay mineral sorption studies

a The pH and temperature are the conditions used for enzyme activity measurement. RT: room temperature

b Compared to respective properties of the free enzyme

The most commonly used model to study enzyme kinetics is the Michaelis–Menten equation:

(1) v = [ S ] K m + [ S ] V m a x

where v is the velocity of the reaction, [S] the substrate concentration, Km the Michaelis–Menten constant, and V max the maximum velocity. The mechanism of how a clay mineral, i.e. the inhibitor, inhibits the enzyme activity can be represented by changes in Km and V max (Johnson & Goody, Reference Johnson and Goody2011). Four types of reversible inhibition mechanisms are generally acknowledged: competitive inhibition, non-competitive inhibition, uncompetitive inhibition, and mixed inhibition. Competitive inhibition is observed with unchanged V max and increased Km, and the inhibitor competes with the substrate generally in a mutually exclusive form, and often for the same active site on the enzyme. The noncompetitive inhibitor binds with the enzyme or the enzyme–substrate complex with the same affinity, thus leading to unchanged Km and decreased V max. The uncompetitive inhibitor binds only to the enzyme–substrate complex to prevent product formation and, therefore, decreases both the V max and Km. Mixed inhibition is often observed with increased or decreased Km and decreased V max, and the inhibitor can bind to the enzyme or the enzyme–substrate complex with different affinity (Copeland, Reference Copeland2000; Cornish-Bowden, Reference Cornish-Bowden and Cornish-Bowden1979; Fange et al., Reference Fange, Lovmar, Pavlov and Ehrenberg2011).

Mixed inhibition is the most widely observed mechanism for clay minerals with regard to their influence on phosphatase activity (Quiquampiox & Mousain, Reference Quiquampiox, Mousain, Turner, Frossard and Baldwin2005; Shirvani et al., Reference Shirvani, Khalili, Kalbasi, Shariatmadari and Nourbakhsh2020). For instance, kaolinite, kaolin, mica, montmorillonite, tannic acid-montmorillonite complex, goethite, allophane, Fe/Al/Mn oxides, uncalcined and calcined Mg/Al-CO3 layered double hydroxide, and δ-MnO2 have been reported to be inhibitors for many acid phosphatases (Gianfreda & Bollag, Reference Gianfreda and Bollag1994; Huang & Shindo, Reference Huang and Shindo2000; Makboul & Ottow, Reference Makboul and Ottow1979; Paul et al., Reference Paul, Datta, Bera, Math, Bhattacharyya and Dahuja2022; Rao et al., Reference Rao, Violante and Gianfreda2000; Shindo et al., Reference Shindo, Watanabe, Onaga, Urakawa, Nakahara and Huang2002; Zhu et al., Reference Zhu, Huang, Pigna and Violante2010), and palygorskite, sepiolite, goethite, and montmorillonite for some alkaline phosphatases (Ghiaci et al., Reference Ghiaci, Aghaei, Soleimanian and Sedaghat2009; Sedaghat et al., Reference Sedaghat, Ghiaci, Aghaei and Soleimanian-zad2009; Shirvani et al., Reference Shirvani, Khalili, Kalbasi, Shariatmadari and Nourbakhsh2020; Wang et al., Reference Wang, Li, Tan, He, Xie, Megharaj and Wei2017; Zhu et al., Reference Zhu, Wu, Feng, Liu and Giesy2016). Discrepancies and other inhibition mechanisms have also been reported. For example, illite and montmorillonite can be uncompetitive inhibitors for some acid phosphatases (Kelleher et al., Reference Kelleher, Willeford, Simpson, Simpson, Stout, Rafferty and Kingery2004; Makboul & Ottow, Reference Makboul and Ottow1979), but noncompetitive inhibitors for other acid phosphatases (Dick & Tabatabai, Reference Dick and Tabatabai1987). Montmorillonite has also been found to be an uncompetitive inhibitor for alkaline phosphatase (Wang et al., Reference Wang, Li, Tan, He, Xie, Megharaj and Wei2017).

Other Properties Affected by the Adsorption

After being adsorbed, the enzymes experience a different microenvironment (e.g. altered diffusion, charge, or steric hindrance) from that experienced by free enzymes (Zimmerman & Ahn, Reference Zimmerman, Ahn, Shukla and Varma2011). Therefore, in addition to catalytic activity, other properties including optimum pH, enzyme conformation, and thermal and proteolytic stability can be influenced.pH-dependent phosphatase adsorption has been observed by several researchers. For instance, the optimum pH of acid phosphatase was shifted to more alkaline values when they were adsorbed on negatively charged mineral surfaces, which was also attributed to the pH-dependent modification of the enzyme conformation (Leprince & Quiquampoix, Reference Leprince and Quiquampoix1996). Instead of exhibiting an optimum pH, the activity of a montmorillonite-adsorbed alkaline phosphatase continued to decrease with pH increasing from 4.0 to 8.0 (Rao et al., Reference Rao, Violante and Gianfreda2000). No change in the optimum pH was reported when an alkaline phosphatase was immobilized on sepiolite, bentonite, and protamine-intercalated bentonite in some studies. This was possibly due to high ionic strength and the compression of the diffuse double layer of the mineral, leading to less local pH variation on the clay mineral surface from the solution pH (Carrasco et al., Reference Carrasco, Rad and González-Carcedo1995; Ghiaci et al., Reference Ghiaci, Aghaei, Soleimanian and Sedaghat2009; Sedaghat et al., Reference Sedaghat, Ghiaci, Aghaei and Soleimanian-zad2009; Tang et al., Reference Tang, Shen, Suib, Coughlin and Vinopal1993). Adsorption of acid phosphatase on goethite, kaolinite, and clay minerals from an Ultisol failed to affect significantly the enzyme’s optimum pH, and the adsorbed phosphatases expressed less sensitivity toward pH changes (Huang et al., Reference Huang, Liang and Cai2005). Unlike the more widely investigated reasons for shifted optimum pH, the factors contributing to unchanged phosphatase optimum pH are poorly understood and have not been studied extensively; studies in the past decade did not probe the optimum pH of adsorbed phosphatase.

Conformational changes of adsorbed enzymes have been recognized (Table 1, Fig. 3). The attractive electrostatic force between minerals and the enzyme could induce the unfolding of the enzyme, exposing the hydrophobic chain, and leading to a hydrophobic interaction with the siloxane layers of phyllosilicates (Staunton & Quiquampoix, Reference Staunton and Quiquampoix1994). Hydrophobic interaction between the mineral and enzyme can also induce a conformational change, as discussed in the section Enzyme adsorption mechanisms in clays and clay minerals. The extent of conformational change depends on the balance between several factors including the intramolecular forces of the enzyme, the effects from the mineral surface, and the solvent molecule (Boyd & Mortland, Reference Boyd, Mortland, Bollag and Stotzky1990; Datta et al., Reference Datta, Anand, Moulick, Baraniya, Pathan, Rejsek, Vranova, Sharma, Sharma, Kelkar and Formanek2017). Some enzyme structures, e.g. fibrous proteins, are more likely to unfold due to more interaction with the mineral surface upon adsorption (Boyd & Mortland, Reference Boyd, Mortland, Bollag and Stotzky1990).

An increased ordered secondary structure of phosphatase upon mineral adsorption has been reported. The conformational change of an acid phosphatase when adsorbed on montmorillonite, kaolinite, and inorganic and organic colloids from an Alfisol were studied by Huang et al. (Reference Huang, Zhu, Qiao, Cai, Rong, Liang and Chen2009). This change increased the ordered secondary structure of the phosphatase and the 2:1 clay led to a more significant change than the 1:1 clay. Such a difference was attributed to the difference in the hydrophobicity between kaolinite and montmorillonite. All of the external planar surfaces of montmorillonite are siloxane surfaces which are typically hydrophobic. As for kaolinite, half of the external planar surface is siloxane and, therefore, the kaolinite surface is considered to be less hydrophobic. The more hydrophobic surface of montmorillonite promoted the formation of intramolecular hydrogen bonds.

Adsorption of enzymes on clay minerals can serve as a means to protect the enzymes against, for example, proteolysis, heat, light, and other inhibitors (Huang et al., Reference Huang, Zhu, Qiao, Cai, Rong, Liang and Chen2009; Rao et al., Reference Rao, Violante and Gianfreda2000; Sarkar & Leonowicz, Reference Sarkar and Leonowicz1989). Different minerals vary in their ability to affect the sensitivity of phosphatases toward proteolysis (Table 1). For example, adsorption by montmorillonite was found to provide more proteolytic protection than kaolinite (Huang et al., Reference Huang, Zhu, Qiao, Cai, Rong, Liang and Chen2009). Increased proteolytic resistance was attributed to the conformational change of the enzyme when immobilized, and to the alteration of the cleavage site, disturbing the recognition from the proteinase. A more significant extent of the conformational change induced by montmorillonite than kaolinite led to greater resistance of montmorillonite-adsorbed phosphatase. In another study, increased hydrolytic stability of an acid phosphatase was induced in the order: adsorption on calcined Mg/Al-CO3 layered double hydroxide (LDH) > adsorption on uncalcined Mg/Al-CO3 LDH > enzyme-free (Zhu et al., Reference Zhu, Huang, Pigna and Violante2010). The latter authors suggested that the enzymes penetrated into the pores on the porous surface of LDH, rendering them less accessible for protease and, therefore, the greater porosity of calcined minerals accounted for more elevated proteolytic resistance of the enzyme. The difference in the proteolytic stability may be a result of the specific mineral site where the enzyme is adsorbed. If the enzyme is fixed in the interlayer space, it would be inaccessible to the protease; hence there is an increased proteolytic resistance. On the other hand, if the enzyme is immobilized on the surface or edges of the mineral which can act as a support for concentrated enzymes and proteases, the proteolysis would be enhanced, resulting in decreased proteolytic stability of immobilized enzyme (Boyd & Mortland, Reference Boyd and Mortland1985).

As for the effects on thermal stability, less temperature sensitivity from montmorillonite-adsorbed acid phosphatase was reported by Rao et al. (Reference Rao, Violante and Gianfreda2000). The acid phosphatase adsorbed on kaolinite and goethite had greater thermal stability than the free enzyme (Huang et al., Reference Huang, Liang and Cai2005). The greater thermal stability of kaolinite-adsorbed than goethite-adsorbed acid phosphatase was probably a result of tighter binding on kaolinite as less desorption was detected from kaolinite. They also reported greater thermal stability of an organic clay-adsorbed acid phosphatase, which was attributed to the protective effect of organic matter. Zhu et al. (Reference Zhu, Huang, Pigna and Violante2010) observed enhanced thermal stability of uncalcined and calcined Mg/Al-CO3 LDH-adsorbed acid phosphatase than the free enzyme, which was attributed to the loss of enzyme conformational flexibility. Increased thermal stability was also observed for the alkaline phosphatase adsorbed on a bilayer-surfactant-covered sepiolite (Sedaghat et al., Reference Sedaghat, Ghiaci, Aghaei and Soleimanian-zad2009). Decreased thermal stability of Na-sepiolite-adsorbed alkaline phosphatase was attributed to a rapid loss of moisture from the clay during the incubation, which affected the immobilized enzyme (Carrasco et al., Reference Carrasco, Rad and González-Carcedo1995). No significant changes in the thermal stability of an alkaline phosphatase immobilized on bentonite were observed (Ghiaci et al., Reference Ghiaci, Aghaei, Soleimanian and Sedaghat2009). Light sensitivity has been studied less, and Tietjen and Wetzel (Reference Tietjen and Wetzel2003) found less light sensitivity and photodegradation of montmorillonite- and lake basin clay-immobilized alkaline phosphatase.

Adsorption on clay minerals may protect the enzyme against other inhibitors, e.g. metal cations (aq) or oxyanions. For instance, a decreased inhibitory effect of Cd2+ on the activity of palygorskite-, sepiolite-, goethite-, and montmorillonite-adsorbed alkaline phosphatase than free phosphatase has been observed (Shirvani et al., Reference Shirvani, Khalili, Kalbasi, Shariatmadari and Nourbakhsh2020; Tan et al., Reference Tan, He, Wang, Li, Kong, Tian, Shen, Megharaj and He2018). Wang et al. (Reference Wang, Li, Tan, He, Xie, Megharaj and Wei2017) reported less sensitivity of goethite- and montmorillonite-adsorbed alkaline phosphatase toward the inhibitory effects of arsenate. Rosas et al. (Reference Rosas, Luz Mora, Jara, López, Rao and Gianfreda2008) reported the protective effect of an allophanic clay on an acid phosphatase against Mn2+ and Mo7O4 inhibition. Huang and Shindo (Reference Huang and Shindo2000) observed different effects from kaolin, goethite, and δ-MnO2 in protecting an acid phosphatase against Cu2+ inhibition. At the same CuCl2 concentration, the goethite-adsorbed enzyme showed greater activity than the δ-MnO2-adsorbed enzyme and this, in turn, was greater than the activity of the kaolin-adsorbed enzyme; the activity of the latter was similar to that of the free enzyme. When copper citrate was used, however, its inhibitory effect was more significant on the clay-adsorbed enzymes than on the free ones. The observed difference induced by varied copper forms was postulated to be a result of the interaction between the carboxylic group of the citrate and the hydroxyl group on the mineral surface, modifying the enzyme conformation.

The protection was attributed to the adsorption of the inhibitors by clay minerals, resulting in lowered inhibitor concentration in the solution and decreased interaction between inhibitors and enzyme active sites. Another factor contributing to the diminished inhibition effect from the inhibitors may be the conformational change of the enzyme after adsorption, burying the active site, which would limit access by the inhibitors (Shirvani et al., Reference Shirvani, Khalili, Kalbasi, Shariatmadari and Nourbakhsh2020; Wang et al., Reference Wang, Li, Tan, He, Xie, Megharaj and Wei2017). The extent of protection depends on many factors, including the concentration and identity of the inhibitor, exposure time, order of addition of the inhibitor, and clay mineral type. For example, adsorption on montmorillonite and goethite diminished the sensitivity of an alkaline phosphatase toward Cd2+ inhibition with an exposure time of >1 h. The opposite effect was observed when the exposure to Cd2+ was <1 h (Tan et al., Reference Tan, He, Wang, Li, Kong, Tian, Shen, Megharaj and He2018). When Mn2+ was added together with allophane to an acid phosphatase, its activity decreased with increasing Mn2+ concentration. When added to allophane-adsorbed acid phosphatase, Mn2+ did not affect significantly the enzyme activity. On the contrary, the opposite trend was observed with Mo7O4 (Rosas et al., Reference Rosas, Luz Mora, Jara, López, Rao and Gianfreda2008).

The adsorption of phosphatases on clay minerals is, therefore, usually a tradeoff between decreased activity and protection against environmental stress (Fig. 4). From the literature, it is not surprising to find that even with the same mineral, the inhibition mechanism and the way in which the enzyme properties are affected can be different. Effects of different forces in the adsorption and the variety of enzyme sources and, thus, enzyme species and behavior (e.g. conformation, orientation on the mineral surface) with respect to different reaction conditions make enzyme adsorption a complex process and increase the difficulty of predicting the results. Specific enzyme structure and characteristics and reaction conditions including pH and temperature could all lead to different properties and behavior of the mineral surface and enzyme. If one were to investigate or compare the effects of a certain mineral on an enzyme, careful attention should, therefore, be paid to clay and enzyme species and individual reaction conditions.

Fig. 4 Phosphatase activities in the terrestrial environment. The role of clays and clay minerals as phosphatase sinks is highlighted. The interaction could suppress P mineralization and enhance enzyme stability

The Phosphatase–Mineral Interaction affects the Fate of P in Soils

The fate of P in soils is dependent on the biological and geochemical activities in the soil and, from the aspect of biology, the release of bioavailable P is achieved through mineralization (e.g. by phosphatase) and solubilization (e.g. by phosphorus-solubilizing microorganisms) (Alori et al., Reference Alori, Glick and Babalola2017; Hussain et al., Reference Hussain, Phillips, Hu, Frey, Geuder, Edwards, Lapen, Ptacek and Blowes2021). Up to 85% of the total P in soils may exist as organic P and a meta-analysis study that summarized results from 149 soils showed that ~30–60% of the organic P in the studied soils can be mineralized by phosphatases and/or phytases (Bünemann, Reference Bünemann2008; Dalal, Reference Dalal1977; Tarafdar et al., Reference Tarafdar, Yadav and Meena2001).

An enhanced phosphatase activity promotes the mineralization of organic P and the release of phosphate (Fig. 2), which is a more bioavailable fraction. For instance, Tarafdar et al. (Reference Tarafdar, Yadav and Meena2001) observed a positive linear relationship between the activity of acid phosphatase and the amount of phosphate released from different organic P compounds. Zou et al. (Reference Zou, Binkley and Caldwell1995) found a positive correlation between phosphatase activity and gross P mineralization rate. Accordingly, depletion of organic P was usually observed with increasing phosphatase activity in the bulk soil (McConnell et al., Reference McConnell, Kaye and Kemanian2020; Schaap et al., Reference Schaap, Fuchslueger, Hoosbeek, Hofhansl, Martins, Valverde-Barrantes, Hartley, Lugli and Quesada2021). An enhanced depletion of different organic P forms (e.g. sodium hydroxide-extractable, bicarbonate-extractable) was also found to be related to increased activity of phosphatases in the rhizosphere of rape, onion, wheat, clover, barley, Norway spruce, and radiata pine, indicating the important role of phosphatases in the organic P mineralization and in providing bioavailable P for plants (Asmar et al., Reference Asmar, Singh, Gahoonia and Nielsen1995; Chen et al., Reference Chen, Condron, Davis and Sherlock2002; Häussling & Marschner, Reference Häussling and Marschner1989; Liu et al., Reference Liu, Loganathan, Hedley and Skinner2004; Tarafdar & Jungk, Reference Tarafdar and Jungk1987).

As well as for phosphate, other P forms are also affected by phosphatase activities. Margalef et al. (Reference Margalef, Sardans, Fernández-Martínez, Molowny-Horas, Janssens, Ciais, Goll, Richter, Obersteiner, Asensio and Peñuelas2017) collated data from 183 studies from 378 sites around the world and found that some P forms may be correlated to the activity of phosphatases in the soil. They found a positive correlation between alkaline phosphatase activity and Olsen-P (the “Olsen-P test” provides estimation on the available P content in the soil), and between organic P and acid phosphatase activity. Yu et al. (Reference Yu, He, Stoffella, Calvert, Yang, Banks and Baligar2006) observed significant correlations between acid and/or alkaline phosphatase activity and different P fractions, including hydrochloric acid-extractable P, Olsen P, and total P, and significant correlations between neutral, as well as natural phosphatase activity (i.e. measured at soil pH) and P fractions such as Olsen P, water-extractable inorganic P, and sodium bicarbonate-extractable organic P. Compton and Cole (Reference Compton and Cole2001) observed a positive relationship between phosphatase activity in the bulk soil and some organic P forms (e.g. the organic P after sodium bicarbonate extraction and labile organic P) at pH 5, which might be attributed to the stimulation of phosphatase excretion induced by high organic P content.

The promoted release of phosphate and other labile P by phosphatases, while more bioavailable, are also prone to leaching. Yu et al. (Reference Yu, He, Stoffella, Calvert, Yang, Banks and Baligar2006) found natural phosphatase activity to be related to the concentration of total P, total dissolved P, and phosphate in surface runoff. A negative correlation was observed between alkaline phosphatase activity and total P and total dissolved P. No correlations were found between acid phosphatase activity and any P fractions, which was probably due to the neutral to slightly alkaline nature of the runoff water. Yu et al. (Reference Yu, He, Stoffella, Calvert, Yang, Banks and Baligar2006) proposed, therefore, that soil natural phosphatase activity may be an index for the P-loss potential by surface runoff. A positive correlation was observed between neutral phosphatase activity and total P in the surface water of a rice paddy field, indicating a possible enhanced P runoff loss (Wang et al., Reference Wang, Liang, Chen, Luo, Liang, Li, Huang, Li, Wan, Li and Shao2012).

The interaction between soil minerals and phosphatases clearly affects phosphatase activity, which in turn influences both soil P availability and loss. Depending on which results are desired, either promoting or suppressing the adsorption of phosphatases on the mineral surfaces could be the goal of soil management. For instance, in strongly weathered soils, the organic P and secondary minerals (e.g. sasaite, wavellite, crandallite, cacoxenite) are usually the predominant P forms and these soils are often observed with limited P availability (Kovács et al., Reference Kovács, Farics, Szabó and Sajó2020; Margalef et al., Reference Margalef, Sardans, Fernández-Martínez, Molowny-Horas, Janssens, Ciais, Goll, Richter, Obersteiner, Asensio and Peñuelas2017; Rocha et al., Reference Rocha, Menegale, Rodrigues, de M Gonçalves, Pavinato, Foltran, Harrison and James2019; Wang et al., Reference Wang, Xie, Zhang and Wang2023). Furthermore, with the long-term development of ecosystems (e.g. forest), available P is gradually lost through surface runoff, for example (Turner et al., Reference Turner, Lambers, Condron, Cramer, Leake, Richardson and Smith2013). Phosphorus depletion and limitation in agricultural soils not only affect yield and decrease biomass but also influence plant community composition and diversity (Turner & Condron, Reference Turner and Condron2013). For such soils, mineralization by extracellular enzymes and maintaining their stability (e.g. through clay adsorption) and activity are particularly crucial in providing bioavailable P (Fig. 4). Some studies observed enhanced phosphatase activity when adsorbed on the mineral surface, which can lead to useful application. Many heavy metals (e.g. Cd, Pb, Cu, Hg, As, W) inhibit phosphatase activity and in some heavy metal-polluted soils, the inhibitory effect, therefore, becomes a severe problem (Hong et al., Reference Hong, Kim, Lee, Yang and Kim2020; Huang & Shindo, Reference Huang and Shindo2000; Mao et al., Reference Mao, Tang, Feng, Gao, Zhou, Xu and Wang2015; Yang et al., Reference Yang, Liu, Zheng and Feng2006). Phosphatase–clay interaction may be used to enhance organic P mineralization or to protect phosphatase against inhibitors or environmental stress, but the real-world application is still limited. For example, an analysis of the inhibition constants and ecological dose of arsenate (Tian et al., Reference Tian, Zhao, Megharaj and He2018) showed that the adsorbents (montmorillonite or soil) protected alkaline phosphatase from arsenate deactivation. Li and Xu (Reference Li and Xu2018) found an increase in phosphatase activity and available P level in a Cd-polluted rice field when sepiolite was added. Although this could also be a result of Cd immobilization by clay minerals and, thus, decreased Cd, the addition of adsorbed or immobilized enzymes to soils still has the potential as one of the applications of enzyme–clay mineral association.

Industrial Application of Enzyme-Mineral Complexes

Many natural and anthropogenic activities contribute to the contamination of soils and waters, and the application of free enzymes or adsorbed enzymes has been developed as a means of bioremediation. One of the main requirements for effective bioremediation by enzymes is the stability of the enzyme under the environmental conditions of the contaminated site (Gianfreda & Rao, Reference Gianfreda and Rao2004). The indigenous extracellular enzymes or free enzymes added to soils are prone to rapid denaturation or degradation due to, for example, microbial and enzymatic degradation or chemical-mediated denaturation, thus generally having a short lifetime (Gianfreda et al., Reference Gianfreda, Rao, Sannino, Saccomandi and Violante2002; Sarkar & Leonowicz, Reference Sarkar and Leonowicz1989). The association of minerals and enzymes, however, can largely contribute to the stabilization and persistence of the extracellular enzymes in soils (Boyd & Mortland, Reference Boyd, Mortland, Bollag and Stotzky1990). As mentioned above, greater thermal stability, proteolysis stability, storage stability, and protection against other inhibitors may result from the adsorption to minerals. Upon adsorption and with increased stability, although usually at a cost of reduced activity, the enzymes can survive longer with long-term retention of activity and contribute to promoting the degradation of xenobiotic substances in aqueous and terrestrial environments to produce less toxic compounds (Gianfreda & Rao, Reference Gianfreda and Rao2004; Gianfreda et al., Reference Gianfreda, Rao, Sannino, Saccomandi and Violante2002; Hoehamer et al., Reference Hoehamer, Mazur and Wolfe2005; Meng et al., Reference Meng, Jiang, Li, Huang, Zhai, Zhang, Guan, Cai and Xiangru2019). Compared to the use of free enzymes, other advantages of applying adsorbed enzymes include improved enzyme reusability, enhanced enantioselectivity, and they can usually be recovered at the end of the process (Burns & Dick, Reference Burns and Dick2002; Chen et al., Reference Chen, Guo, Xin, Gu, Zhang and Guo2023; Gianfreda & Rao, Reference Gianfreda and Rao2004).

Organophosphorus pesticides (e.g. glyphosate), accounting for ~40% of the global pesticide market, have been used commonly in agriculture and forestry (Chen et al., Reference Chen, Guo, Xin, Gu, Zhang and Guo2023). They are highly toxic and disposal methods might be inefficient and produce toxic pollution (Chen et al., Reference Chen, Guo, Xin, Gu, Zhang and Guo2023). As an alternative, using enzymes as a bioremediation method could be a more sustainable, renewable, and green approach that is more environmentally friendly and less invasive (Dzionek et al., Reference Dzionek, Wojcieszynska and Guzik2016; Somu et al., Reference Somu, Narayanasamy, Gomez, Rajendran, Lee and Balakrishnan2022). For instance, organophosphorus hydrolases (e.g. phosphotriesterase, parathion hydrolase), organophosphate acid anhydrase, and methyl parathion hydrolase can mineralize organophosphorus compounds such as (methyl) paraoxon, (methyl) parathion, and soman. Nitrilases can be used for degrading herbicides, polymers, plastics, and cyanide (Mousavi et al., Reference Mousavi, Hashemi, Iman, Ravan, Gholami, Lai, Chiang, Omidifar, Yousefi and Behbudi2021). Enzymes such as tyrosinase, phenoloxidases, and peroxidases have been used for the removal of phenols, aromatic amines, and related compounds (Gianfreda et al., Reference Gianfreda, Rao, Sannino, Saccomandi and Violante2002; Mousavi et al., Reference Mousavi, Hashemi, Iman, Ravan, Gholami, Lai, Chiang, Omidifar, Yousefi and Behbudi2021; Somu et al., Reference Somu, Narayanasamy, Gomez, Rajendran, Lee and Balakrishnan2022).

The application of free and immobilized enzymes has, therefore, become more common. Daumann et al. (Reference Daumann, Larrabee, Ollis, Schenk and Gahan2014) found that glycerophosphodiesterases immobilized on functionalized magnetite nanoparticles had prolonged storage time and can be used for bioremediation of organophosphate pollution. Laccases have wide substrate specificity and can be used for degrading a variety of contaminants such as dyes, pesticides, benzenediol, and polycyclic aromatic hydrocarbons (Mousavi et al., Reference Mousavi, Hashemi, Iman, Ravan, Gholami, Lai, Chiang, Omidifar, Yousefi and Behbudi2021). Montmorillonite- and kaolinite-adsorbed laccase and peroxidase have been used to transform and detoxify 2,4-dichlorophenol, a degradation product from a herbicide (Gianfreda & Bollag, Reference Gianfreda and Bollag1994; Ruggiero et al., Reference Ruggiero, Sarkar and Bollag1989). The removal efficiency has been confirmed both with and without the presence of soils. The authors also reported that the immobilized enzyme can be separated from the reaction mixture and used repeatedly. Masaphy et al. (Reference Masaphy, Fahima, Levanon, Henis and Mingelgrin1996) used a crude parathion-degrading enzyme extract from Xanthomonas sp. to degrade parathion, an organophosphate pesticide. To simulate the soil environment, the enzyme was added to a Na-montmorillonite suspension for parathion degradation. Their study suggested the potential of using the enzyme as an approach in soil decontamination.

Many enzymes have also been used for the bioremediation of haloorganics and heavy metal (e.g. Se, As, Cr, Hg) pollution, which can be an indication of the future application of phosphatases (Burns & Dick, Reference Burns and Dick2002). With the enzyme stabilization brought about via enzyme–clay mineral interaction, the beneficial influences such as eco-friendly remediation and, in some cases, enhanced mineralization can be prolonged with reusable enzymes. To further induce the excretion of enzymes and, therefore, to improve the remediation effect and mineralization efficiency, methods such as adjusting soil properties (e.g. moisture, pH, and fertilizer N amendment) can be applied (Arenberg & Arai, Reference Arenberg, Arai and Sparks2019).

Enzymes adsorbed on the mineral surfaces are also applied in other fields such as the biomedical, biosensor, biocatalyst, food, pharmaceutical, textile, and detergent industries, and in wastewater treatment (An et al., Reference An, Zhou, Zhuang, Tong and Yu2015; Basso & Serban, Reference Basso and Serban2019; Jesionowski et al., Reference Jesionowski, Zdarta and Krajewska2014; Maghraby et al., Reference Maghraby, El-Shabasy, Ibrahim and Azzazy2023). Adsorption can increase thermal stability and protect enzymes against denaturation during the purification process in the food industry (Hassan et al., Reference Hassan, Yang, Xiao, Liu, Wang, Cui and Yang2019). In the pharmaceutical industry, a loss of activity during shelf life can be a problem, and enzymes adsorbed to, e.g. silica surfaces, with greater stability may help alleviate the problem (Norde, Reference Norde2008). Amorphous clay minerals such as allophane can adsorb water from the surrounding environment, providing an aqueous environment for adsorbed enzymes, which can be applied to the biotechnological field (Calabi-Floody et al., Reference Calabi-Floody, Velásquez, Gianfreda, Saggar, Bolan, Rumpel and Mora2012).

Future Perspectives

Previous studies of the phosphatase–mineral interaction investigated the macroscopic behavior of mineralization, such as phosphatase activity and its thermal and/or proteolysis stability. The results showed that the interaction between phosphatases and minerals was a trade-off between mostly inhibited enzyme activity and protection against the environmental stress. Various and complex phosphatase behavior in the literature was attributed to the variety of specific properties of the enzymes and minerals involved. The wide application of adsorbed enzymes, the complexity of the phosphatase–mineral reactions, and the significant role of phosphatase in affecting P fate in soils have, however, made it important to further investigate the microscopic characteristics of the interaction and reaction mechanisms (e.g. the three-dimensional conformational change of the enzyme, effects on detailed biochemical pathway) between phosphatases and clay minerals. Spectroscopic techniques such as nuclear magnetic resonance (NMR), electron paramagnetic resonance (EPR), and in situ attenuated total reflectance Fourier-transform infrared (ATR-FTIR) are robust in probing such characteristics, and their recent applications have provided more insights into the microscopic aspects of the protein–clay mineral interactions (Reardon et al., Reference Reardon, Chacon, Walter, Bowden, Washton and Kleber2016; Schmidt & Martínez, Reference Schmidt and Martínez2016; Ustunol et al., Reference Ustunol, Coward, Quirk and Grassian2021). The future microscopic study of phosphatase–clay mineral interaction could, therefore, benefit from these techniques. Furthermore, description of the specific reaction conditions and well-characterized macro- and microscopic properties of the specific phosphatase and minerals could be very helpful when investigating their interaction and comparing results from different studies.

Acknowledgements

The authors gratefully acknowledge the USDA National Institute of Food and Agriculture (Hatch 875-967) for supporting this project financially.

Data Availability

Data available on request from the authors.

Declarations

Conflict of Interest

The authors declare no conflict of interest.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Footnotes

Associate Editor: Chun-Hui Zhou

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

References

Abel, S., Ticconi, C. A., Delatorre, C. A. Phosphate sensing in higher plants Physiologia Plantarum 2002 115 18 10.1034/j.1399-3054.2002.1150101.xCrossRefGoogle ScholarPubMed
Abelson, P. H. A potential phosphate crisis Science 1999 283 2015 10.1126/science.283.5410.2015CrossRefGoogle ScholarPubMed
Alef, K., Nannipieri, P. Alef, K., Nannipieri, P. Enzyme activities Methods in applied soil microbiology and biochemistry 1995 Acedemic Press 311373Google Scholar
Alkorta, I., Aizpurua, A., Riga, P., Albizu, I., Amézaga, I., Garbisu, C. Soil enzyme activities as biological indicators of soil health Reviews on Environmental Health 2003 18 6573 10.1515/REVEH.2003.18.1.65CrossRefGoogle ScholarPubMed
Allison, S. D. Soil minerals and humic acids alter enzyme stability: Implications for ecosystem processes Biogeochemistry 2006 81 361373 10.1007/s10533-006-9046-2CrossRefGoogle Scholar
Alori, E. T., Glick, B. R., Babalola, O. O. Microbial phosphorus solubilization and its potential for use in sustainable agriculture Frontiers in Microbiology 2017 8 971 10.3389/fmicb.2017.00971CrossRefGoogle ScholarPubMed
An, N., Zhou, C. H., Zhuang, X. Y., Tong, S. D., Yu, H. W. Immobilization of enzymes on clay minerals for biocatalysts and biosensors Applied Clay Science 2015 114 283296 10.1016/j.clay.2015.05.029CrossRefGoogle Scholar
Anderson, G. Khasawneh chairman, F. E., Sample, E. C., Kamprath, E. J. Assessing organic phosphorus in soils The role of phosphorus in agriculture 1980 American Society of Agronomy 411431Google Scholar
Arenberg, M. R., Arai, Y. Sparks, D. L. Uncertainties in soil physicochemical factors controlling phosphorus mineralization and immobilization processes Advances in agronomy, 154 2019 Acedemic Press 153200Google Scholar
Asmar, F., Singh, T., Gahoonia, , Nielsen, N. E. Barley genotypes differ in activity of soluble extracellular phosphatase and depletion of organic phosphorus in the rhizosphere soil Plant and Soil 1995 172 117122 10.1007/BF00020865CrossRefGoogle Scholar
Bariola, P. A., Howard, C. J., Crispin, B., Verburg, M. T., Jaglan, V. D., Green, P. J. The Arabidopsis ribonuclease gene RNS1 is tightly controlled in response to phosphate limitation The Plant Journal 1994 6 673685 10.1046/j.1365-313X.1994.6050673.xCrossRefGoogle ScholarPubMed
Basso, A., Serban, S. Industrial applications of immobilized enzymes— A review Molecular Catalysis 2019 479 110607 10.1016/j.mcat.2019.110607CrossRefGoogle Scholar
Beghin, J. C., & Nogueira, L. (2021). A perfect storm in fertilizer markets. https://cap.unl.edu/crops/perfect-storm-fertilizer-markets. Accessed 25 Apr 2023.Google Scholar
Blake, R. E., O’Neil, J. R., Surkov, A. V. Biogeochemical cycling of phosphorus: Insights from oxygen isotope effects of phosphoenzymes American Journal of Science 2005 305 596620 10.2475/ajs.305.6-8.596CrossRefGoogle Scholar
Bollag, J. M., Mertz, T., Otjen, L. Anderson, T. A., Coats, J. R. Role of microorganisms in soil bioremediation Bioremediation through rhizosphere technology 1994 American Chemical Society 210 10.1021/bk-1994-0563.ch001CrossRefGoogle Scholar
Boyd, S. A., Mortland, M. M. Urease activity on a clay-organic complex Soil Science Society of America Journal 1985 49 619622 10.2136/sssaj1985.03615995004900030018xCrossRefGoogle Scholar
Boyd, S. A., Mortland, M. M. Bollag, J. M., Stotzky, G. Enzyme interactions with clays and clay-organic matter complexes Soil biochemistry 1990 Marcel Dekker, Inc. 128Google Scholar
Bünemann, E. K. Enzyme additions as a tool to assess the potential bioavailability of organically bound nutrients Soil Biology and Biochemistry 2008 40 21162129 10.1016/j.soilbio.2008.03.001CrossRefGoogle Scholar
Burns, R. G. Enzyme activity in soil: Location and a possible role in microbial ecology Soil Biology and Biochemistry 1982 14 423427 10.1016/0038-0717(82)90099-2CrossRefGoogle Scholar
Burns, R. G., Dick, R. P. Enzymes in the environment: Activity, ecology and applications 2002 Marcel DekkerCrossRefGoogle Scholar
Burns, R. G., DeForest, J. L., Marxsen, J., Sinsabaugh, R. L., Stromberger, M. E., Wallenstein, M. D., Weintraub, M. N., Zoppini, A. Soil enzymes in a changing environment: Current knowledge and future directions Soil Biology and Biochemistry 2013 58 216234 10.1016/j.soilbio.2012.11.009CrossRefGoogle Scholar
Calabi-Floody, M., Velásquez, G., Gianfreda, L., Saggar, S., Bolan, N., Rumpel, C., Mora, M. L. Improving bioavailability of phosphorous from cattle dung by using phosphatase immobilized on natural clay and nanoclay Chemosphere 2012 89 648655 10.1016/j.chemosphere.2012.05.107CrossRefGoogle ScholarPubMed
Carrasco, M. S., Rad, J. C., González-Carcedo, S. Immobilization of alkaline phosphatase by sorption on Na-sepiolite Bioresource Technology 1995 51 175181 10.1016/0960-8524(94)00115-HCrossRefGoogle Scholar
Chen, A., Arai, Y. Sparks, D. L. Current uncertainties in assessing the colloidal phosphorus loss from soil Advances in agronomy, 163 2020 Acedemic Press 117151Google Scholar
Chen, C. R., Condron, L. M., Davis, M. R., Sherlock, R. R. Phosphorus dynamics in the rhizosphere of perennial ryegrass (Lolium perenne L.) and radiata pine (Pinus Radiata D. Don.) Soil Biology and Biochemistry 2002 34 487499 10.1016/S0038-0717(01)00207-3CrossRefGoogle Scholar
Chen, J., Guo, Z., Xin, Y., Gu, Z., Zhang, L., Guo, X. Effective remediation and decontamination of organophosphorus compounds using enzymes: From rational design to potential applications Science of the Total Environment 2023 867 161510 10.1016/j.scitotenv.2023.161510CrossRefGoogle ScholarPubMed
Chin, J. P., McGrath, J. W., Quinn, J. P. Microbial transformations in phosphonate biosynthesis and catabolism, and their importance in nutrient cycling Current Opinion in Chemical Biology 2016 31 5057 10.1016/j.cbpa.2016.01.010CrossRefGoogle ScholarPubMed
Chu, H. M., Guo, R. T., Lin, T. W., Chou, C. C., Shr, H. L., Lai, H. L., Tang, T. Y., Cheng, K. J., Selinger, B. L., Wang, AHJ Structures of Selenomonas ruminantium phytase in complex with persulfated phytate: DSP phytase fold and mechanism for sequential substrate hydrolysis Structure 2004 12 20152024 10.1016/j.str.2004.08.010CrossRefGoogle ScholarPubMed
Coleman, J. E. Structure and mechanism of alkaline phosphatase Annual Review of Biophysics and Biomolecular Structure 1992 21 441483 10.1146/annurev.bb.21.060192.002301CrossRefGoogle ScholarPubMed
Compton, J. E., Cole, D. W. Fate and effects of phosphorus additions in soils under N2-fixing red alder Biogeochemistry 2001 53 225247 10.1023/A:1010646709944CrossRefGoogle Scholar
Copeland, R. A. Enzymes: A practical introduction to structure, mechanism, and data analysis 2000 Wiley-VCH, Inc. 10.1002/0471220639CrossRefGoogle Scholar
Cordell, D., White, S. Life’s bottleneck: Sustaining the world’s phosphorus for a food secure future Annual Review of Environment and Resources 2014 39 161188 10.1146/annurev-environ-010213-113300CrossRefGoogle Scholar
Cornish-Bowden, A. (1979). Inhibitors and activators. In Cornish-Bowden, A. (ed.), Fundamentals of enzyme kinetics (pp. 7398). Butterworths, Oxford, UK.CrossRefGoogle Scholar
Cross, H., & Yariv, S. (1979). Colloid geochemistry of clay minerals. In Cross, H., & Yariv, S. (eds.), Geochemistry of colloid systems (pp. 247285). Berlin Heidenberg, New York: Springer-Verlag.Google Scholar
Dalal, R. C. Soil organic phosphorus Advances in Agronomy 1977 29 83117 10.1016/S0065-2113(08)60216-3CrossRefGoogle Scholar
Datta, R., Anand, S., Moulick, A., Baraniya, D., Pathan, S. I., Rejsek, K., Vranova, V., Sharma, M., Sharma, D., Kelkar, A., Formanek, P. How enzymes are adsorbed on soil solid phase and factors limiting its activity: A review International Agrophysics 2017 31 287302 10.1515/intag-2016-0049CrossRefGoogle Scholar
Daughton, C. G., Cook, A. M., Alexander, M. Biodegradation of phosphonate toxicants yields methane or ethane on cleavage of the C-P bond FEMS Microbiology Letters 1979 5 9193 10.1111/j.1574-6968.1979.tb03254.xCrossRefGoogle Scholar
Daumann, L. J., Larrabee, J. A., Ollis, D., Schenk, G., Gahan, L. R. Immobilization of the enzyme GpdQ on magnetite nanoparticles for organophosphate pesticide bioremediation Journal of Inorganic Biochemistry 2014 131 17 10.1016/j.jinorgbio.2013.10.007CrossRefGoogle ScholarPubMed
Dick, W. A., Tabatabai, M. A. Kinetics and activities of phosphatase-clay complexes Soil Science 1987 143 515 10.1097/00010694-198701000-00002CrossRefGoogle Scholar
Dionisio, G., Madsen, C. K., Holm, P. B., Welinder, K. G., Jorgensen, M., Stoger, E., Arcalis, E., Brinch-pedersen, H. Cloning and characterization of purple acid phosphatase phytases from wheat, barley, maize, and rice Plant Physiology 2011 156 10871100 10.1104/pp.110.164756CrossRefGoogle ScholarPubMed
Dzionek, A., Wojcieszynska, D., Guzik, U. Natural carriers in bioremediation: A review Electronic Journal of Biotechnology 2016 23 2836 10.1016/j.ejbt.2016.07.003CrossRefGoogle Scholar
Faba-Rodriguez, R., Gu, Y., Salmon, M., Dionisio, G., Brinch-Pedersen, H., Brearley, C. A., Hemmings, A. M. Structure of a cereal purple acid phytase provides new insights to phytate degradation in plants Plant Communications 2022 3 100305 10.1016/j.xplc.2022.100305CrossRefGoogle ScholarPubMed
Fahs, S., Lujan, P., Kohn, M. Approaches to study phosphatases ACS Chemical Biology 2016 11 29442961 10.1021/acschembio.6b00570CrossRefGoogle ScholarPubMed
Fange, D., Lovmar, M., Pavlov, M. Y., Ehrenberg, M. Identification of enzyme inhibitory mechanisms from steady-state kinetics Biochimie 2011 93 16231629 10.1016/j.biochi.2011.05.031CrossRefGoogle ScholarPubMed
Garske, B., Ekardt, F. Economic policy instruments for sustainable phosphorus management: Taking into account climate and biodiversity targets Environmental Sciences Europe 2021 33 56 10.1186/s12302-021-00499-7CrossRefGoogle Scholar
Ghiaci, M., Aghaei, H., Soleimanian, S., Sedaghat, M. E. Enzyme immobilization: Part 2. Immobilization of alkaline phosphatase on Na-bentonite and modified bentonite Applied Clay Science 2009 43 308316 10.1016/j.clay.2008.09.011CrossRefGoogle Scholar
Gianfreda, L., Bollag, J. M. Effect of soils on the behavior of immobilized enzymes Soil Science Society of America Journal 1994 58 1672 10.2136/sssaj1994.03615995005800060014xCrossRefGoogle Scholar
Gianfreda, L., Rao, M. A. Potential of extra cellular enzymes in remediation of polluted soils: A review Enzyme and Microbial Technology 2004 35 339354 10.1016/j.enzmictec.2004.05.006CrossRefGoogle Scholar
Gianfreda, L., Rao, M. A., Sannino, F., Saccomandi, F., Violante, A. Enzymes in soil: Properties, behavior and potential applications Developments in Soil Science 2002 28B 301327 10.1016/S0166-2481(02)80027-7CrossRefGoogle Scholar
Gil-Sotres, F., Trasar-Cepeda, C., Leirós, M. C., Seoane, S. Different approaches to evaluating soil quality using biochemical properties Soil Biology and Biochemistry 2005 37 877887 10.1016/j.soilbio.2004.10.003CrossRefGoogle Scholar
Hassan, M. E., Yang, Q., Xiao, Z., Liu, L., Wang, N., Cui, X., Yang, L. Impact of immobilization technology in industrial and pharmaceutical applications 3 Biotech 2019 9 440 10.1007/s13205-019-1969-0CrossRefGoogle ScholarPubMed
Häussling, M., Marschner, H. Organic and inorganic soil phosphates and acid phosphatase activity in the rhizosphere of 80-year-old Norway spruce [Picea abies (L.) Karst.] trees Biology and Fertility of Soils 1989 8 128133 10.1007/BF00257756CrossRefGoogle Scholar
Hegeman, C. E., Grabau, E. A. A novel phytase with sequence similarity to purple acid phosphatases is expressed in cotyledons of germinating soybean seedlings Plant Physiology 2001 126 15981608 10.1104/pp.126.4.1598CrossRefGoogle ScholarPubMed
Hoehamer, C. F., Mazur, C. S., Wolfe, N. L. Purification and partial characterization of an acid phosphatase from Spirodela oligorrhiza and its affinity for selected organophosphate pesticides Journal of Agricultural and Food Chemistry 2005 53 9097 10.1021/jf040329uCrossRefGoogle ScholarPubMed
Hong, Y. K., Kim, J. W., Lee, S. P., Yang, J. E., Kim, S. C. Heavy metal remediation in soil with chemical amendments and its impact on activity of antioxidant enzymes in Lettuce (Lactuca sativa) and soil enzymes Applied Biological Chemistry 2020 63 42 10.1186/s13765-020-00526-wCrossRefGoogle Scholar
Huang, Q., Shindo, H. Effects of copper on the activity and kinetics of free and immobilized acid phosphatase Soil Biology and Biochemistry 2000 32 18851892 10.1016/S0038-0717(00)00162-0CrossRefGoogle Scholar
Huang, Q., Liang, W., Cai, P. Adsorption, desorption and activities of acid phosphatase on various colloidal particles from an Ultisol Colloids and Surfaces B: Biointerfaces 2005 45 209214 10.1016/j.colsurfb.2005.08.011CrossRefGoogle ScholarPubMed
Huang, Q., Zhu, J., Qiao, X., Cai, P., Rong, X., Liang, W., Chen, W. Conformation, activity and proteolytic stability of acid phosphatase on clay minerals and soil colloids from an Alfisol Colloids and Surfaces B: Biointerfaces 2009 74 279283 10.1016/j.colsurfb.2009.07.031CrossRefGoogle ScholarPubMed
Hughes, J. D., Simpson, G. H. Arylsulphatase–clay interactions. II The effect of kaolinite and montmorillonite on arylsulphatase activity Australian Journal of Soil Research 1978 16 3540 10.1071/SR9780035CrossRefGoogle Scholar
Hussain, S. I., Phillips, L. A., Hu, Y., Frey, S. K., Geuder, D. S., Edwards, M., Lapen, D. R., Ptacek, C. J., Blowes, D. W. Differences in phosphorus biogeochemistry and mediating microorganisms in the matrix and macropores of an agricultural clay loam soil Soil Biology and Biochemistry 2021 161 108365 10.1016/j.soilbio.2021.108365CrossRefGoogle Scholar
Ito, A., Wagai, R. Global distribution of clay-size minerals on land surface for biogeochemical and climatological studies Scientific Data 2017 4 111 10.1038/sdata.2017.103CrossRefGoogle ScholarPubMed
Jesionowski, T., Zdarta, J., Krajewska, B. Enzyme immobilization by adsorption: A review Adsorption 2014 20 801821 10.1007/s10450-014-9623-yCrossRefGoogle Scholar
Johnson, K. A., Goody, R. S. The original Michaelis constant: Translation of the 1913 Michaelis−Menten paper Biochemistry 2011 50 82648269 10.1021/bi201284uCrossRefGoogle Scholar
Kamat, S. S., Raushel, F. M. The enzymatic conversion of phosphonates to phosphate by bacteria Current Opinion in Chemical Biology 2013 17 589596 10.1016/j.cbpa.2013.06.006CrossRefGoogle ScholarPubMed
Kehler, A., Haygarth, P., Tamburini, F., Blackwell, M. Cycling of reduced phosphorus compounds in soil and potential impacts of climate change European Journal of Soil Science 2021 72 25172537 10.1111/ejss.13121CrossRefGoogle Scholar
Kelleher, B. P., Willeford, K. O., Simpson, A. J., Simpson, M. J., Stout, R., Rafferty, A., Kingery, W. L. Acid phosphatase interactions with organo-mineral complexes: Influence on catalytic activity Biogeochemistry 2004 71 285297 10.1023/B:BIOG.0000049348.53070.6fCrossRefGoogle Scholar
Kim, E. E., Wyckoff, H. W. Reaction mechanism of alkaline phosphatase based on crystal structures. Two-metal ion catalysis Journal of Molecular Biology 1991 218 449464 10.1016/0022-2836(91)90724-KCrossRefGoogle ScholarPubMed
Kim, Y. O., Kim, H. K., Bae, K. S., Yu, J. H., Oh, T. K. Purification and properties of a thermostable phytase from Bacillus sp. DS11 Enzyme and Microbial Technology 1998 22 27 10.1016/S0141-0229(97)00096-3CrossRefGoogle Scholar
Kome, G. K., Enang, R. K., Tabi, F. O., Yerima, BPK Influence of clay minerals on some soil fertility attributes: A review Open Journal of Soil Science 2019 9 155188 10.4236/ojss.2019.99010CrossRefGoogle Scholar
Kovács, J., Farics, É, Szabó, P., Sajó, I. Fe-Al phosphate microcrystals in pedogenic goethite pisoliths Minerals 2020 10 114 10.3390/min10040357CrossRefGoogle Scholar
Lehmann, J., Lehmann, J., Lilienfein, J., Rebel, K., do Carmo Lima, S., Wilcke, W. Subsoil retention of organic and inorganic nitrogen in a Brazilian savanna Oxisol Soil Use and Management 2004 20 163172 10.1079/SUM2004240CrossRefGoogle Scholar
Leprince, F., Quiquampoix, H. Extracellular enzyme activity in soil: Effect of pH and ionic strength on the interaction with montmorillonite of two acid phosphatases secreted by the ectomycorrhizal fungus Hebeloma cylindrosporum European Journal of Soil Science 1996 47 511522 10.1111/j.1365-2389.1996.tb01851.xCrossRefGoogle Scholar
Li, J., Xu, Y. Effects of clay combined with moisture management on Cd immobilization and fertility index of polluted rice field Ecotoxicology and Environmental Safety 2018 158 182186 10.1016/j.ecoenv.2018.04.031CrossRefGoogle ScholarPubMed
Liu, Q., Loganathan, P., Hedley, M. J., Skinner, M. F. The mobilisation and fate of soil and rock phosphate in the rhizosphere of ectomycorrhizal Pinus radiata seedlings in an Allophanic soil Plant and Soil 2004 264 219229 10.1023/B:PLSO.0000047758.77661.57CrossRefGoogle Scholar
Liu, P., Cai, Z., Chen, Z., Mo, X., Ding, X., Liang, C., Liu, G., Tian, J. A root-associated purple acid phosphatase, SgPAP23, mediates extracellular phytate-P utilization in Stylosanthes guianensis Plant Cell and Environment 2018 41 28212834 10.1111/pce.13412CrossRefGoogle ScholarPubMed
Maghraby, Y. R., El-Shabasy, R. M., Ibrahim, A. H., Azzazy, HME Enzyme immobilization technologies and industrial applications ACS Omega 2023 8 51845196 10.1021/acsomega.2c07560CrossRefGoogle ScholarPubMed
Makboul, H. E., Ottow, JCG Alkaline phosphatase activity and Michaelis constant in the presence of different clay minerals Soil Science 1979 128 129135 10.1097/00010694-197909000-00001CrossRefGoogle Scholar
Mamo, M., Taylor, R. W., Shuford, J. W. Ammonium fixation by soil and pure clay minerals Communications in Soil Science and Plant Analysis 1993 24 11151126 10.1080/00103629309368864CrossRefGoogle Scholar
Mao, L., Tang, D., Feng, H., Gao, Y., Zhou, P., Xu, L., Wang, L. Determining soil enzyme activities for the assessment of fungi and citric acid-assisted phytoextraction under cadmium and lead contamination Environmental Science and Pollution Research 2015 22 1986019869 10.1007/s11356-015-5220-1CrossRefGoogle ScholarPubMed
Margalef, O., Sardans, J., Fernández-Martínez, M., Molowny-Horas, R., Janssens, I. A., Ciais, P., Goll, D., Richter, A., Obersteiner, M., Asensio, D., Peñuelas, J. Global patterns of phosphatase activity in natural soils Scientific Reports 2017 7 1337 10.1038/s41598-017-01418-8CrossRefGoogle ScholarPubMed
Masaphy, S., Fahima, T., Levanon, D., Henis, Y., Mingelgrin, U. Parathion degradation by Xanthomonas sp. and Its crude enzyme extract in clay suspensions Journal of Environmental Quality 1996 25 12481255 10.2134/jeq1996.00472425002500060012xCrossRefGoogle Scholar
McBride, M. B. Dixon, B., Weed, S. B. Surface chemistry of soil minerals Minerals in soil environments 1989 Soil Science Society of America, Inc. 3588Google Scholar
McConnell, C. A., Kaye, J. P., Kemanian, A. R. Reviews and syntheses: Ironing out wrinkles in the soil phosphorus cycling paradigm Biogeosciences 2020 17 53095333 10.5194/bg-17-5309-2020CrossRefGoogle Scholar
Meng, D., Jiang, W., Li, J., Huang, L., Zhai, L., Zhang, L., Guan, Z., Cai, Y., Xiangru, L. An alkaline phosphatase from Bacillus amyloliquefaciens YP6 of new application in biodegradation of five broad-spectrum organophosphorus pesticides Journal of Environmental Science and Health, Part B 2019 54 336343 10.1080/03601234.2019.1571363CrossRefGoogle ScholarPubMed
Mousavi, S. M., Hashemi, S. A., Iman, MSM, Ravan, N., Gholami, A., Lai, C. W., Chiang, W. H., Omidifar, N., Yousefi, K., Behbudi, G. Recent advances in enzymes for the bioremediation of pollutants Biochemistry Research International 2021 2021 5599204 10.1155/2021/5599204CrossRefGoogle ScholarPubMed
Naidja, A., Huang, P. M., Bollag, J. M. Enzyme-clay interactions and their impact on transformations of natural and anthropogenic organic compounds in soil Journal of Environmental Quality 2000 29 677691 10.2134/jeq2000.00472425002900030002xCrossRefGoogle Scholar
Nannipieri, P., Giagnoni, L., Landi, L., Renella, G. Bünemann, E., Oberson, A., Frossard, E. Role of phosphatase enzymes in soil Phosphorus in action: Biological processes in soil phosphorus cycling 2011 26 Springer Berlin 215243 10.1007/978-3-642-15271-9_9CrossRefGoogle Scholar
Nasrabadi, G. R., Greiner, R., Yamchi, A., Nourzadeh Roshan, E. A novel purple acid phytase from an earthworm cast bacterium Journal of the Science of Food and Agriculture 2018 98 36673674 10.1002/jsfa.8845CrossRefGoogle Scholar
Norde, W. My voyage of discovery to proteins in flatland …and beyond Colloids and Surfaces B: Biointerfaces 2008 61 19 10.1016/j.colsurfb.2007.09.029CrossRefGoogle ScholarPubMed
Odom, I. E. Smectite clay minerals: Properties and uses Philosophical Transactions of the Royal Society of London 1984 311 391409 10.1098/rsta.1984.0036Google Scholar
Oh, B. C., Choi, W. C., Park, S., Kim, Y. O., Oh, T. K. Biochemical properties and substrate specificities of alkaline and histidine acid phytases Applied Microbiology and Biotechnology 2004 63 362372 10.1007/s00253-003-1345-0CrossRefGoogle ScholarPubMed
Olagoke, F. K., Kalbitz, K., Vogel, C. Control of soil extracellular enzyme activities by clay minerals — perspectives on microbial responses Soil Systems 2019 3 116 10.3390/soilsystems3040064CrossRefGoogle Scholar
Oliverio, A. M., Bissett, A., McGuire, K., Saltonstall, K., Turner, B. L., Fierer, N. The role of phosphorus limitation in shaping soil bacterial communities and their metabolic capabilities American Society for Microbiology 2020 11 e01718e1720Google ScholarPubMed
Ostanin, K., Harms, E. H., Stevis, P. E., Kuciel, R., Zhou, M., Van Ettene, R. L. Overexpression, site-directed mutagenesis, and mechanism of Escherichia coli acid phosphatase The Journal of Biological Chmistry 1992 267 2283022836 10.1016/S0021-9258(18)50022-3CrossRefGoogle ScholarPubMed
Paul, R., Datta, S. C., Bera, T., Math, M. K., Bhattacharyya, R., Dahuja, A. Interaction of phosphatase with soil nanoclays: Kinetics, thermodynamics and activities Geoderma 2022 409 115654 10.1016/j.geoderma.2021.115654CrossRefGoogle Scholar
Quiquampiox, H., Mousain, D. Turner, B. L., Frossard, E., Baldwin, D. S. Enzymatic hydrolysis of organic phosphorus Organic phosphorus in the environment 2005 CAB International 89112 10.1079/9780851998220.0089CrossRefGoogle Scholar
Quiquampoix, H., Abadie, J., Baron, M. H., Leprince, F., Matumoto-Pintro, P. T., Ratcliffe, R. G., Staunton, S. Horbett, T. A., Brash, J. L. Mechanisms and consequences of protein adsorption on soil mineral surfaces Proteins at interfaces II: Fundamentals and applications 1995 American Chemical Society 321333 10.1021/bk-1995-0602.ch023CrossRefGoogle Scholar
Rao, M. A., Violante, A., Gianfreda, L. Interaction of acid phosphatase with clays, organic molecules and organo-mineral complexes: Kinetics and stability Soil Biology and Biochemistry 2000 32 10071014 10.1016/S0038-0717(00)00010-9CrossRefGoogle Scholar
Reardon, P. N., Chacon, S. S., Walter, E. D., Bowden, M. E., Washton, N. M., Kleber, M. Abiotic Protein fragmentation by manganese oxide: Implications for a mechanism to supply soil biota with oligopeptides Environmental Science & Technology 2016 50 34863493 10.1021/acs.est.5b04622CrossRefGoogle ScholarPubMed
Redel, Y., Cartes, P., Velásquez, G., Poblete-Grant, P., Poblete-Grant, P., Bol, R., Mora, M. L. Assessment of phosphorus status influenced by Al and Fe compounds in volcanic grassland soils Journal of Soil Science and Plant Nutrition 2016 16 490506Google Scholar
Rejsek, K., Vranova, V., Pavelka, M., Formanek, P. Acid phosphomonoesterase (E.C. 3.1.3.2) location in soil Journal of Plant Nutrition and Soil Science 2012 175 196211 10.1002/jpln.201000139CrossRefGoogle Scholar
Richardson, A. E., Lynch, J. P., Ryan, P. R., Delhaize, E., Smith, F. A., Smith, S. E., Harvey, P. R., Ryan, M. H., Veneklaas, E. J., Lambers, H., Oberson, A., Culvenor, R. A., Simpson, R. J. Plant and microbial strategies to improve the phosphorus efficiency of agriculture Plant and Soil 2011 349 121156 10.1007/s11104-011-0950-4CrossRefGoogle Scholar
Rocha, JHT, Menegale, MLC, Rodrigues, M., de M Gonçalves, J. L., Pavinato, P. S., Foltran, E. C., Harrison, R., James, J. N. Impacts of timber harvest intensity and P fertilizer application on soil P fractions Forest Ecology and Management 2019 437 295303 10.1016/j.foreco.2019.01.051CrossRefGoogle Scholar
Rodríguez, R. F. Structure-function studies of a purple acid phytase 2018 University of East AngliaGoogle Scholar
Rodríguez, H., Fraga, R., Gonzalez, T., Bashan, Y. Genetics of phosphate solubilization and its potential applications for improving plant growth-promoting bacteria Plant and Soil 2006 287 1521 10.1007/s11104-006-9056-9CrossRefGoogle Scholar
Rosas, A., Luz Mora, M., Jara, A. A., López, R., Rao, M. A., Gianfreda, L. Catalytic behaviour of acid phosphatase immobilized on natural supports in the presence of manganese or molybdenum Geoderma 2008 145 7783 10.1016/j.geoderma.2008.02.008CrossRefGoogle Scholar
Roth, C. M., Lenhoff, A. M. Electrostatic and van der Waals contributions to protein adsorption: Comparison of theory and experiment Langmuir 1995 11 35003509 10.1021/la00009a036CrossRefGoogle Scholar
Ruggiero, P., Sarkar, J. M., Bollag, J. M. Detoxification of 2,4-dichlorophenol by a laccase immobilized on soil or clay Soil Science 1989 147 361370 10.1097/00010694-198905000-00007CrossRefGoogle Scholar
Sarkar, J. M., Leonowicz, A. Immobilization of enzymes on clays and soils Soil Biology and Biochemistry 1989 21 223230 10.1016/0038-0717(89)90098-9CrossRefGoogle Scholar
Schaap, K. J., Fuchslueger, L., Hoosbeek, M. R., Hofhansl, F., Martins, N. P., Valverde-Barrantes, O. J., Hartley, I. P., Lugli, L. F., Quesada, C. A. Litter inputs and phosphatase activity affect the temporal variability of organic phosphorus in a tropical forest soil in the Central Amazon Plant and Soil 2021 469 423441 10.1007/s11104-021-05146-xCrossRefGoogle Scholar
Schmidt, M. P., Martínez, C. E. Kinetic and conformational insights of protein adsorption onto montmorillonite revealed using in situ ATR-FTIR/2D-COS Langmuir 2016 32 77197729 10.1021/acs.langmuir.6b00786CrossRefGoogle ScholarPubMed
Sedaghat, M. E., Ghiaci, M., Aghaei, H., Soleimanian-zad, S. Enzyme immobilization. Part 4 . Immobilization of alkaline phosphatase on Na-sepiolite and modi fi ed sepiolite Applied Clay Science 2009 46 131135 10.1016/j.clay.2009.07.021CrossRefGoogle Scholar
Sepelyak, R. J., Feldkamp, J. R., Moody, T. E., White, J. L., Hem, S. L. Adsorption of pepsin by aluminum hydroxide I: Adsorption mechanism Journal of Pharmaceutical Sciences 1984 73 15141517 10.1002/jps.2600731104Google ScholarPubMed
Shin, S., Ha, N. C., Oh, B. C., Oh, T. K., Oh, B. H. Enzyme mechanism and catalytic property of β propeller phytase Structure 2001 9 851858 10.1016/S0969-2126(01)00637-2CrossRefGoogle ScholarPubMed
Shindo, H., Watanabe, D., Onaga, T., Urakawa, M., Nakahara, O., Huang, Q. Adsorption, activity, and kinetics of acid phosphatase as influenced by selected oxides and clay minerals Soil Science and Plant Nutrition 2002 48 763767 10.1080/00380768.2002.10409268CrossRefGoogle Scholar
Shirvani, M., Khalili, B., Kalbasi, M., Shariatmadari, H., Nourbakhsh, F. Adsorption of alkaline phosphatases on palygorskite and sepiolite: A tradeoff between enzyme protection and inhibition Clays and Clay Minerals 2020 68 287295 10.1007/s42860-020-00066-wCrossRefGoogle Scholar
Sims, J. T., Sharpley, A. N. Phosphorus: Agriculture and the environment 2005 American Society of Agronomy 10.2134/agronmonogr46CrossRefGoogle Scholar
Somu, P., Narayanasamy, S., Gomez, L. A., Rajendran, S., Lee, Y. R., Balakrishnan, D. Immobilization of enzymes for bioremediation: A future remedial and mitigating strategy Environmental Research 2022 212 113411 10.1016/j.envres.2022.113411CrossRefGoogle ScholarPubMed
Sparks, D. L. (2003). Sorption phenomena on soils. In Sparks, D. L. (ed.), Environmental soil chemistry (pp. 133186). San Diego, CA: Academic Press.CrossRefGoogle Scholar
Sposito, G. The surface chemistry of soils 1984 Oxford University Press, Inc.Google Scholar
Staunton, S., Quiquampoix, H. Adsorption and conformation of bovine serum albumin on montmorillonite: Modification of the balance between hydrophobic and electrostatic interactions by protein methylation and pH variation Journal of Colloid and Interface Science 1994 166 8994 10.1006/jcis.1994.1274CrossRefGoogle Scholar
Stec, B., Holtz, K. M., Kantrowitz, E. R. A revised mechanism for the alkaline phosphatase reaction involving three metal ions Journal of Molecular Biology 2000 299 13031311 10.1006/jmbi.2000.3799CrossRefGoogle ScholarPubMed
Stosiek, N., Talma, M., Klimek-ochab, M. Carbon-Phosphorus Lyase — the state of the art Applied Biochemistry and Biotechnology 2020 190 15251552 10.1007/s12010-019-03161-4CrossRefGoogle ScholarPubMed
Stotzky, G. Huang, P. M., Schnitzer, M. Influence of soil mineral colloids on metabolic processes, growth, adhesion, and ecology of microbes and viruses Interactions of soil minerals with natural organics and microbes 1986 Soil Science Society of America 305428Google Scholar
Sumner, M. E. Handbook of soil sciences: Properties and processes 2000 CRC PressGoogle Scholar
Tan, X., He, Y., Wang, Z., Li, C., Kong, L., Tian, H., Shen, W., Megharaj, M., He, W. Soil mineral alters the effect of Cd on the alkaline phosphatase activity Ecotoxicology and Environmental Safety 2018 161 7884 10.1016/j.ecoenv.2018.05.069CrossRefGoogle ScholarPubMed
Tang, X., Shen, Y., Suib, S. L., Coughlin, R. W., Vinopal, R. Immobilization of alkaline phosphatase on protamine-intercalated bentonite Microporous Materials 1993 2 6571 10.1016/0927-6513(93)80063-ZCrossRefGoogle Scholar
Tapia-Torres, Y., Rodríguez-torres, M. D., Elser, J. J., Islas, A., Souza, V., García-Oliva, F., Olmedo-Álvarez, G. How to live with phosphorus scarcity in soil and sediment: Lessons from bacteria Applied and Environmental Microbiology 2016 82 46524662 10.1128/AEM.00160-16CrossRefGoogle ScholarPubMed
Tarafdar, J. C., Jungk, A. Phosphatase activity in the rhizosphere and its relation to the depletion of soil organic phosphorus Biology and Fertility of Soils 1987 3 199204 10.1007/BF00640630CrossRefGoogle Scholar
Tarafdar, J. C., Yadav, R. S., Meena, S. C. Comparative efficiency of acid phosphatase originated from plant and fungal sources Journal of Plant Nutrition and Soil Science 2001 164 279282 10.1002/1522-2624(200106)164:3<279::AID-JPLN279>3.0.CO;2-L3.0.CO;2-L>CrossRefGoogle Scholar
Ternan, N. G., Mc Grath, J. W., Mc Mullan, G., Quinn, J. P. Review: Organophosphonates: Occurrence, synthesis and biodegradation by microorganisms World Journal of Microbiology and Biotechnology 1998 14 635647 10.1023/A:1008848401799CrossRefGoogle Scholar
Theng, B. K. G. (2012) Proteins and enzymes. In Formation and properties of clay-polymer complexes (pp. 245318). 2nd edition. Elsevier B.V.CrossRefGoogle Scholar
Tian, H., Zhao, Y., Megharaj, M., He, W. Arsenate inhibition on kinetic characteristics of alkaline phosphatase as influenced by pH Ecological Indicators 2018 85 11011106 10.1016/j.ecolind.2017.11.041CrossRefGoogle Scholar
Tiessen, H., Stewart, JWB, Cole, C. V. Pathways of phosphorus transformations in soils of differing pedogenesis Soil Science Society of America Journal 1984 48 853858 10.2136/sssaj1984.03615995004800040031xCrossRefGoogle Scholar
Tietjen, T., Wetzel, R. G. Extracellular enzyme-clay mineral complexes: Enzyme adsorption, alteration of enzyme activity, and protection from photodegradation Aquatic Ecology 2003 37 331339 10.1023/B:AECO.0000007044.52801.6bCrossRefGoogle Scholar
Turner, B. L., Condron, L. M. Pedogenesis, nutrient dynamics, and ecosystem development: The legacy of T.W. Walker and J.K. Syers Plant and Soil 2013 367 110 10.1007/s11104-013-1750-9CrossRefGoogle Scholar
Turner, B. L., Richardson, A. E., Mullaney, E. J. Inositol phosphates: Linking agriculture and the environment 2007 CAB International 10.1079/9781845931520.0000CrossRefGoogle Scholar
Turner, B. L., Lambers, H., Condron, L. M., Cramer, M. D., Leake, J. R., Richardson, A. E., Smith, S. E. Soil microbial biomass and the fate of phosphorus during long-term ecosystem development Plant and Soil 2013 367 225234 10.1007/s11104-012-1493-zCrossRefGoogle Scholar
Ullah, AHJ, Cummins, B. J., Dischinger, H. C. Cyclohexanedione modification of arginine at the active site of Aspergillusficuum phytase Biochemical and Biophysical Research Communications 1991 178 4553 10.1016/0006-291X(91)91777-ACrossRefGoogle ScholarPubMed
Ustunol, I. B., Coward, E. K., Quirk, E., Grassian, V. H. Interaction of beta-lactoglobulin and bovine serum albumin with iron oxide (α-Fe2O3) nanoparticles in the presence and absence of pre-adsorbed phosphate Environmental Science: Nano 2021 8 28112823Google Scholar
Vance, C. P., Uhde-Stone, C., Allan, D. L. Phosphorus acquisition and use: Critical adaptations by plants for securing a nonrenewable resource New Phytologist 2003 157 423447 10.1046/j.1469-8137.2003.00695.xCrossRefGoogle ScholarPubMed
Vincent, J. B., Crowder, M. W., Averill, B. A. Hydrolysis of phosphate monoesters: A biological problem with multiple chemical solutions Trends in Biochemical Sciences 1992 17 105110 10.1016/0968-0004(92)90246-6CrossRefGoogle ScholarPubMed
Wackett, L. P., Shames, S. L., Venditti, C. P., Walsh, C. T. Bacterial carbon-phosphorus lyase: Products, rates, and regulation of phosphonic and phosphinic acid metabolism Journal of Bacteriology 1987 169 710717 10.1128/jb.169.2.710-717.1987CrossRefGoogle ScholarPubMed
Wang, P. Nanoscale biocatalyst systems Current Opinion in Biotechnology 2006 17 574579 10.1016/j.copbio.2006.10.009CrossRefGoogle ScholarPubMed
Wang, S., Liang, X., Chen, Y., Luo, Q., Liang, W., Li, S., Huang, C., Li, Z., Wan, L., Li, W., Shao, X. Phosphorus loss potential and phosphatase activity under phosphorus fertilization in long-term paddy wetland agroecosystems Soil Science Society of America Journal 2012 76 161167 10.2136/sssaj2011.0078CrossRefGoogle Scholar
Wang, Z. Q., Li, Y. B., Tan, X. P., He, W. X., Xie, W., Megharaj, M., Wei, G. H. Effect of arsenate contamination on free, immobilized and soil alkaline phosphatases: Activity, kinetics and thermodynamics European Journal of Soil Science 2017 68 126135 10.1111/ejss.12397CrossRefGoogle Scholar
Wang, Y., Xie, H., Zhang, L., Wang, C. Restrictions of weathering on phosphorus migration and enrichment of the lower Cambrian phosphorus-bearing rock series in eastern Guizhou ACS Omega 2023 8 98159831 10.1021/acsomega.2c06160CrossRefGoogle ScholarPubMed
Wei, S., Dai, Y., Duan, Q., Liu, B., Yuan, H. A global soil data set for earth system modeling Journal of Advances in Modeling Earth Systems 2014 6 249263 10.1002/2013MS000293Google Scholar
White, A. K., Metcalf, W. W. Microbial metabolism of reduced phosphorus compounds Annual Review of Microbiology 2007 61 379400 10.1146/annurev.micro.61.080706.093357CrossRefGoogle ScholarPubMed
Yang, Z. X., Liu, S. Q., Zheng, D. W., Feng, S. D. Effects of cadium, zinc and lead on soil enzyme activities Journal of Environmental Sciences 2006 18 11351141 10.1016/S1001-0742(06)60051-XCrossRefGoogle ScholarPubMed
Yanke, L. J., Selinger, L. B., Cheng, K. J. Phytase activity of Selenomonas ruminantium: A preliminary characterization Letters in Applied Microbiology 1999 29 2025 10.1046/j.1365-2672.1999.00568.xCrossRefGoogle Scholar
Yu, S., He, Z. L., Stoffella, P. J., Calvert, D. V., Yang, X. E., Banks, D. J., Baligar, V. C. Surface runoff phosphorus (P) loss in relation to phosphatase activity and soil P fractions in Florida sandy soils under citrus production Soil Biology and Biochemistry 2006 38 619628 10.1016/j.soilbio.2005.02.040CrossRefGoogle Scholar
Zhang, W., Tang, X., Feng, X., Wang, E., Li, H., Shen, J., Zhang, F. Management strategies to optimize soil phosphorus utilization and alleviate environmental risk in China Journal of Environmental Quality 2019 48 11671175 10.2134/jeq2019.02.0054CrossRefGoogle ScholarPubMed
Zhu, J., Huang, Q., Pigna, M., Violante, A. Immobilization of acid phosphatase on uncalcined and calcined Mg/Al-CO3 layered double hydroxides Colloids and Surfaces B: Biointerfaces 2010 77 166173 10.1016/j.colsurfb.2010.01.020CrossRefGoogle Scholar
Zhu, Y., Wu, F., Feng, W., Liu, S., Giesy, J. P. Interaction of alkaline phosphatase with minerals and sediments: Activities, kinetics and hydrolysis of organic phosphorus Colloids and Surfaces A: Physicochemical and Engineering Aspects 2016 495 4653 10.1016/j.colsurfa.2016.01.056CrossRefGoogle Scholar
Zimmerman, A. R., Ahn, M. Y. Shukla, G., Varma, A. Organo-mineral-enzyme interaction and soil enzyme activity Soil enzymology 2011 Springer 271292Google Scholar
Zou, X., Binkley, D., Caldwell, B. A. Effects of dinitrogen-fixing trees on phosphorus biogeochemical cycling in contrasting forests Soil Science Society of America Journal 1995 59 14521458 10.2136/sssaj1995.03615995005900050035xCrossRefGoogle Scholar
Figure 0

Fig. 1 Classification of enzymes that can release phosphate from organic P compounds, based on the enzyme database from the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology (NC-IUBMB). Alkaline (EC 3.1.3.1) and acid (EC 3.1.3.2) phosphatases are the main focus of the subsequent sections

Figure 1

Fig. 2 Phosphomonoester hydrolysis reaction catalyzed by alkaline and acid phosphatases

Figure 2

Fig. 3 a The enzyme is in its active conformation and the substrate enters the active site. b Substrate binds to the active site, forming an enzyme–substrate complex. c Conformational change of the enzyme alters the shape of the active site, preventing substrate binding

Figure 3

Table 1 Experimental conditions and changes in phosphatase properties in phosphatase-clay mineral sorption studies

Figure 4

Fig. 4 Phosphatase activities in the terrestrial environment. The role of clays and clay minerals as phosphatase sinks is highlighted. The interaction could suppress P mineralization and enhance enzyme stability