Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-17T14:45:40.539Z Has data issue: false hasContentIssue false

Clay-Polymer Interactions: Summary and Perspectives

Published online by Cambridge University Press:  02 April 2024

B. K. G. Theng*
Affiliation:
Soil Bureau, Department of Scientific and Industrial Research, Lower Hutt, New Zealand

Abstract

The adsorption of uncharged polymers by clays is largely “entropy-driven.” Polymer conformation changes from a random coil in solution to an extended form at the surface in which adsorbed polymer segments or trains alternate with loops and tails extending away from the surface. Although the net interaction energy, ε, per segment-surface contact is small (~1 kT unit), the total energy of adsorption is large because the fraction of train segments, p, is commonly between 0.3 and 0.5. The adsorption isotherms are typically of the high-affinity type, and there is an apparent lack of desorption on dilution. Positively charged polymers (polycations) are adsorbed largely through electrostatic interactions between the cationic groups of the polymer and the negatively charged sites at the clay surface. Here ε ≫ 1 kT unit and p > 0.7, leading to an almost complete collapse of the polymer chain onto the surface. Indeed, beyond a given level of adsorption charge reversal can occur in that the clay-polycation system effectively behaves as an anion exchanger. Little adsorption occurs with negatively charged polymers (polyanions) due to initial charge repulsion between the polymer and the clay surface. Acid pH, a high ionic strength, and the presence of polyvalent cations in the system enhance and promote polyanion adsorption. Uncharged polymers and polycations can enter the interlayer space of expanding 2:1 type layer silicates but polyanions generally fail to intercalate.

The interactions of clays with biopolymers, such as proteins, nucleic acids, and polysaccharides, can be rationalized in similar terms. When intercalation occurs, the interlayer biopolymer is further stabilized against microbial (enzymatic) degradation giving rise to practical applications of clay-polymer complexes as flocculants and soil conditioners. Polyanions are effective as flocculants because of their large “grappling distance,” whereas uncharged polymers are better suited as soil conditioners because they can spread over adjacent clay/soil particle surfaces like a coat of paint.

Резюме

Резюме

Движущей силой адсорбции незаряженных полимеров глинами является, главным образом, энтропия. Конформация полимера изменяется от случайной спирали в растворе до вытянутой формы на поверхности, в которой сегменты адсорбированных полимеров либо их последовательные ряды чередуются с изогнутостями и концевыми частями, простирающимися от поверхности. Хотя чистая энергия взаимодействия, е, на контакт сегмент, поверхность небольшая (~1 кТ), общая энергия адсорбции большая вследствие того, что фракция серийных сегментов, р, обычно находится в пределах между 0,3 и 0,5. Адсорбционные изотермы обычно большого родства и имеется также кажущееся отсутствие десорбции при разбавленнии. Положительно заряженные полимеры (поликатионы) адсорбируются, главным образом, посредством электростатических взаимодействий между катионными группами полимера и отрицательно заряженными местами на поверхности глины. При этом е *** 1 кТ и р >0,7, что приводит к почти полному разрушению полимерных цепочек на поверхности. Действительно, вне данного уровна адсорбции может выступить перемена знака заряда с тем, что система глина—поликатион эффективно ведет себя как обменник анионов. Вследствие начального отталкивания заряда между полимером и поверхностью глины появляется небольшая адсорбция с отрицательно заряженными полимерами (полианионами). Кислотность рН, большая ионная сила и присутствие многовалентных катионов в системе благоприятствуют и ускоряют адсорбцию полианионов. Незаряженные полимеры и поликатионы могут входить в межслойные пространства расширяющихся слоевых силикатов типа 2:1, только полианиоиы, в основном, отказываются переслаиваться.

Взаимодействия глин с биополимерами, такими как белки, нуклеиновые кислоты и полисака-риды могут быть рационализированы подобным образом. Когда пояавляется прослойка, промежуточный биополимер стабилизируется в дальнейшем от микробного (ензиматического) распада приводя к практическому применению глинополимерных комплексов в качестве флоккулянтов и кондиционеров почвы. Полианионы эффективны как флоккулянты из-за их большого „схватывающего расстояния”, в то время как незаряженные полимеры лучше удовлетворяют требованиям в качестве кондиционеров почвы, так как они могут растекаться по прилегающих частицах глины/ почвы подобно растеканию краски. [Е.С.]

Resümee

Resümee

Die Adsorption von ungeladenen Polymeren durch Tone hängt hauptsächlich von der Entropie ab. Die Polymerkonformation ändert sich von einer Spirale in der Lösung zu einer gestreckten Form an der Oberfläche, an der adsorbierte Polymer-Segmente oder Züge mit Polymer-Schleifen und -Enden abwechseln, die von der Oberfläche wegstehen. Obwohl die Vernetzungsenergie, ε, pro Segmentoberflächenkontakt gering ist (~1 kT), ist die gesamte Adsorptionsenergie groß, da die Spaltung der Polymersegmente, p, meistens zwischen 0,3 und 0,5 ist. Die Adsorptionsisotherme sind typisch vom Hochaffinitätstyp, wobei auch bei der Verdünnung eine Desorption offensichtlich fehlt. Positiv geladene Polymere (Polyka-tionen) werden hauptsächlich durch elektrostatische Wechselwirkungen zwischen den kationischen Gruppen des Polymers und den negativ geladenen Stellen auf der Tonoberfläche adsorbiert. Dabei ist ε ≫ 1 kT und p > 0,7, was zu einem nahezu vollständigen Zusammenbrechen der Polymerketten auf der Oberfläche führt. Tatsächlich kann über einem bestimmten Grad der Adsorption eine Ladungsumkehr auftreten, wobei das Ton-Polykation-System wie ein Anionenaustauscher wirkt. Eine geringe Adsorption tritt bei negativ geladenen Polymeren (Polyanionen) auf, die auf dem anfänglichen Ladungsabstoß zwischen dem Polymer und der Tonoberfläche beruht. Ein saurer pH, eine groß Ionenstärke und die Anwesentheit von polyvalenten Kationen im System erhöhen und fördern die Adsorption von Polyanionen. Ungeladene Polymere und Polykationen können in den Zwischenraum von quellfähigen 2:1 Schichtsilikaten gehen, Polyanionen können dagegen im allgemeinen nicht eingebaut werden.

Die Wechselwirkungen von Tonen mit Biopolymeren, wie Proteinen, Nukleinsäuren und Polysaccharide, können auf ähnlichem Weg verwirklicht werden. Wenn ein Einbau stattfindet, ist das Zwischenschichtpolymer gegen mikrobiologischen (enzymischen) Abbau geschützt. Dies führt dazu, daß man Ton-Polymerkomplexe als Flockungsmittel und Bodenstabilisierungsmittel verwenden kann. Polykationen wirken wegen ihres großen “grappling distance” als Flockungsmittel, während ungeladene Polymere besser als Bodenstabilisierungsmittel geeignet sind, da sie sich wie eine Farbhaut über aneinander grenzende Ton/ Boden-Teilchen breiten können. [U.W.]

Résumé

Résumé

L'adsorption de corps polymères sans charge par des argiles est pour la plupart “conduite par entropie.” La conformation de corps polymères change d'une bobine au hasard en solution en une forme étendue à la surface dans laquelle des segments ou des trains de corps polymères adsorbés alternent avec des boucles et des queues s’éloignant de la surface. Malgré que l’énergie nette d'interaction, ε, par contact segment-surface, est petite (~1 kT) l’énergie totale d'adsorption est grande parce que la fraction des segments de train, p, est souvent entre 0,3 et 0,5. Les isothermes d'adsorption sont typiquement de la sorte à haute affinité et il y a un manque apparent de désorption lors de la dilution. Les corps polymères chargés positivement (polycations) sont pour la plupart adsorbés par interactions électrostatiques entre les groupes cationiques du corps polymère et les sites chargés négativement à la surface de l'argile. Dans ce cas, ε ≫ 1 kT et p > 0,7, menant à un effondrement presque complet de la chaîne polymère sur la surface. En effet, au delà d'un niveau donné d'adsorption, un renversement de charge peut se produire en le fait que le système argile-polycation se conduit effectivement comme un échangeur d'anions. Peu d'adsorption se produit avec les corps polymères chargés négativement (polyanions) à cause d'une répulsion de charge initiale entre le corps polymère et la surface argileuse. Un pH acide, une force ionique élevée et la présence de cations polyvalents dans le système accroit et promeut l'adsorption de polyanions. Des corps polymères et des polycations peuvent entrer dans l'espace interfeuillet des silicates en expansion à couches du type 2:1, mais les polyanions ne s'intercalent généralement pas.

Les interactions des argiles avec des corps biopolymères tels que des protéines, des acides nucléiques et des polysaccharides peuvent être rationalisées en des termes semblables. Lorsque l'intercalation se produit, le corps polymère interfeuillet est stabilisé d'avantage face à la dégradation microbiale (enzy-matique), menant à des applications pratiques de complexes argile-polymère en tant que flocculants et de conditionneurs de sols. Les polyanions sont de bons flocculants à cause de leur grande “distance étreig-nante,” tandis que les corps polymères sans charge sont de meilleurs conditionneurs de sols parcequ'ils peuvent s’étendre par dessus des particules argile/sol adjacentes comme une couche de peinture. [D.J.]

Type
Research Article
Copyright
Copyright © 1982, The Clay Minerals Society

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Almon, W. R., Johns, W. D. and Bailey, S. W., 1976 Petroleum forming reactions: clay catalyzed fatty acid decarboxylation Proc. Int. Clay Conf, Mexico City, 1975 Wilmette, Illinois Applied Publishing 399409.Google Scholar
de la Aragon Cruz, F. and Viton Barbolla, C., 1979 Formation of amino acids from hydrolysis of potassium cyanide in the presence of montmorillonite or graphite oxide An. Quim. 75 820826.Google Scholar
Black, A. P., Birkner, F. B. and Morgan, J. J., 1966 The effect of polymer adsorption on the electrokinetic stability of dilute clay suspensions J. Colloid Interface Sci. 21 626648.CrossRefGoogle Scholar
Cady, S. S. and Pinnavaia, T. J., 1978 Porphyrin intercalation in mica-type silicates Inorg. Chem. 17 15011507.CrossRefGoogle Scholar
Clapp, C. E. and Emerson, W. W., 1972 Reactions between Ca-montmorillonite and polysaccharides Soil Sci. 114 210216.CrossRefGoogle Scholar
Cohen Stuart, M. A., Scheutjens, J. M. H. M. and Fleer, G. J., 1980 Polydispersity effects and the interpretation of polymer adsorption isotherms J. Polymer Sci., Polymer Physics Ed. 18 559573.CrossRefGoogle Scholar
Czarnecka, E. and Gillott, J. E., 1980 Formation and characterization of clay complexes with bitumen from Athabasca oil sand Clays & Clay Minerals 28 197203.CrossRefGoogle Scholar
De Boodt, M. and Gabriels, D. (eds.) (1976) Third International Symposium on Soil Conditioning. Mededelingen Fakulteit Landbouwwetenschappen Rijksuniversiteit Gent 41, 464 pp.Google Scholar
De Leeuw, J. W., Correia, V. A. and Schenck, P. A., 1973 On the decomposition of phytol under simulated geological conditions and in the top-layer of natural sediments Advan. Org. Geochem., Proc. 6th Int. Congr. 9931004.Google Scholar
Doner, H. E. and Mortland, M. M., 1969 Benzene complexes with Cu(II)-montmorillonite Science 166 14061407.CrossRefGoogle Scholar
Estermann, E. F., Peterson, G. H. and McLaren, A. D., 1959 Digestion of clay-protein, lignin-protein, and silicaprotein complexes by enzymes and bacteria Soil Sci. Soc. Amer. Proc. 23 3136.CrossRefGoogle Scholar
Foscolos, A. E., Powell, T. G. and Gunther, P. R., 1976 The use of clay minerals and inorganic and organic geochemical indicators for evaluating the degree of diagenesis and oil generating potential of shales Geochim. Cosmochim. Acta 40 953966.CrossRefGoogle Scholar
Fripiat, J. J. and Poncelet, G., 1973 Clays and prebiotic synthesis Advan. Org. Geochem., Proc. 6th Internat. Congr. 875881.Google Scholar
Garvey, M L T Th F and Vincent, B., 1976 A comparison of the adsorbed layer thickness obtained by several techniques of various molecular weight fractions of poly(vinyl alcohol) on aqueous polystyrene latex particles J. Colloid Interface Sci. 55 440453.CrossRefGoogle Scholar
Greaves, M. P. and Wilson, M. J., 1969 The adsorption of nucleic acids by montmorillonite Soil Biol. Biochem. 1 317323.CrossRefGoogle Scholar
Greenland, D. J., 1963 Adsorption of polyvinyl alcohols by montmorillonite J. Colloid Sci. 18 647664.CrossRefGoogle Scholar
Hawthorne, D. G., Hodgkin, J. H., Loft, B. C. and Solomon, D.H., 1974 Polymerization of vinyl monomers on mineral surfaces; a novel method of preparing reinforcing fillers J. Macromol. Sci. Chem. A. 8 649657.CrossRefGoogle Scholar
Helmer, B. M., Prescott, P. I. and Sennett, P., 1976 Surface-modified kaolin in plastic 31st Ann. Tech. Conf. Reinforced Plastics/Composites Inst., Soc. Plastics Ind. Section 8-G 14.Google Scholar
Hepper, C. M. and Walker, N., 1975 Extracellular polysaccharides of soil bacteria Soil Microbiology London Butterworths 93110.Google Scholar
Hesslink, F. Th., 1977 On the theory of polyelectrolyte adsorption. The effect on adsorption behavior of the electrostatic contribution to the adsorption free energy J. Colloid Interface Sci. 60 448466.Google Scholar
Ibanez, J. D., Kimball, A. P. and Oro, J., 1971 Possible prebiotic condensation of mononucleotides by cyanamide Science 173 444446.CrossRefGoogle ScholarPubMed
Jacks, G. V., 1963 The biological nature of soil productivity Soils Fertilizers 26 147150.Google Scholar
Jordan, J. W., Swineford, A. and Franks, P. C., 1963 Organophilic clay-base thickeners Clays and Clay Minerals, Proc. 10th Natl. Conf, Austin, Texas, 1961 New York Pergamon Press 299308.Google Scholar
Kaas, R. L. and Gardlund, Z. G. (1976) Surface modification of fillers by radiation polymerization: Res. Publ. General Motors Res. 2304, 15 pp.Google Scholar
Kobayashi, Y. and Aomine, S., 1967 Mechanism of inhibitory effect of allophane and montmorillonite on some enzymes Soil Sci. Plant Nutr. 13 189194.CrossRefGoogle Scholar
Lahav, N. and Chang, S., 1976 The possible role of solid surface area in condensation reactions during chemical evolution: re-evaluation J. Mol. Evol. 8 357380.CrossRefGoogle Scholar
Libby, P. W., Iannicelli, J. and McGill, C. R., 1976 Elastomer reinforcement with amino silane grafted kaolin Amer. Chem. Soc, Div. Rubber Chem., Spring Meeting .Google Scholar
Makboul, H. E. and Ottow, J. C. G., 1979 Michaelis constant (Km) of acid phosphatase as affected by montmorillonite, illite, and kaolinite clay minerals Microbial Ecol. 5 207213.CrossRefGoogle ScholarPubMed
Malina, J. F. Jr. and Sagik, B. P., 1974 Virus Survival in Water and Wastewater Systems .Google Scholar
McLaren, A. D., Peterson, G. H. and Barshad, I., 1958 The adsorption and reactions of enzymes and proteins on clay minerals. IV. Kaolinite and montmorillonite Soil Sci. Soc. Amer. Proc. 22 239244.CrossRefGoogle Scholar
Morrissey, B. W. and Stromberg, R. R., 1974 The confirmation of adsorbed blood proteins by infrared bound fraction measurements J. Colloid Interface Sci. 46 152164.CrossRefGoogle Scholar
Mortensen, J. L. and Ada, S., 1962 Adsorption of hydrolysed polyacry-lonitrile on kaolinite Clays and Clay Minerals, Proc. 9th Natl. Conf, West Lafayette, Indiana, 1960 New York Pergamon Press 530545.Google Scholar
Mortland, M. M., 1970 Clay-organic complexes and interactions Advan. Agron. 22 75117.CrossRefGoogle Scholar
Mortland, M. M. and Halloran, L. J., 1976 Polymerization of aromatic molecules on smectite Soil Sci. Soc. Amer. J. 40 367370.CrossRefGoogle Scholar
Nahin, P. G., Ada, S. and Franks, P. C., 1963 Perspectives in applied organo-clay chemistry Clays and Clay Minerals, Proc. 10th Natl. Conf., Austin, Texas, 1961 New York Pergamon Press 257271.Google Scholar
Olness, A. and Clapp, C. E., 1975 Influence of polysaccharide structure on dextran adsorption by montmorillonite Soil Biol. Biochem. 7 113118.CrossRefGoogle Scholar
Paecht-Horowitz, M., 1978 The influence of various cations on the catalytic properties of clays J. Mol. Evol. 11 101107.CrossRefGoogle ScholarPubMed
Parfitt, R. L. and Greenland, D. J., 1970 The adsorption of poly(ethylene glycols) on clay minerals Clay Miner. 8 305315.CrossRefGoogle Scholar
Parfitt, R. L. and Greenland, D. J., 1970 Adsorption of polysaccharides by montmorillonite Soil Sci. Soc. Amer. Proc. 34 862866.CrossRefGoogle Scholar
Quirk, J. P. and Williams, B. G., 1974 Disposition of organic materials in relation to stable aggregation Trans. 10th Int. Congr. Soil Sci., Moscow 1 165173.Google Scholar
Ramirez-Martinez, J. R. and McLaren, A. D., 1966 Some factors influencing the determination of phosphatase activity in native soils and in soils sterilized by irradiation Enzymologia 31 2338.Google ScholarPubMed
Roberts, K., Kowalewska, J. and Friberg, S., 1974 The influence of interactions betwen hydrolyzed aluminum ions and Polyacrylamides on the sedimentation of kaolin suspensions J. Colloid Interface Sci. 48 361367.CrossRefGoogle Scholar
Ruehrwein, R. A. and Ward, D. W., 1952 Mechanism of clay aggregation by polyelectrolytes Soil Sci. 73 485492.CrossRefGoogle Scholar
Schamp, N. and Huylebroeck, J., 1973 Adsorption of polymers on clays J. Polymer Sci., Symposium 42 553562.CrossRefGoogle Scholar
Schloesing, Th., 1874 Détermination de l’argile dans la terre arable C. R. Acad. Sci., Paris 78 12761279.Google Scholar
Schnitzer, M. and Kodama, H., 1966 Montmorillonite: effect of pH on its adsorption of a soil humic compound Science 153 7071.CrossRefGoogle ScholarPubMed
Shimoyama, A., Blair, N. and Ponnamperuma, C., 1977 Synthesis of amino acids under primitive Earth conditions in the presence of clay Origin Life, Proc. 2nd Meet. Internat. Soc. Study Origin Life 9599.Google Scholar
Sieskind, O., Joly, G. and Albrecht, P., 1979 Simulation of the geochemical transformations of sterols: superacid effect of clay minerals Geochim. Cosmochim. Acta 43 16751679.CrossRefGoogle Scholar
Sohn, J. R. and Ozaki, A., 1980 Acidity of nickel silicate and its bearing on the catalytic activity for ethylene dimerization and butene isomerization J. Catalysis 61 2938.CrossRefGoogle Scholar
Stoessel, F., Guth, J. L. and Wey, R., 1977 Polymerisation de benzene en polyparaphenylene dans une montmorillonite cuivrique Clay Miner. 12 255259.CrossRefGoogle Scholar
Tate, K. R., Theng, B. K. G. and Theng, B. K. G., 1980 Organic matter and its interactions with inorganic soil constituents Soils with Variable Charge 225249.Google Scholar
Theng, B. K. G., 1974 The Chemistry of Clay-Organic Reactions London Adam Hilger.Google Scholar
Theng, B.K.G., 1976 Interactions between montmorillonite and fulvic acid Geoderma 15 243251.CrossRefGoogle Scholar
Theng, B. K. G., 1979 Formation and Properties of Clay-Polymer Complexes .Google Scholar
Thompson, T. D. and Tsunashina, A., 1973 The alteration of some aromatic amino acids and polyhydric phenols by clay minerals Clays & Clay Minerals 21 351361.CrossRefGoogle Scholar
Tricker, M. J., Tennakoon, D. T. B., Thomas, J. M. and Graham, S. H., 1975 Novel reactions of hydrocarbon complexes of metal-substituted sheet silicates: thermal dimerization of trans-stilbene Nature 253 110111.CrossRefGoogle Scholar
Ueda, T. and Harada, S., 1968 Adsorption of cationic poly-sulfone on bentonite J. Appl. Polymer Sci. 12 23952401.CrossRefGoogle Scholar
van der Velden, W., Chittenden, G. J. F. and Schwartz, A. W., 1973 Studies on the geochemistry of purines and py-rimidines Advan. Org. Geochem., Proc. 6th Int. Congr. 293304.Google Scholar
Vincent, B., 1974 The effect of adsorbed polymers on dispersion stability Advan. Colloid Interface Sci. 4 193277.CrossRefGoogle Scholar
Wang, T. S. C., Kao, M. M. and Huang, P. M., 1980 The effect of pH on the catalytic synthesis of humic substances by illite Soil Sci. 129 333338.CrossRefGoogle Scholar