Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Chemical Processes at the Water-Manganite (γγ-MnOOH) Interface by Madeleine Ramstedt Akademisk avhandling Som med tillstånd av rektorsämbetet vid Umeå Universitet för erhållande av Filosofie Doktorsexamen framlägges till offentlig granskning vid Kemiska institutionen, sal KB3B1, KBC-huset, onsdagen den 19 maj 2004, kl 13.00. Fakultetsopponent: Professor Steven A. Banwart, Groundwater Protection and Restoration Group, Department of Civil & Structural Engineering, University of Sheffield, England i TITLE Chemical Processes at the Water-Manganite (γ-MnOOH) Interface AUTHOR Madeleine Ramstedt ADDRESS Department of Chemistry; Inorganic Chemistry, Umeå University, SE-901 87 Umeå, Sweden ABSTRACT The chemistry of mineral surfaces is of great importance in many different areas including natural processes occurring in oceans, rivers, lakes and soils. Manganese (hydr)oxides are one important group to these natural processes, and the thermodynamically most stable trivalent manganese (hydr)oxide, manganite (γ-MnOOH), is studied in this thesis. This thesis summarises six papers in which the surface chemistry of synthetic manganite has been investigated with respect to surface acid-base properties, dissolution, and adsorption of Cd(II) and the herbicide N-(phosphonomethyl)glycine (glyphosate, PMG). In these papers, a wide range of analysis techniques were used, including X-ray photoelectron spectroscopy (XPS), extended X-ray absorption fine structure (EXAFS) spectroscopy, Fourier transform infra-red (FTIR) spectroscopy, atomic force microscopy (AFM), scanning electron microscopy (SEM), X-ray diffraction (XRD), potentiometry, electrophoretic mobility measurements and wet chemical techniques, in order to obtain a more complete understanding of the different processes occurring at the manganite-water interface. From the combined use of these techniques, a 1-pKa acid-base model was established that is valid at pH>6. The model includes a Na+ interaction with the surface: =MnOH2+½ log β0 (intr.) = -8.20 = -pHiep =MnOH-½ + H+ =MnOHNa+½ + H+ log β0 (intr.) = -9.64 =MnOH2+½ + Na+ At pH<6 the manganite crystals dissolve and disproportionate into pyrolusite (β-MnO2) and Mn(II)-ions in solution according to: 2 γ-MnOOH + 2H+ β-MnO2 + Mn2+ + 2H2O log K0 = 7.61 ± 0.10 The adsorption and co-adsorption of Cd(II) and glyphosate at the manganite surface was studied at pH>6. Cd(II) adsorption displays an adsorption edge at pH~8.5. Glyphosate adsorbs over the entire pH range, but the adsorption decreases with increasing pH. When the two substances are co-adsorbed, the adsorption of Cd(II) is increased at low pH but decreased at high pH. The adsorption of glyphosate is increased in the entire pH range in the presence of Cd(II). From XPS, FTIR and EXAFS it was found that glyphosate and Cd(II) form inner sphere complexes. The binary Cd(II)-surface complex is bonded by edge sharing of Mn and Cd octahedra on the (010) plane of manganite. Glyphosate forms inner-sphere complexes through an interaction between the phosphonate group and the manganite surface. The largest fraction of this binary glyphosate complex is protonated throughout the pH range. A ternary surface complex is also present, and its structure is explained as type B ternary surface complex (surface-glyphosate-Cd(II)). The chelating rings between the Cd(II) and glyphosate, found in aqueous complexes, are maintained at the surface, and the ternary complex is bound to the surface through the phosphonate group of the ligand. KEYWORDS manganite, γ-MnOOH, mineral surface, acid-base, adsorption, Cd(II), N-phosphonomethylglycine, glyphosate, PMG, surface complex, dissolution, disproportionation, XPS, EXAFS, infrared spectroscopy, SEM, AFM, potentiometry, electrophoresis ISBN 91-7305-634-0 74 pages and 6 papers ii How can I say thanks for the things you have done for me? Things so undeserved, that you gave to prove your love for me. The voices of a million angels could not express my gratitude. All that I am and ever hope to be, I owe it all to Thee… (Andrae Crouch) iii Printed by: Print & Media, Umeå University, Umeå, Sweden ISBN 91-7305-634-0 iv Chemical Processes at the Water-Manganite (γ- MnOOH) Interface Madeleine Ramstedt Department of Chemistry, Inorganic chemistry Umeå University SE-901 87 Umeå SWEDEN This thesis contains a summary and a discussion of the following papers, referred to in the text by their numbers 1-6. 1. Characterization of hydrous manganite (γ-MnOOH) surfaces – an XPS study. Madeleine Ramstedt, Andrei Shchukarev, Staffan Sjöberg, Surface and Interface Analysis, 34, 632-636, 2002 2. Surface Properties of Hydrous Manganite (γ-MnOOH) – A potentiometric, electroacoustic and XPS study. Madeleine Ramstedt, Britt M Andersson, Andrei Shchukarev, Staffan Sjöberg, Manuscript submitted to Langmuir 3. Low pH Phase Transformation and Proton Promoted Dissolution of Hydrous Manganite (γ-MnOOH). Madeleine Ramstedt, Staffan Sjöberg, Manuscript 4. Chemical speciation of N-(phosphonomethyl)glycine, PMG, in solution and at mineral surfaces. Madeleine Ramstedt, Caroline Norgren, Julia Sheals, Andrei Shchukarev, Staffan Sjöberg, Surface and Interface analysis, In Press, 2004 5. Thermodynamic and Spectroscopic Studies of Cadmium(II) – N- (phosphonomethyl)glycine (PMG) complexes. Madeleine Ramstedt, Caroline Norgren, Julia Sheals, Dan Boström, Staffan Sjöberg, Per Persson, Inorganica Chimica Acta, 357, 1185-1192, 2004 6. Co-adsorption of Cadmium(II) and Glyphosate at the Water-Manganite (γ-MnOOH) Interface. Madeleine Ramstedt, Caroline Norgren, Staffan Sjöberg, Per Persson, Manuscript Publications of interest but not included in the thesis: • A study of the Aqueous Geochemistry in the Udden Pit lake – Complete Dataset. Madeleine Ramstedt, Erik Carlsson, Lars Lövgren, MiMi-Report 2002:1, Stockholm, 2002, (ISSN 1403-9478, ISBN 91-89350-19-7) • Aqueous geochemistry in the Udden pit lake, northern Sweden. Madeleine Ramstedt, Erik Carlsson, Lars Lövgren, Applied Geochemistry, 18, 97-108, 2003 v Index Populärvetenskaplig sammanfattning.............................................................................................. 1 Introduction .................................................................................................................................. 3 Aim of the Thesis .......................................................................................................................... 4 Manganese - General Chemistry .................................................................................................... 5 Manganese - Geochemistry ........................................................................................................... 7 Manganese - Geochemical Cycle........................................................................................... 8 Manganese Oxides - Structures ..................................................................................................... 9 Manganite (γ-MnOOH) ................................................................................................................ 12 Surface Structure ................................................................................................................. 13 Mineral Synthesis................................................................................................................. 15 Methods ..................................................................................................................................... 17 Spectroscopy and Diffraction .............................................................................................. 17 X-ray Photoelectron Spectroscopy (XPS) ....................................................................... 18 X-ray Absorption Spectroscopy....................................................................................... 29 Infrared Spectroscopy ...................................................................................................... 32 X-ray Diffraction.............................................................................................................. 34 Microscopy........................................................................................................................... 36 Equilibrium Analysis............................................................................................................ 38 Potentiometric Titrations ................................................................................................. 38 Batch Adsorption Experiments ........................................................................................ 39 Electrophoretic Mobility Measurements.......................................................................... 39 Modelling......................................................................................................................... 40 Manganite Surface Processes...................................................................................................... 42 Surface Charge and Acid-base Properties .......................................................................... 42 Dissolution at Low pH ......................................................................................................... 47 Surface Complexes with Glyphosate and Cd(II).................................................................. 52 Manganite-Glyphosate ..................................................................................................... 52 Manganite-Glyphosate-Cadmium(II)............................................................................... 56 Summary.................................................................................................................................... 65 Future Research ......................................................................................................................... 67 References................................................................................................................................. 69 Acknowledgements………………………………………………………………………………………………74 vi Populärvetenskaplig sammanfattning Syftet med denna avhandling är att beskriva ytkemin hos ett manganmineral med namnet manganit ( -MnOOH). Ytkemin beskrivs med avseende på det rena mineralets* syra-bas egenskaper* och dess upplösning i sur miljö. Dessutom beskrivs hur manganit reagerar med kadmiumjoner och glyfosat. Glyfosat och kadmium är båda intressanta ämnen i ett miljöperspektiv, glyfosat eftersom det är en aktiv ingrediens i vissa växtbekämpningsmedel och kadmium pga att det är en giftig tungmetall som förekommer i vår miljö. Fördelningen i naturen av till expempel tungmetaller och näringsämnen påverkas av ett flertal processer, bland annat kan de adsorberas* på ytan av mikroskopiskt små mineralkorn. I vattendrag följer i sådana fall dessa kemikalier med mineralerna när de sedimenterar eller transporteras iväg i vattenströmmar. I naturliga vattendrag finns ofta stora mängder av dessa mineralpartiklar och de kan därför i stor grad påverka olika kemiska processer i vattnet. Ute i naturen är mineraler som innehåller aluminium, järn och mangan vanliga. Därför valde vi att studera ytkemin hos manganit som kan bildas i sjövatten och i jordar. För att studera manganits beteende och egenskaper användes ett flertal metoder. Spektroskopiska metoder ger information om ett provs sammansättning med avseende på mängd och struktur på molekylnivå. Mikroskopi beskriver den yttre formen hos manganitkristallerna och våtkemiska metoder användes för att undersöka kemiska förändringar i lösningarna och på ytan av manganit. I avhandlingen visas att manganitpartiklar får en negativ ytladdning vid alkaliska pH-värden. Ytans negativa laddning minskar då pH sjunker och vid pH 8.2 är ytan oladdad. Under detta pH-värde är ytan positivt laddad och under pH 6 löser manganit upp sig och omvandlas till ett mineral med namnet pyrolusit (β-MnO2). Vid alkaliska pH-värden adsorberar manganit gärna metalljoner men i närvaro av glyfosat minskar denna adsorption. Däremot ökar glyfosat metalljonadsorptionen vid neutrala pH värden. Adsorptionen av glyfosat ökade både vid neutrala och alkaliska pH-värden. I ett * mineral ~ sten- eller lermaterial syra-bas egenskaper ~ ett materials egenskaper vid olika pH adsorbera ~ fastna på ytan 1 naturligt vatten med pH 6-7, skulle detta till exempel kunna betyda att manganit adsorberar mer metalljoner om föreningar som glyfosat finns närvarande. Tidigare undersökningar har visat att nedbrytningen av glyfosat blir långsammare hos adsorberade molekyler. Våra resultat kan därför tyda på att glyfosatmolekylerna troligen kommer att brytas ner långsammare i närvaro av manganit. Dessutom, om även kadmiumjoner finns närvarade, kan nedbrytningen minska ännu mer eftersom kadmiumjoner binder upp den del av molekylen där nedbrytningen börjar. Däremot kommer adsorptionen att leda till en minskad rörlighet av både glyfosat och kadmiumjoner i jordar som innehåller manganit. 2 Introduction The chemistry of mineral surfaces is of great importance in many different areas including natural processes occurring in rivers, lakes and soils. Metal ions and other compounds can adsorb onto mineral surfaces and, thus, the fate of these substances can be greatly influenced by the surface chemistry of the mineral. In water, mineral particles can settle to the bottom of e.g. a lake, or be transported with water currents. In soils the adsorption onto stationary mineral surfaces can reduce the transport of different substances and also affect the bioavailability and degradation of these substances. Minerals can also dissolve and thereby release environmental contaminants or nutrients that previously were part of their structure or adsorbed onto the mineral surface.1,2,3 The scavenging of metal ions and other substances by mineral surfaces can proceed by surface complexation, surface precipitation, and formation of solid solutions. In surface complexation processes, adsorbents can form bonds to specific sites on the surface or be electrostatically attracted to the surface. Surface precipitation can be preceded by formation of surface complexes, and as the coverage of adsorbents is increased, the species at the particle surface polymerise to form a surface coating. The formation of solid solutions involves an exchange or insertion of substances into the crystal lattice of a mineral, thereby altering the composition of the solid.2,4 The dissolution of minerals also affects the content and distribution of different compounds in water solutions. This process can be beneficial if the dissolved ions are nutrients but harmful if they are environmentally hazardous substances. Many minerals dissolve through redox processes, such as the reductive dissolution of iron and manganese (hydr)oxides by the degradation of organic material.5,6,7 Fe, Al and Mn are abundant elements in the Earth’s crust and readily form (hydr)oxide minerals. Consequently, minerals such as iron(II, III), manganese(II, III, IV) and aluminium(III) (hydr)oxides form one important group in all of the processes mentioned above1. Although manganese is the least abundant of these three, it is reported to be the most mobile in some areas of the world8. This mobility leads to relatively high concentrations of 3 Mn(II) in surface waters9. The oxidation of Mn(II) and the formation of different manganese(hydr)oxides can effect the distribution of other substances. For example, in ferromanganese nodules, some heavy metals are enriched within the manganese fraction of the nodule instead of the iron fraction, indicating that these metals have a stronger affinity for manganese (hydr)oxides8. Thus, knowledge of the chemistry of manganese (hydr)oxides is essential to an understanding of different processes occurring in the environment. Manganese (hydr)oxides are a large and heterogeneous group of minerals with manganese oxidation states ranging from +II to +IV under natural conditions8. The oxidation state depends on the ambient conditions, and the chemical characteristics of the different manganese valences are reflected in the chemistry of the manganese (hydr)oxides. The ability to change oxidation state gives manganese (hydr)oxides the possibility to interact with other redox sensitive substances and participate in processes such as degradation of organic material6,7. The most stable of the trivalent manganese hydroxides, manganite (γ-MnOOH), is also a naturally occurring mineral8. Due to its composition, it can function both as an oxidising and a reducing agent, making it an interesting mineral in geochemical processes. Aim of the Thesis The objective of this thesis was to investigate the surface characteristics and reactivity of synthetically prepared manganite with respect to acid-base properties and metal/ligand adsorption reactions. Glyphosate (N-(phosphonomethyl)glycine) and cadmium(II) were chosen as model substances to investigate the reactivity of the manganite surface with small organic ligands and heavy metals, respectively. The two model substances are interesting from an environmental perspective and, furthermore, the molecular structure of glyphosate allows for the study of surface and metal ion interactions of different functional groups within the same molecule. 4 The experimental strategy employed here was to combine a range of different techniques to obtain a more complete understanding of the system studied. Various spectroscopic methods were used, together with x-ray diffraction, microscopy techniques, potentiometry, electrophoresis measurements, and batch adsorption experiments. From these techniques, information about the molecular structure of adsorbates, morphological structure of the mineral particles, and quantitative thermodynamic data were acquired and used to describe the chemical processes at the water-manganite interface. Manganese - General Chemistry Manganese is a 3d transition metal with an electronic configuration of [Ar] 3d5 4s2. In nature the oxidation states of (II), (III) and (IV) predominate10,11, but Mn can display any oxidation state from -III to +VII10. Mn shows typical behaviour of the 3d transition metals10, which is partly explained by the crystal field theory, with splitting of the d orbitals to form nongenerate energy levels10. The high spin species of Mn(IV) and Mn(II) predominate in nature and have electron configurations (d3 and d5 respectively) that give Mn(IV) but not Mn(II) a large stabilisation energy. Mn(III) generally displays Jahn-Teller distorted octahedra with two long apical distances and four short axial distances due to its d4 valence electron configuration10. Mn(II) is the most stable oxidation state of manganese in solution and it can form a number of different metal complexes with ligands12. Mn(III) ions hydrolyse readily according to (eq. 1-2)13 and are thermodynamically unstable in aqueous solutions (eq. 3)10, unless stabilised by a strong complexing ligand10 such as pyrophosphate14, EDTA or citrate15. Mn(III) can also be stabilised in the form of trivalent hydroxides at higher pH. Mn(IV) mainly occurs in the form of solid manganese oxides11. Mn3+ + H2O Mn3+ + 2H2O MnOH2+ + H+ Mn(OH)2+ + 2H+ 5 log K = 0.4 (1) log β = 0.1 (2) 2Mn3+ + 2H2O Mn2+ + MnO2(s) + 4H+ log K = 9.5 (3) The thermodynamic stability of the different manganese (hydr)oxides is determined by oxygen partial pressure, redox conditions (pe), pH and temperature. Figure 1a shows the distribution of predominant oxidation states and compositions with pH and redox conditions1,11 and Figure 1b shows the disproportionation of manganite at low pH. a) b) pe 120 MnOOH(s) 3 2- mmoles/dm MnO4 Mn(OH)3 80 2+ Mn & MnO2(s) 40 - 0 2 3 4 5 6 7 8 pH pH -5 -3 Figure 1. a) pe – pH diagram for the Mn-H2O-CO2 system at 25 °C. Mntot = 10 M, Ctot = 10 M, the dashed line represents the equilibrium: H2O + - 1 ½ O2 + 2H + 2e (modified from Stumm and Morgan ). b) Distribution of Mn species with pH showing the disproportionation of manganite forming β-MnO2 2+ 11 and Mn at low pH (paper 3). The oxidation of Mn(II) is complex and the final product is dependent on the prevailing conditions during the oxidation process. Extensive efforts have been made to describe the oxidation from Mn(II) to Mn(IV) 16,17,18,19,20. These studies show that under most laboratory conditions, manganese does not form Mn(IV), but more often the end product is one of the Mn(III) hydroxide polymorphs11,21,22. In lake water, a mixture of hausmannite (Mn3O4) and feitknechite (β-MnOOH) has been reported to precipitate, which with time transforms into manganite 20,21,16,19,22. This indicates that also in natural waters the oxidation of Mn(II) forms MnOOH instead of MnO2. However, at low manganese concentrations and in the presence of bacteria, amorphous Mn(IV) oxides have also been reported to form30,18,23. 6 The Mn(II) oxidation process is also affected by different ligands. In the presence of weak complexing ligands, such as nitrate or chloride, a mixture of hausmannite and feitknechite has been shown to form with pyrochroite (Mn(OH)2) as a precursor, but if sulphate was present, the precipitate consisted of manganite16. To explain this phenomena, it was suggested that the ion pair MnSO40 inhibits the formation of hexagonal Mn(OH)6 plates and, in this manner, prevents β-MnOOH from being formed16. In contrast, oxalate has not been shown to influence the solid phase speciation, although it slows down the oxidation process due to the strong complex formation22. High concentrations of oxalate were also shown to decrease the transformation of hausmannite and feitknechite into manganite22. Manganese - Geochemistry One of the reasons for the environmental interest in manganese (hydr)oxides is that they act as scavengers for heavy metals and other substances in natural waters and soils1,8,24,25. Furthermore, Mn can form oxide coatings on other minerals and bacteria and in this manner form reactive surfaces24. It has been reported that heavy metals, such as Cd or Co, have a higher affinity for manganese oxides compared to iron- and aluminium oxides26,27,28. Also, accumulated amounts of heavy metals have been reported in ferromanganese nodules, where some metals in particular, Ni, Co, Pb and Mo, seem to be concentrated within the manganese fraction of the nodules8. These findings emphasise the importance of studies of metal adsorption onto the different types of manganese (hydr)oxides. Manganese occurs mainly in sedimentary-, metamorphic- and igneous rock, but Mn is also present in the form of colloidal manganese oxides in most soils and waters. The average rock contains about 0.1 % manganese which makes manganese the tenth most abundant element in the Earth’s crust24,25. Although it is less abundant than both iron and aluminium (1/50 and 1/25 respectively), it is the most mobile of the three8. The mobility can partly be explained by the higher solubility of the Mn (hydr)oxides compared to Al or Fe (hydr)oxides under ambient conditions10, but also by the slower oxidation rate for Mn(II) compared to Fe(II)1,8. This mobility is highest in sub-arctic and temperate soils, which have high organic content and acidity8. In these regions, Mn leaches from the soil and is transported in dissolved forms or as 7 small colloids in lakes and rivers9. Tropical or arid soils are generally more alkaline and have a lower relative content of organic material. Thus, in these regions, Mn(II) oxidises within the soil or forms immobile coatings on rock surfaces on top of the soil, which results in low mobility of Mn8. Apart from weathering of Mn containing rock, Mn also originates from anthropogenic or biogenic sources, although to a much smaller extent. The metal industry and combustion of fossil fuel are the main anthropogenic sources25, while the biogenic sources are mainly decomposed plants and animals8,25. The manganese content in plants is due to its catalytic position in photo system II where it oxidises water to form oxygen during the photosynthesis29. For animals, manganese is an essential nutrient25 since it forms reactive sites in different enzymes29, e.g. in superoxide dismutas that react with oxygen radicals in the body1. Manganese - Geochemical Cycle In the geochemical cycle of manganese, dissolved Mn(II) is produced mainly from weathering of metamorphic and igneous rock24,25. Mn(II) ions are transported in water where they form complexes with different ligands, adsorb onto particle surfaces, or become oxidised to different manganese oxides or hydroxides8,25 (Figure 2). The oxidation process can be autocatalytic, but most commonly the oxidation of manganese is mediated by different microbes8,25,30. When the particles grow in water they eventually settle to the bottom of the lake and become part of the sediments. However, if the redox conditions are altered, e.g. in the presence of a redoxcline, the manganese particles are reduced back into Mn(II) ions. These ions will diffuse along the redox gradient until they reach oxidising conditions where they will be reoxidised and form (hydr)oxide particles8. This cycling around the redox boundary leads to an enrichment of Mn at specific depths, which can be detected in natural lakes25. The redox boundary is usually found close to or within the sediments in shallow lakes while, in deeper stratified lakes or lakes with high concentration of organic matter, it is present within the water column25. In rivers, where the water is well oxygenated and constantly moving, these redox boundaries are rare and the manganese is mainly present in the form of colloidal (hydr)oxide particles, although, dissolved Mn(II)† can also be present depending on season9,31. In the Kalix river in northern Sweden dissolved Mn(II) dominated † filtered through 45µm filter 8 during winter and particulate Mn during summer9,31. The manganese (hydr)oxides settle to become part of the sediments, and with time they form metamorphic or igneous rock as a consequence of sediment compression, which completes the geochemical cycle for Mn. Uptake by biota MnL(aq) Mn(II) Mn2+ MnL(s) X:Mn2+ MnOx MnOx redoxcline Mn2+ Figure 2. The fate of manganese in natural water where manganese can form aqueous complexes 2+ (MnL(aq)) or solid phases(MnL(s)) with different ligands, adsorb onto particulate matter (X:Mn ), oxidise to form different manganese (hydr)oxides (MnOx), or be consumed by biota. The solids formed slowly settle and end up in the sediment. Manganese Oxides - Structures Manganese oxides and hydroxides are among the most common manganese containing minerals32. More than 30 manganese oxides are known, although not all of their structures have been determined24. The oxidation of Mn results in precipitation of different oxides depending on surrounding conditions, and often an intergrowth of two or more oxides are found in natural deposits24. The variety of different structures in the manganese oxides is illustrated in Figure 3. Most manganese minerals have a framework constructed of MnO6 octahedra sharing corners or edges8. From these different architectures, two large groups can 9 be formed that include most of the structures i.e. tunnel and chain structures and layered structures (Figure 3 and Table 1)24. a d b e c f Figure 3. The arrangement of MnO6 octahedra in a few selected manganese oxides; a) pyrolusite, b) ramsdellite, c) hollandite, d) romanechite, e) birnessite (layers of MnO2 with metal ions in the interspace), f) lithophorite (layers of (Al,Li)(OH)2 and MnO2 alternating in the structure). In c) and d) counter ions and water have been omitted for clarity. 10 The structure with the smallest tunnel size is found in pyrolusite (Figure 3a), which is the most thermodynamically stable Mn(IV) oxide under ambient conditions11. Pyrolusite has a rutile (TiO2) type structure with rows of MnIVO6 octahedra sharing edges and corners. Manganite has a similar structure. However, the MnIIIO6 octahedra are Jahn-Teller distorted8,24, and half of the oxygen atoms are replaced by OH-groups38,33,34. Ramsdellite is isostructural with groutite (α-MnOOH), goethite (α-FeOOH) and diaspore (AlOOH) and displays tunnels which are slightly larger compared to pyrolusite (Figure 3b)24. An intergrowth between the pyrolusite and ramsdellite structures is found in the nsutite mineral, which has single and double chains randomly alternating in the structure, forming a mixture of small tunnel sizes35. Larger tunnels are found in e.g. todorokite, romanechite and the minerals in the hollandite group (e.g. hollandite, cryptomelane, coronadite, manjiroite) (Figure 3c & d). In these latter minerals, the tunnels are built from Mn octahedra with a mixture of Mn valences and, thus, the tunnels have cations present inside the tunnels to balance the charge from the manganese framework24. Lithophorite, chalcophanite and the minerals in the birnessite group are examples of manganese minerals that have a layered structure (Fig 3e & f). They are all constructed from charged layers of manganese octahedra and have different counterions in the interspace between the layers. Pyrochroite also has a layered structure but without cations in the interspace and with sheets constructed from MnII(OH)6 octahedra. It has not been possible to determine the crystal structure of several of these layered minerals by the analysis of natural samples24. Instead, the existing structures have been determined using synthetic analogues. For example, birnessite, which is present in most soils, has been structurally determined in the form of a synthetic potassium rich birnessite. There also exist Mn oxides that have different structures from the ones discussed above. The most common, hausmannite, has a spinel type structure. This structure consists of oxygen atoms in cubic close packed arrangement, Mn(III) in ½ of the octahedral holes, and Mn(II) in ¼ of the tetrahedral holes. Apart from the layered pyrochroite, two Mn(II) structures are included in Table 1. Manganosite has a halite (NaCl) type structure. It is a mineral that can be found in metamorphic ore and can be produced by heating rhodochrosite. Rhodocrosite has a calcite (CaCO3) type structure36, and although it is not a manganese oxide, it is included here since it is commonly present in sediments and rearranges to form manganese (hydr)oxides11. 11 Table 1. Some manganese minerals with chemical composition and structure24,8. Mineral name Chemical formula Alternative name Mn valence Manganosite MnO +II Pyrochroite Mn(OH)2 +II Rhodocrosite MnCO3 +II Hausmannite Mn3O4 +II, +III Groutite +III α-MnOOH Feitknechite# β-MnOOH +III Manganite +III γ-MnOOH Birnessite (Na,K,Ca)4Mn7O14 ×2.8 H2O δ-MnO2, +II, +III +IV 7Å manganate Vernadite* # MnO2 ×nH2O Lithophorite (Al,Li)(OH)2 × MnO2 +III, +IV Chalcophanite (Zn, Mn, Fe)Mn3O7 × 3H2O +II, +IV Todorokite (H2O,…)2-xMn8(O,OH)16 10Å manganate +III, +IV +III, +IV Romanechite (Ba,Mn2+,…)3(O,OH)6Mn8O16 Psilomelane Hollandite Ba2-xMn8O16 +III, +IV Cryptomelane K2-xMn8O16 +III, +IV α-MnO2 Coronadite Pb 2-xMn8O16 +III, +IV Manjiroite Na 2-xMn8O16 +III, +IV +IV (+II, +III, Nsutite Mn(O,OH)2 × H2O, γ-MnO2 +IV)35 (Mn4+1-xMn2+xO2-2x(OH)2x) Ramsdellite MnO2 +IV Pyrolusite +IV β-MnO2 Str L T L? T L L? L L T T T T T T T T T *) Synthetic δ-MnO2 analogue24. # ) no crystal structure has been determined for this mineral. Str = structure, T = tunnel structure, L = layered structure, L?=layered structure is suggested in the literature Manganite (γ-MnOOH) Manganite (γ-MnOOH), is the most stable of the trivalent manganese hydroxides8 and is the hydroxide most likely to be present in lakes and rivers with high manganese concentrations and/or high sulphate concentrations22. It occurs in the form of colloidal particles in water and sediments. However, due to the difficulties of sampling colloidal particles, it has been suggested that the presence of manganite has been overlooked in many routine samplings37. In sediments, manganite is commonly found as a layer between the anoxic MnCO3 phase and the fully oxidised pyrolusite (β-MnO2) layer8. Careful studies of the manganite and pyrolusite boundary show that the manganite is pseudomorphously replaced by pyrolusite during oxidation8,11. A similar process can be observed when manganite is heated in air and oxidises to pyrolusite. The only morphological change observed from this heat induced process is a 12 decrease in the unit cell length38 corresponding to the contraction of the elongated MnIIIO6 octahedra. Proton promoted dissolution of manganite at low pH has been observed by several authors5,39,40. The dissolution occurs through disproportionation of Mn(III) into Mn(II) ions in solution and one of the solid Mn(IV) oxide polymorphs (eq. 4). 2 MnOOH(s) + 2 H+ MnO2(s) + Mn2+ + 2 H2O log K~ 7.0-8.6 (4) To our knowledge only two studies have been performed to understand this proton promoted dissolution11,41 and in many low pH adsorption studies onto manganite the dissolution has not been addressed although the adsorption experiments have been conducted in a pH region where dissolution is significant. Surface Structure The adsorption of different substances onto a mineral surface is often interpreted using the structure and composition of the surface. The structure of the mineral surface in aqueous solution is often not known in detail but a fairly good assumption is that it resembles the bulk structure42. Since the three dimensional framework in the bulk is lost at the surface, metal ions will be under-saturated with respect to bond formation2. Thus, oxygen atoms positioned at the metal hydroxide surface can be pictured as having fractional charges according to Pauling’s valence bond theory43. According to this theory, the metal ion displays a fraction of its charge at each coordinating atom or group. For example, a trivalent metal ion within an octahedron of oxygen atoms would have a formal charge of +½ at each bond. Due to the presence of water, the surface of a hydroxide or oxide mineral will not be coordinatively unsaturated but will be covered by hydroxyl groups once immersed into a pure aqueous environment. At the surface this results in hydroxyl groups, displaying a charge of -½ and doubly protonated oxygen atoms with a charge of +½ (since OH2 is neutral). If the hydroxyl groups are coordinated to two metal centres instead of one (i.e. doubly coordinated) they will display a charge of 0 or, when protonated, +1. The net charge of the surface will be positive when the number of protonated groups is higher than the number of deprotonated groups, and vice versa. The pH at which the total positive charge is equal to the negative charge is called pHpzc 13 (pzc = point of zero charge). Thus, the change in charge at the surface can be seen as a result of protonation and deprotonation of hydroxyl groups. Assuming that the surface displays the same structure as the bulk, a theoretical number of adsorption sites can be estimated for specific planes of the crystals. The dominating planes for small manganite crystals have not been determined and may differ compared to larger crystals44. However, since the (010) plane has been shown to be predominant in large manganite crystals and since it also constitutes a plane of cleavage34,45, we have assumed that the small manganite crystals of this work are dominated by the (010) plane. On the (010) plane, the surface will display singly coordinated oxygen (=O-3/2) or hydroxide groups (=OH-½) (Figure 4). In water suspensions these groups will be protonated44 to form =OH-½ or =OH2+½, depending on surrounding pH (paper 2). The singly coordinated oxygen atoms can be assumed to be responsible for the reactivity at the (010) plane42 since the doubly coordinated oxygens display lower charge. The triply coordinated oxygens are positioned further down in the structure, and are here assumed to be less reactive than the singly coordinated oxygens. The (010) plane hosts 7.9 singly coordinated oxygens /nm2 (13 µmol/m2). doubly singly triply (001) (010) 34 Figure 4. The manganite structure with the (010) plane visualised by dotted lines. The plane of the paper is (010) in the left structure and (001) in the right structure. The arrows show singly-, doublyand triply coordinated oxygen atoms at the surface. 14 This density of crystallographic sites on the manganite surface was used in the work presented here due to the difficulties and absence of good methods for experimentally determining site densities. In the literature there is, to our knowledge, only one previously reported study on manganite site density. This was a fluoride adsorption study, which yielded 5 sites /nm2 or 8.3 µmol sites /m2 of manganite5. Mineral Synthesis The manganite used in this thesis work was synthesised according to Giovanoli and Leuenberger46, by adding 300 cm3 of 0.2 M NH3 (Merck p.a.) to a solution of 20.4 cm3 30 % H2O2 (JT Baker p.a.) and 1 dm3 0.06 M MnSO4 (BDH). The mixture was rapidly heated and thereafter kept at a temperature of 95 °C for six hours with constant stirring. The formed suspension was filtered, still hot, through a G4 glass filter, washed with equal amounts (with respect to the initial solution) of hot deionised water, collected on watch glass and dried under reduced pressure in presence of P2O5 (Riedel-de Haën). The dried manganite was ground to a powder, the presence of crystalline phases checked using X-ray powder diffraction and the crystallinity with High Resolution Transmission Electron Microscopy (HRTEM) (Figure 5). 4 nm Figure 5. Transmission electron microscopy image of part of a manganite crystal showing the inner crystalline structure. The scale bar in the upper left corner represents 4 nm. The width of the needle in this picture is ~20 nm (paper 1). 15 XRD of the synthesised manganite gives the typical d-spacings for manganite, at 3.41, 2.64, 2.52, 2.41, 2.20, 1.78, 1.70, 1.67, 1.50, 1.44 and 1.32 Å. The synthesis usually yielded only pure manganite but, at a few occasions, a trace of hausmannite was found in the diffraction patterns. When this happened, the specific batches were discarded and not used further. The manganite was stored as a powder, from which suspensions with known amounts of solid were prepared. Suspensions were placed in an ultrasonic bath for a few minutes where they were shaken from time to time to avoid over heating and to keep the particles from settling. The suspensions were thoroughly bubbled with Ar(g) at the time of preparation and were stored with Ar atmosphere in the bottles. Typically 10 g of manganite was used per dm3 suspension. The manganite crystals seem to disperse well in suspensions, although they form aggregates when dried as seen in microscopy images (Figure 6). During the work of this thesis it was found that the potentiometric titrations were reproducible if the suspensions were allowed to age for a few months, which most likely is due to Oswald ripening processes i.e. the growth of larger crystals on the expense of smaller ones. a) b) Figure 6. Scanning electron microscopy images of dried manganite crystals. The scale bars represent 3µm in image a and 750 nm in image b. Length of scale bar: The surface area of manganite was experimentally determined after synthesis using the BET N2 adsorption method47 at 70 °C. This temperature was chosen in order to prevent phase transformations40. It was discovered that heating manganite under a nitrogen atmosphere converts it into hausmannite. However, if heated in air, then pyrolusite is formed, which supports the findings of e.g. Giovanoli and Leuenberger46. Since the tests did not show any changes in XRD patterns below 100 °C, a temperature of 70 °C was chosen as the degassing 16 temperature to ensure that no phase transformation occurred. The BET area varied between 30 and 51 m2/g for the manganite synthesised in this work, and the difference in surface area between the batches is ascribed to the size difference between the manganite crystals46. The synthetic manganite is crystalline and needle-shaped with an average length of approximately 500 nm and a width of about 20 - 40 nm (Figure 5 & 6). Methods In this study, chemical processes at the manganite surface were investigated using a number of different analysis methods. This was done in order to have a broader understanding of the manganite surface and reactions occurring at the manganite-water interface. Spectroscopy, x-ray diffraction, microscopy and wet chemical techniques were used to examine the different properties of the material and to interpret the chemical processes at the surface both quantitatively and at a molecular level. The combination of different techniques gives the possibility to validate data and also a better understanding of the surface processes can be obtained. Spectroscopy and Diffraction Different spectroscopic and diffraction techniques can be used to obtain molecular level information about a sample. The common theme in most spectroscopic methods is that electromagnetic radiation is used and the effect from the interaction between the electromagnetic radiation and the sample is recorded48. Three different types of spectroscopic techniques were used in this thesis, X-ray Photoelectron Spectroscopy (XPS), X-ray Absorption Spectroscopy (XAS) and Fourier Transform Infrared (FTIR) Spectroscopy. From these three techniques, quantitative and qualitative information about the composition of surface species, the molecular structure and coordination around a specific atom, and the molecular structure and identity of compounds can be obtained. Furthermore, X-ray diffraction (XRD) was used to identify and characterise the structure of the crystalline phases. 17 X-ray Photoelectron Spectroscopy (XPS) XPS is used to determine and quantify the chemical composition of a sample surface. Theory When X-rays strike a surface, high energy photons cause ejection of electrons from the atoms in the sample. By measuring the kinetic energy of the emitted electrons and knowing the energy of the incident beam, the electron binding energy can be calculated using eq. 5. EB = hν - EK - φ (5) where EB is the binding energy in eV hν is the energy of the incoming photon EK is the kinetic energy in eV and φ is the spectrometer work function, determined in separate measurements. The binding energy for core level electrons is specific for each element and orbital. This makes it possible to elucidate the elemental composition of a surface from the binding energies obtained in XPS. In addition, the binding energy of an element is influenced by the chemical surroundings and oxidation state of the atom, making it possible to distinguish between different chemical forms of an element e.g. between protonated and deprotonated amine groups (Figure 7). All elements except H and He can be directly detected using XPS. However, this does not mean that no information about H can be obtained, since the presence of H will shift the binding energy of atoms bound to H (Figure 7). The quantitative estimate of the different surface components are given in atomic % of the surface (< ± 10 %) and the detection limit is approximately 0.1 atomic %49. 18 x 10 120 1 Name NH2 + NH Pos. 402.0 399.9 FWHM 1.38 1.56 Area % 659.6 48.5 701.0 51.5 110 Counts/s 100 90 80 412 408 404 400 Binding Energy (eV) 396 Figure 7. Nitrogen 1s spectrum of aqueous glyphosate at pH 9 showing the chemical shifts as a result of the protonation (paper 4). XPS spectra are created by electrons emitted from the top atomic layers of a sample, making it a surface sensitive technique. Electrons ejected from atoms deeper down in the sample collide with atoms or electrons on their way to the surface and loose some of their specific kinetic energy. These electrons form the background intensity in XPS spectra. The analysis depth (or information depth) is dependent on the kinetic energy of the measured photoelectron together with the sample density and the type of elements present. For an electron with a kinetic energy around that of O 1s, the analysis depth would be roughly 4.5 nm for a mineral such as manganite, 3 nm for Mn metal and 6.5 nm for an organic substance such as glyphosate‡,50. However, since the likelihood of electrons escaping the surface falls off exponentially to the depth, the major part (63 %) of the signal originates from a layer about 1/3 of the analysis depth51. ‡ Estimated as 3 × attenuation length for an electron with KE=650 eV in the material in question. Attenuation length calculated using NIST database50 19 x 104 40 35 O 30 Mn CPS 25 20 15 10 5 1000 800 600 400 Binding Energy (eV) 200 0 Figure 8. Wide spectrum for a sample of manganite with glyphosate adsorbed at the surface. The arrows show the broader Auger lines from Mn and O atoms. The narrow lines represent photoelectron lines from the different elements present in the sample. In addition to the specific photoelectron lines in the spectra, Auger lines can be seen (Figure 8). These lines originate from rearrangements of electrons within the atoms. When an electron is ejected from a core level, a vacancy is created, which can be filled by a relaxation process. An electron from a higher orbital moves to fill the empty site (Figure 9), and energy is released either in the form of X-rays or as another ejected electron, i.e. an Auger electron. The Auger lines are usually broader and have a more complex substructure, compared to photoelectron lines. From Auger lines information about oxidation state of elements can be extracted. 20 (a) (b) h h 4 4 1 3 2 Figure 9. Schematic drawing of an oxygen atom exposed to X-rays (h ). (a) shows the photo-ejection of an electron from the O 1s orbital. (b) shows a subsequent reorganisation resulting in emission of X-rays or an Auger electron. The sequence of the processes is shown by the numbers 1-4. Instrument For this work, monochromatic Al Kα X-rays (h = 1486.6 eV) have been used as incident radiation, but it is also common to use non monochromatic Mg Kα or Al Kα radiation for XPS analyses. The non monochromatic X-rays have a higher intensity, but can contain other X-ray lines apart from Kα, which can give rise to additional peaks in the spectra. Furthermore, a monochromatic X-ray beam gives better energy resolution. In this work, the monochromator used for the Al Kα X-rays consists of a set of quartz crystals positioned to reflect the beam so that only a specific wavelength of the beam reaches the sample surface. If the surface is not electrically conductive, it will build up a positive charge as a consequence of the ejection of electrons. This charge will influence the electrons leaving the surface and to prevent this, charge neutralising equipment is used. In this work, one of the most common techniques for charge neutralisation was used, which is to flood the surface with low energy electrons (< 20 eV). Despite this treatment, there are differences in charging between samples, and to be able to compare data, the binding energy scale has to be calibrated. Most surfaces analysed with XPS display an amount of hydrocarbons due to surface contamination. This adventitious carbon can be used as an internal standard, and the difference in measured EB for the aliphatic carbon component and the literature value (285.0 eV) is used to shift all the 21 spectra to obtain a more accurate scale. This approach assumes that the surface contamination exhibits the same charging behaviour as the rest of the sample and that it does not interact with sample surface. This is not always the case and suggestions to deposit a small gold dot on the surface, or to use a well characterised component from the sample matrix as internal standard have been made51,52. The first suggestion requires a gold evaporator and careful control of the thickness of the gold since different thickness of gold films have been proven to influence the BE for the Au lines. The second of these suggestions requires that the reference component always has the same BE independent of the surrounding conditions in the sample. In this thesis, the approach with the adventitious carbon has been used. The absence of interactions between the aliphatic carbon and the manganite cannot be guaranteed, but should be more or less identical in all samples. Thus, this internal calibration enables comparisons to be made between samples. Furthermore, the low concentration of carbon at the surface indicates that the manganite does not have a strong affinity for hydrocarbons. XPS measurements take place under ultra-high vacuum conditions, i.e. at pressures below 10-8 Torr. There are several reasons for this; one is to prevent the electrons leaving the sample surface from colliding with gaseous atoms and molecules. These collisions would result in loss of the specific kinetic energy of the electrons and, thus, loss of signal intensity. Another reason is that the high vacuum keeps the sample and the analyzer (f-i in Figure 10) clean from surface contaminations. Unfortunately, the vacuum conditions result in “disappearence” of volatile substances at room temperature. However, this disadvantage can often be overcome by freezing the samples using N2(l) in order to reduce the evaporation process. 22 _ h + g i f e d c b a Figure 10. Schematic picture of the pathway of the electrons leaving the sample in XPS a) b) c) d) e) f) Sample Magnetic lens Iris Scan plates Spot size aperture Electrostatic lens g) Retention grid h) Hemispherical electron energy analyzer i) Detector system – electron multiplier = electron path In this work, the kinetic energy of the electrons is detected in a concentric hemispherical analyser (Figure 10h). The analyser consists of two hemispheres, one positive and one negative, concentrically located on an insulating pedestal. On one end there is an entrance slit and a retarding grid, and on the other there is an exit slit and an electron multiplier. By changing the potential between the two hemispheres, a path between them is created which only allows electrons with a specific kinetic energy (EK) to reach the electron multiplier, while electrons with different EK will collide with surfaces on either side of the path. The energy an electron must have in order to pass through the analyser is called the pass energy, and this energy is proportional to the voltage difference between the two hemispheres. The number of electrons passing through the path is, thus, decided by the pass energy. Using a low pass energy, e.g. 20 eV, results in lower intensity of the output signal but better spectral resolution. The retarding grid serves to maintain constant analyser energy, by accelerating 23 electrons in the beginning of a scan, when the electrons have low energy, and decelerating electrons in the end of the scan. The electron multiplier consists (for the work presented here) of a set of 8 channeltrons (channel electron multipliers). The channeltrons are constructed of glass tubes coated on the inside with a resistive material that has a high secondary emission coefficient. Thus, an electron entering a tube generates many more electrons when colliding with the tube walls. The secondary electrons produced continue to multiply in a cascade manner, and at the end of the channeltron the resulting electrical current is measured and fed to software where it is displayed as intensity in a spectrum. To control the analysis area, different apertures (e in Figure 10) are used. For the work presented here, a slot was employed that gives an analysis area of 0.3 × 0.7 mm. The energy resolution for the spectra is estimated from the width of the Ag 3d5/2 line measured on a Ag standard sample. At 20 eV pass energy and with the above slot, this line had a FWHM (full width at half maximum) of 0.63 eV for the instrument used in this study. For more information about the XPS technique, the reader is suggested to consult reviews by Hochella51 Paterson & Swaffield53 and Ratner & Castner49. Experimental procedure For the XPS analyses, a relatively thin layer of the wet paste, or a small drop of solution, was applied onto a sample holder and directly placed on the pre-cooled stage (-170 °C) in the sample introduction chamber of the spectrometer. The samples visibly froze within approximately 5 - 10 seconds and, subsequently, the pressure in the sample compartment was reduced (while maintaining the sample at approximately -170 °C). When required pressure was obtained (~5 × 10-7 Torr), the sample was introduced into the analysis chamber, placed onto the pre-cooled stage and the measurements started. The temperature of the sample manipulator was kept at approximately -155 °C or below throughout the entire measurement. This treatment maintained water in the sample and kept the pressure in the analysis chamber below 10-8 Torr. Next, the frozen paste was left overnight inside the spectrometer, with the main valve open (to increase pumping capacity) and high voltage switches off (for safety). After reaching room temperature, the sample was remeasured the following day in the exact same spot (Figure 11). 24 x 102 x 102 20 Name Pos. FWHM O 1.10 530.3 OH 531.5 1.35 H2O 533.5 1.60 18 Area 1774.7 2350.0 573.2 % 37.8 50.0 12.2 30 Name Pos. O 530.2 OH 531.5 OH2 532.8 a) FWHM 1.07 1.38 1.72 Area 2707.7 3775.3 242.7 % 40.3 56.1 3.61 b) 25 14 20 12 15 10 8 10 6 4 5 2 544 540 536 532 544 528 540 Binding Energy (eV) 536 532 528 Binding Energy (eV) Figure 11. O 1s XPS spectra of a) frozen manganite paste and b) room temperature paste the subsequent day, showing that water is maintained at the interface in the frozen samples (paper 1). NH Intensity Intensity (counts/s) 16 + NH2 407 405 403 401 399 397 395 Binding Energy (eV) Figure 12. Spectrum showing the changes in N 1s with loss of water. The solid line represents glyphosate adsorbed onto goethite analysed as a frozen paste. The dotted line represent the same sample but at room temperature (paper 4). 25 This approach enables observations of changes caused by the disappearance of water (Figure 11 & 12) and the ability to estimate the water layer thickness. Typically this water layer was estimated to 0.6 nm for the manganite paste. The rough estimate of the water layer thickness on manganite was made by using an equation for calculating the thickness of SiO2 on Si54, shown in eq. 6. Ipaste = Iwarmed sample × e-d(aq)/(λ sinθ) (6) Ipaste is the intensity of the Mn 2p peaks in the frozen paste, Iwarmed sample is the intensity of the same peaks when the water layer is removed, d(aq) is the thickness of the water layer, λ is the attenuation length of Mn 2p electrons in water and θ the take-off angle (θ = 90 and, consequently, sin θ =1). The attenuation length, λ, was calculated using the NIST standard database50. However, this thickness is only a rough approximation since eq. 6 assumes a flat surface, but the samples used in this thesis are in the form of powders. In XPS analyses of frozen droplets of solution, a large concentration effect was observed in relation to the actual concentration of the solution. In an aqueous solution containing 1 mM glyphosate, NaOH (for pH adjustment) and H2O, the ratio between the area of the water peak (~55 M) and the components of the glyphosate molecule (1 mM) should be very large (55/0.001=55000). However, XPS spectra displayed a much lower ratio between water and glyphosate suggesting that the glyphosate was concentrated through the freezing procedure and/or analysis procedure. A third option could be that the solution-air interface displays a different composition than the bulk of the droplet. The concentration effect might arise through the evaporation of water from the frozen sample droplet. This would leave a higher percentage of glyphosate at the surface than originally was present. XPS spectra supported this hypothesis. It was observed that an increase in the analysis time decreased the O/N ratio, which indicates that water molecules leave the sample in the vacuum even at these low temperatures (-155 °C). This evaporation could also be detected as a slight increase of pressure in the analysis chamber from ~5×10-10 to ~5× 10-9 Torr. 26 In spectra of frozen pastes, a contribution from the small amounts of remaining bulk solution in the paste is possible. This has been observed for the ionic medium and raises the question if non-adsorbed ligands in solution can contribute to the XPS spectra. The problem with NaCl originating from the solution can be quite easily removed by subtracting e.g. the Na+ concentration by the Cl- concentration, if Na+ is in excess in the sample and Cl- does not interact with the surface this would give the excess of Na+ at the surface. However, since the concentration of adsorbing ligands is much lower than the concentration of ionic media, (usually 100 mM ionic medium compared to < 5 mM ligand), the contribution in spectra from non-adsorbed ligands should be much smaller smaller. In paper 4, the concentration of protonated and deprotonated glyphosate was investigated as a function of pH using XPS. It was observed that at higher surface loadings the concentration of protonated glyphosate at the manganite surface increased, while the concentration of deprotonated glyphosate stayed at the same level (Figure 13). This was suggested to be an effect from surface saturation with respect to the deprotonated species, similar to previous observations of glyphosate adsorption onto goethite at neutral pH values94 (paper 4). However, this increase could also be due to an increased amount of protonated glyphosate in the solution that contributed to the XPS spectra. Atomic ratio N/Mn 0.08 0.06 0.04 0.02 0 5 6 7 8 9 10 pH Figure 13. Atomic ratio of N 1s components in glyphosate adsorbed to manganite. The amount of protonated ( ) and deprotonated ( ) amine at the surface of: manganite with 14.6 µmol glyphosate (empty symbols), 9.8 µmol glyphosate (shaded symbols) and 2.4 µmol glyphosate (filled symbols) 2 added per m of manganite (paper 4). 27 To check this, closer examinations were made of the glyphosate-MnOOH system. Throughout the pH range, the XPS spectra indicated the presence of both protonated and deprotonated amine groups (Figure 13). This could also be observed in FTIR spectra (in the form of two types of vibrations for the carboxylic group and the amine group). Since the contribution of solutes in FTIR can be removed, these infrared data represent only the glyphosate adsorbed onto the manganite surface. The difference in the concentration of protonated amine groups observed in the XPS spectra should also be seen by subtracting infared spectra of different surface loadings of glyphosate at the same pH (Figure 14). If the intensity due to protonated glyphosate is greater in one spectrum, then this would give rise to a subtraction spectrum identical to the spectrum of the protonated glyphosate (Figure 14c). However, no difference was found in the regions for the two carboxylate-amine vibrations (~1600 & ~1400 cm-1). Thus, the increase in protonated amine detected in the XPS spectra can also be explained by a contribution from remaining solution in the paste. This solution effect is increased since solutes are concentrated by the vacuum (as described above). a) b) c) 1700 1600 1500 1400 /cm 1300 1200 -1 Figure 14. Difference between FTIR spectra of different surface loadings of glyphosate onto manganite at pH 6.4. Vibrations indicating the protonation of the glyphosate should be observed -1 -1 2 around 1600 cm and 1400 cm . Spectrum (a) represents 14.6 µmoles glyphosate/m manganite, 2 spectrum (b) 9.8 µmoles glyphosate/m manganite and (c) the difference between spectra a and b. The contribution from solution should increase in importance as the concentration of free ligand is increased in solution. In paper 6, the Cd(II)-glyphosate-manganite system was 28 studied using XPS. The concentration of free protonated glyphosate was 0.5 mM in these experiments, but no contribution from the solution could be detected in XPS spectra. For the pure glyphosate-manganite system, the concentrations of free ligand were similar to this for the lowest surface loading reported in paper 4 but increased to 1.5 mM for the higher surface loading. In these latter spectra, the contribution from solution was seen, which indicates that samples should be prepared so that the concentration of free ligand in solution is 0.5 mM. Despite the above mentioned problems, the low temperature XPS method can be used successfully when using mineral systems with low concentration of ligand in solution. This is clear from data for the Cd-glyphosate-manganite system presented in paper 6. By comparing solution data with XPS data of the adsorbing species, it can be seen that these two methods are in good agreement (Figure 42). X-ray Absorption Spectroscopy Extended X-ray Absorption Fine Structure (EXAFS) spectroscopy and X-ray Absorption Near Edge Structure (XANES) spectroscopy can be used to characterise the local chemical structure and geometry of a compound. A great advantage with these techniques over XRD (described below) is that they can be used to determine the local structure in non crystalline samples, as was done for the different Cd-glyphosate complexes in papers 5 and 6. XANES is sensitive to changes in oxidation state of atoms and coordination geometry around the atom, but this method has not been used in this thesis. In both XANES and EXAFS, information about local structure is obtained from the fine structure at the high energy side of the so called absorption edge. When atoms are irradiated by high energy x-ray photons, electrons are ejected out of the inner orbitals. If a range of xray energies are scanned and the absorbance plotted vs. energy, this ejection is seen as an absorption edge at the binding energy of the electron (Figure 15). On the high energy side of the edge, an oscillating fine structure can be seen in the absorption curve. This fine structure, extending from ~50 eV up to as much as ~1000 eV, is used in EXAFS. It is a result of interactions between the ejected electron wave and atoms present around the absorbing atom. By analyzing the oscillations in the spectrum, the local structure around the absorbing atom can be determined in the form of number of coordinating atoms around the absorber and 29 distances (up to ~5 Å) between the absorbing atom and its neighbours. XANES, on the other hand, uses the first region around the absorption edge, i.e. from < 10 eV below the absorption edge to ~50 eV after the edge55. Figure 15. Graph showing the absorption edge and the regions used in XANES and EXAFS 55 spectroscopies (modified from Brown et al. ). Since the fine structure in XANES and EXAFS is fairly weak, intense x-ray radiation is needed to obtain a good signal-to-noise ratio and, thus, synchrotron radiation is used. Synchrotron radiation is emitted by electrons (or positrons) that travel close to the speed of light in a curved pathway. It has a high degree of collimation and is very intense over a broad and continuous energy range. Synchrotron x-rays are monochromated and the intensity of the beam is measured before and after passage through a sample, which enables the amount of absorption to be determined. 30 a) c) b) d) Figure 16. EXAFS data after different steps of the data treatment process. a) Absorption edge fitted with a spline, b) normalisation of step height, c) isolated EXAFS oscillations, d) data in r-space 55 (modified from Brown et al. ). The data treatment in EXAFS starts by extracting the weak EXAFS oscillations, which only represent a few percentage of the total absorption. It is done by fitting the absorption edge with a spline and normalising the intensity of the step height (Figure 16). This isolates the fine structure oscillations at the high energy side of the edge. The pure oscillations can be seen as a sum of several sinusoidal signals, and to make the data more illustrative the signals are often Fourier transformed into r-space. This transformation results in a graph (radial distribution function), and the peaks in this graph represent the different atomic shells surrounding the absorbing atom (Figure 16d). The approximate distance from an atomic shell to the absorbing atom is shown on the x-axis and the height of the peak represents the number of absorbing atoms in that shell. When this graph is constructed, corrections for phase shift are not done and this results in a slight error in the distances showed on the x-axis of the graph. The data are analysed by curve fitting procedures and information about average coordination number (± 0.2 - 0.5 atoms), distance to the absorbing atom (± 0.01 Å), and internal structural disorder (Debye-Waller factor) is obtained55. In this fitting procedure the phase shift is taken into account and, thus, the fitting gives a correct determination of the distances. Theoretical 31 scattering paths, calculated using the FEFF program code56, form the basis of the algorithm used to fit of the data. These paths are calculated for a presumed model structure, e.g. a fragment of a crystal structure. Further reading about XAS can be found in the reviews by Jalilevhand57 and Brown et al.55. Infrared Spectroscopy Infrared (IR) light is used to study molecular vibrations and from these, deduce information about bonds within molecules. Different molecular vibrations absorb different wavelengths of light and, furthermore, the mass of the atoms participating in bonds and the strength of the bonds will influence the wavelength absorbed. IR spectroscopy can be used qualitatively to obtain a “fingerprint” of a substance48. Measured spectra are then compared to a library of previously obtained spectra and the substance identified. IR can also be used to monitor specific functional groups in a substance and how the structure of these groups changes as a consequence of different chemical environments, as was done for the glyphosate molecule in papers 5 and 6. For small simple molecules the molecular vibrations and resulting IR-spectra can be interpreted based on the molecular structure and symmetry. For more complex molecules this is difficult to do, instead the experimental data can be evaluated by comparing them to frequencies from ab initio calculations. But even without these calculations, structural information can be derived from the spectra. When a part of a molecule forms bonds to a surface or metal ion, the molecular vibrations will be affected and the changes are often visible in the IR spectra. By monitoring these changes, information about structure can be obtained. Attenuated Total Reflectance - Fourier Transformed Infrared spectroscopy (ATR-FTIR) was used in this thesis. The FTIR technique is applied in order to speed up the process of collecting spectra and to obtain a better signal to noise ratio compared to conventional IR spectroscopy. The ATR-cell is used since it allows spectra to be collected from strongly absorbing samples or compounds in a strongly absorbing matrix such as aqueous solutions and aqueous pastes. 32 (a) (b) (d) (c) (e) (f) Figure 17. Schematic diagram over a Michelson Interferometer (a) = fixed mirror (d) = monochromatic IR source (b) = moving mirror (e) = sample (f) = detector (c) = beam splitting mirror The FTIR-instrument uses a device known as a Michelson Interferometer (Figure 17). The incident IR beam is collimated by a concave mirror and divided into two beams at the beamsplitter. One of these beams is reflected by a stationary mirror while the other is directed to a mirror moving at constant speed. Reflections from the two beams are recombined at the beamsplitter and the difference in path length between them will result in constructive and destructive interference. In this manner a sinusoidal modulation in radiation intensity is obtained that is passed through the sample and focused onto a detector. The detector records intensity as a function of mirror position and this information is subsequently mathematically converted, by the use of Fourier transformation, to intensity as a function of wavenumber. 33 IR light Sample To detector Crystal Figure 18. Schematic drawing showing the function of the ATR cell used in FTIR measurements. The arrows represent the IR light, which is reflected several times within the diamond crystal on its way to the detector. The ATR cell used, in the present study, consists of a diamond crystal positioned horizontally in the sample compartment and the sample is applied on top of the crystal (Figure 18). The diamond crystal has a high refractive index, which enables IR light to be reflected at the interface between the crystal and the sample (as long as the sample has a lower refractive index). The geometry and size of the ATR crystal result in multiple reflections of the IR light within the crystal. At each reflection, an evanescent field penetrates a few microns into the sample. If the sample absorbs this evanescent wave, the intensity of the radiation is decreased at the absorbing wavelengths giving IR spectra similar to that of transmission measurements but from a path length that is only a few microns. X-ray Diffraction The structure and identity of crystalline samples can be determined using X-ray diffraction (XRD). The X-ray wavelength is of the same order of magnitude as the bond distances between atoms. As a result, the crystal array scatters the X-rays, and this scattering pattern is unique for the particular atomic arrangement in a sample. In this work, X-ray powder diffraction has mainly been used to check the outcome of the manganite synthesis but also to identify phase transformations at low pH (paper 3). Furthermore, single crystal X-ray diffraction was used in paper 5 to determine the crystal structure of a Cd(II)-glyphosate compound, (Cd9(PMG)6(H2O)12 × 6 H2O, colour Figure on p 68). The diffraction by crystalline material can be explained by the model by Bragg. In this model the crystal structure is pictured as consisting of planes that reflect X-rays. The diffraction is dependent on the angle of incidence of the X-rays and the distance between planes in the crystal according to eq. 7. 34 2 d sinθ =n λ Bragg’s law (7) where d is the distance between imaginary planes within the crystal, θ the angle of the incident radiation, the wavelength of the X-ray radiation and n the order of the reflection, commonly n is set to 1. Powder diffraction is often used to identify crystalline substances. Apart from identification of crystalline phases, the one-dimensional powder diffraction pattern can give additional information about the sample, e.g. small crystals can cause broadening of peaks, and the presence of amorphous material can result in very broad peaks, so called halos. Unfortunately, the identity of the amorphous phase can not be determined from the one dimensional diffraction patterns and the presence of amorphous material frequently cause diffuse features in the diffraction pattern. To obtain a complete range of reflections during powder diffraction analysis, the angle θ (theta) is altered stepwise or continuously. The sample is presented as a powder, which ideally will expose all possible orientations (and planes) of crystals in the sample. The obtained one dimensional diffraction pattern can be compared to reference patterns in databases and the identity of the sample determined. In some cases the crystal structure can also be derived from these patterns although single crystal diffraction is often preferred for structural determinations. In single crystal diffraction, a single crystal is synthesized and exposed to X-rays from different angles. The diffraction pattern is recorded in a sphere around the crystal and from this three dimensional pattern, the positions of different atoms within the crystal can be determined. Detailed descriptions of single crystal diffraction can be found in the books by Giacovazzo58 and Hammond59 and about powder diffraction in the book edited by Bish and Post60. 35 Microscopy Images of a sample or a sample surface can be obtained using different microscopy techniques. In conventional optical microscopy, the image is constructed by reflected or transmitted light, which the eye interprets as an image. Thus, what we see is reflected light, and the resolution is limited by the wavelength of light. Consequently, objects in the nm range cannot be imaged using visible light. However, an image of these objects can be constructed by using e.g. scattering of electrons. The three microscopic techniques described below are all used to image particles in the nanometre or sub nanometre range. In this work, Scanning Electron Microscopy (SEM) was used to determine morphological changes and formation of new phases during the dissolution of manganite (paper 3). High Resolution - Transmission Electron Microscopy (HR-TEM) was used to characterize the manganite crystals (Figure 5) for morphology and crystallinity (paper 1). Atomic Force Microscopy (AFM) was used to study morphological changes in dissolving manganite crystal in real-time (paper 3). In Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM) the images are constructed from scattered or transmitted electrons. The image in SEM is obtained by sweeping the surface with a finely focused beam of electrons. This focusing is obtained by a set of magnetic lenses to form a small spot size. For the highest resolution, a special source of electrons (field emission source) is used, resulting in a very small spot size and, thus, better resolution (1 - 5 nm vs. 10 - 50 nm) 61. The scanning is obtained using electromagnetic coils that deflect the beam in x and y directions. The detected image is a result of backscattered and secondary electrons from the sample. In TEM, on the other hand, the image originates from electrons transmitted through the sample and, as a consequence, the inner crystal array and any structural defects will be visible in TEM images62. By using an energy dispersive spectrometer (EDS) the elemental composition of the sample can be determined from x-rays emitted from the sample. When the sample surface is subjected to an electron beam, x-rays with characteristic energies are emitted from the sample, and these x-rays are used to determine the elemental composition of the sample. In AFM, images are obtained by scanning a sharp tip over a surface and recording the deflection of the cantilever (in “contact mode”). The tip is typically constructed from Si or Si3N4 and typically has a radius63 of 2 - 20 nm. The lower resolution limit is determined by 36 the radius of the tip and objects smaller than the tip will only show the geometry and structure of the tip. The software of the instrument controls the movement and the force of the tip against the surface. In order to obtain high resolution images, piezoelectric scanners are used, which give very accurate movements along the x-y axis of the surface. Height changes in the sample will cause deflection of the cantilever as the tip scans the surface. Consequently, height information can be obtained by measuring this deflection. This is done by focusing a laser beam on the back of the cantilever and as the cantilever moves up and down, the reflection of the laser beam will be altered. These changes are measured by a photodiode detector. Through a feedback loop the software maintains a constant deflection by increasing or decreasing the force of tip at the surface. Accordingly, the used force will be directly proportional to the height of the surface and plotting the force at the different points will produce a topographic image of the surface. Tapping Mode is another commonly used method of operation in AFM. In this mode the tip is oscillated near the cantilevers resonance frequency, and as the tip approaches the surface it will start to “tap” the surface. Height changes at the surface cause a decrease or increase in the amplitude of the oscillations and through a feedback loop constant amplitude of the oscillations is maintained. This enables the amplitude to be used as a measure for the topography of the surface, similar to contact mode. Tapping Mode causes less damage to the sample surface compared to contact mode and accordingly, it is often used for softer samples or samples that are easily moved out of place by the tip. SEM and AFM are two techniques that can give very similar information although the spatial resolution is often better with AFM. However, they both have drawbacks and thus are often used complementary to each other to obtain more information and for validation. An advantage with AFM is that morphological changes can be monitored in real time, but in SEM, a larger area can be imaged at once. Furthermore, a deeper field of view is obtained on rough surfaces using SEM. AFM, on the other hand, can give additional physical information of the surface, such as adhesion, friction, viscoelasticity and surface forces present. Also sometimes the height is more easily obtained from AFM than from SEM. A slope in SEM gives an increase of electrons but it might be difficult to say if the slope is directed upwards or downwards. Furthermore, in SEM, non conductive samples will build up charge at the surface due to the surface bombardment of electrons. This charge decreases the resolution in the 37 image and to obtain good electric conductivity, the surface is often covered with a thin carbon or metal film (e.g. gold or iridium) prior to the measurements. Equilibrium Analysis In order to describe and predict the outcome of a chemical reaction, the equilibrium constant for this particular reaction is important (together with knowledge of kinetic effects if such information is available). From thermodynamic models, knowledge about solution composition and concentrations at equilibrium can be obtained. Correspondingly, the exact composition of a solution before and after a reaction enables determination of equilibrium models64. Potentiometric Titrations The experimental setup for precise potentiometric titrations was designed and built at our department and is thoroughly described in the literature65,66,67. In short, it consists of a titration vessel immersed in a thermostated oil bath in order to maintain constant temperature throughout the titration. The titration vessel is purged with Ar gas, to exclude interference from CO2 and O2, and the addition of acid/base is accomplished by an automated burette. The potential is measured using a glass electrode, a “Wilhelm” type bridge and a Ag/AgCl reference electrode. Solutions are stirred with a magnetic stirrer and suspensions by the use of a small propeller. The use of propellers avoids possible grinding of the solid material by the stirrer. The potential of the cell; - Ag, AgCl(s) | ionic medium || equilibrium solution | glass electrode + is measured and the concentration of H+ is calculated using Nernst equation (eq. 8); [ ] E = E0 + g log H + + E j (8) where E is the potential in mV, E0 an apparatus constant determined separately, g=RTln10/nF=59.16 mV at 25°C and n=1, Ej is the liquid potential which occurs at the interface between the “Wilhelm” bridge and the equilibrium solution. 38 Batch Adsorption Experiments Batch experiments for equilibrium analyses were performed in test tubes, an appropriate amount of ligand and/or metal ions were added to the manganite suspension and pH adjusted. Care had to be taken not to lower the pH into the region of manganite dissolution. A combination electrode was used that was filled with 3 M KCl solution. This electrode was calibrated according to a concentration scale, obtaining -log[H+] instead of pH. A similar approach as in the potentiometric titrations was used for the calibration, but E0 was determined from the potential in two diluted H+ solutions with ionic medium. The test tubes were left on an end-over-end-rotator to equilibrate. Thereafter -log[H+] was measured, the solution and solid separated through centrifugation and filtration (0.22 µm filter), and analyses were performed on the solid and the solution. For adsorption and dissolution experiments, the concentration of metal ions or ligand in the solution was measured using Atomic Absorption Spectroscopy or scintillation counting as described in paper 6. The solid, in the form of wet paste, was analysed using the different spectroscopic and microscopic techniques, described earlier, to obtain information about quantities and molecular structures of the substances at the surface of the manganite. Electrophoretic Mobility Measurements The charge displayed at the surface of particles is important for their behaviour and adsorption properties. Unfortunately, the surface charge is very difficult to measure experimentally but it can be indirectly estimated using electrophoretic mobility techniques. In the electrophoretic mobility measurements, movement of the particles are measured and recalculated into a potential (zeta-potential) at the shear plane. In aqueous solutions, mineral surfaces are covered by water molecules. These molecules are electrostatically attracted towards the surface and exhibit somewhat different structural arrangement compared to the bulk. In-between this layer and the bulk solution, a shear plane develops. The potential at this plane reflects the surface charge and potential, but will usually have a smaller numeric value. However, the change of this potential with pH can be used to determine the iso electric point of a surface (pHiep), as was done for manganite in paper 2. 39 The potential at the shear plane was determined using the AcoustoSizer method developed by O’Brian et al.68. This method uses the electroacoustic effect to determine electrophoretic mobility of particles. When an alternating electric field is applied to a suspension of charged particles, the particles start to oscillate and thereby creating sound waves. The oscillation is influenced by the charge and size of the particles together with the frequency of the applied electric field. The sound is detected as pressure waves at the electrodes in the AcoustoSizer and recalculated into mobility, size and zeta potentials by the software68. The experimental setup is explained in O’Brian et al68. In short, it consists of a container equipped with a combination electrode for measuring pH, a probe for measuring conductivity, four ceramic rods for maintaining temperature, and a glass propeller. The electrodes that create the electric field are built into the container walls and are connected to two glass rods extending on either side of the container. The suspension can be titrated with acid or base to study changes in zeta potentials with pH. Modelling Different theoretical surface complexation models exist that aims to explain surface-ligand interactions. The most simple model for surface complexation is the constant capacitance model (CCM) 69 with the 1-pKa approach70,71, which also was the model used to describe the acid-base properties of manganite in paper 2. The CCM treats the mineral-solution interface as a parallel plate capacitor where the plates consist, on one side, of ionised surface groups and, on the other, of counter ions from the electrolyte. The capacitance between the plates is kept constant, and the model is valid for systems with small total charges or for systems with high concentrations of the electrolyte3 i.e. where the double layer is compressed. In this model, adsorbed ions are placed directly at the surface as inner sphere complexes (Figure 19). . 40 a surface + - b + - water molecule Figure 19, Schematic picture of a metal ion in a) inner sphere coordination and b) outer sphere coordination. Outer sphere complexes are usually due to electrostatic attraction between the surface and the adsorbate while covalent bonds between the ligand and different surface sites are formed in inner sphere complexes. The 1-pKa approach involves one deprotonation step at the surface according to eq. 9, where Me represents a metal centre at the metal hydroxide surface. MeOH2+½ MeOH-½ + H+ (9) An advantage of the 1-pKa model is that the logarithm of the equilibrium constant can be directly related to experimental data since it corresponds to the pHpzc of the surface. This also reduces the number of adjustable parameters in the modelling. If the surface species display similar behaviour as corresponding aqueous groups, a second deprotonation step (resulting in MeO-1½) should be approximately 10 pH units from the first72. Consequently, the second deprotonation reaction should be of very little importance in most modelling scenarios. There also exist more complex models that allow ionic strength effects to be modelled and counter ions to be placed further away from the surface. These models also enable placement of different charges of the adsorbate (if e.g. zwitterionic molecules or large organic molecules are modelled) at different distances from the surface depending on expected interactions. However, these models introduce more parameters that have to be adjusted to fit the data. Thus, although the chemical description of the system might be more illustrative with a more complex model, the use of a simple model with few parameters is often be preferred in many 41 systems. For more reading about surface complexation models, the paper by Lützenkirchen et al.72 is suggested. Experimental data were evaluated using the computer codes; LAKE73 and a modified version of Fiteql 2.074. The modified Fiteql version correctly accounts for dilution factors on solid concentrations and allow up to 9 surfaces to coexist in a system (J. Lützenkirchen pers. comm.). The computer codes minimise the error sum of squares, U= (Hcalculated-Hexperimental)2, to provide the best fit to the experimental data (H represent the total concentration of H+). Diagrams showing the distribution of different species were calculated and plotted using WinSGW75, a program package based on the SOLGASWATER algorithm76. Manganite Surface Processes The first part of the thesis work concerns the pure manganite mineral, its acid base and surface charge properties at circum neutral and alkaline pH together with its dissolution at acidic pH. The second part focuses on the surface complexation of manganite in the presence of glyphosate and Cd(II). Surface Charge and Acid-base Properties When manganite is suspended in an aqueous ionic medium, an initial pH of approximately 6 is obtained. Since this pH is below the isoelectric point (pHiep) at pH 8.2 (see below), the surface is positively charged and anions will be present at the surface to balance this charge. In the synthesis MnSO4 is used as starting material and, consequently, sulphate is the most likely anion at the surface. By analysing the sulphate concentration in the suspensions an estimate for the positive charge, i.e. the protonation of the surface, could be made and resulted in 0.55 µmol H+ /m2 manganite (4 % of crystallographic site concentration). The charge of the manganite surface was investigated in paper 2 using the AcoustoSizer technique and is illustrated in Figure 20 in the form of zeta potentials. 42 15 ζ-potential (mV) 10 5 pH 0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 -5 -10 -15 -20 Figure 20. Experimentally determined potentials at the shear plane of manganite at three different ionic strengths, = 10 mM NaCl, = 100 mM NaCl (paper 2). = 30 mM NaCl and No significant ionic strength dependence could be observed in the pH range 6 - 8. In other words, the data at the three ionic strengths more or less overlapped within the experimental uncertainties (Figure 20). The pHiep was recorded at pH 8.2 ± 0.1 in 10 mM NaCl with a small shift to higher pH at the higher ionic strengths. In the pH range above the pHiep, more pronounced ionic strength dependence was indicated suggesting surface interaction with Na+ from the ionic medium. This was supported by XPS data (paper 1 & 2), which confirmed the presence of Na+ at the surface at high pH (Figure 21). 0.08 0.06 Ratio OH/OH Ratio (Na-Cl)/Mn 2 0.04 0.02 1.6 1.2 0.8 0 5.5 6.5 7.5 8.5 9.5 5 10.5 pH 6 7 8 9 10 11 pH + Figure 21. The excess of Na at the manganite surface versus pH as determined by XPS. The Na present as NaCl is removed through the + Na-Cl subtraction. Na and Cl concentrations are normalised to Mn. I = 10 mM NaCl. Error 49 bars represent a 10 % error in atomic concentrations of the components (paper 1 & 2). Figure 22. The ratio of OH/O at the manganite surface versus pH as determined by XPS. I = 10 mM NaCl. Error bars represent a 10 % 49 error in atomic concentrations of the components (paper 1). 43 Furthermore, XPS data showed a slight increase in the ratio between hydroxide (OH) and oxide (O) components with increasing pH (Figure 22). This was interpreted as due to the deprotonation of surface OH2+½-groups forming OH-½ according to eq. 10 (paper 1). =MnOH2+½ =MnOH-½ + H+ (10) Potentiometric titrations of manganite suspensions (paper 2) also indicate a low buffer capacity ( 10-5 M g-1 pH-1) at pH < 7.5 (Figure 23). Furthermore, no or very little ionic strength dependence was observed in accordance to the zeta-potential measurements. At higher pH, the buffer capacity increased with increasing ionic strength, which, again, could be explained as Na+ interactions with the surface. 0.4 H (mM) 0.2 0 6 7 8 9 10 -0.2 -0.4 -0.6 pH 2 Figure 23. Potentiometric titration data for manganite suspensions (10 g/l, 51 m /g) with 100 mM NaCl (♦) and 30 mM NaCl ( ) ionic medium. The solid lines represent the fit to the measured data by the acid-base model presented in eq. 10 & 11 (paper 2). The knowledge of the above mentioned surface characteristics enabled the acid-base properties for the manganite to be modelled. Potentiometric data were modelled considering a surface complex between Na+ and the negatively charged sites on the surface. The 1-pK Model introduced by Bolt and van Riemsdijk70,71 was used, and the proton affinity constant was set equal to the pHiep obtained from the electroacoustic measurements at 10 mM NaCl. 44 This low ionic strength was chosen to avoid the slight Na+ effect seen in pHiep. The constant capacitance model applied contains only two adjustable parameters, the capacitance C and the affinity constant for the interaction between Na+ and singly coordinated oxygen atoms at the surface. The fit between the model and the data is shown in Figure 23 and the established equilibria in eq. 10 and 11, (C = 0.39 F/m2). =MnOH2+½ =MnOH-½ + H+ =MnOH2+½ + Na+ =MnOHNa+½ + H+ log β0 (intr.) = -8.20 (10) log β0 (intr.) = -9.64 (11) As expected, the Na+ complex is weak as indicated by the formation constant for =MnOHNa+½ from =MnOH-½ , which is: logK0 (intr.) = -1.44. The distribution of surface sites with pH can be seen in Figure 24. It can be noted that some 10 % of the surface site concentration is found as the Na+ - surface complex at pH 10 in 100 mM NaCl. This model is also supported by the XPS data showing a low affinity for Na+ at the surface. At pH =10 the Nasurface/Mn ratio is approximately 5 in 10 mM NaCl (Figure 21), which is fairly close to the percentage in the model (paper 2). Fi 0.6 +½ =MnOH2 0.5 -½ 0.4 =MnOH 0.3 0.2 +½ 0.1 =MnOHNa 100 mM 0 0.6 +½ =MnOH2 0.5 0.4 =MnOH -½ 0.3 0.2 0.1 +½ =MnOHNa 10 mM 0 6 7 8 pH 9 10 11 Figure 24. The distribution of surface sites at the manganite surface in 100 mM NaCl and 10 mM NaCl acquired from the acid-base model described in the text. The broken lines illustrate an extrapolation of the model (paper 2). 45 To our knowledge, only one previous publication has been made describing the acid-base characteristics of the manganite surface5. In that study the acid base properties were explained using a 2-pKa model§ with a capacitance of 1.02 ± 0.1 F/m2 in 10 mM ionic medium. The intrinsic protolysis constants were found to be pKa1 = 5.72 and pKa2 = 6.66, which resulted in pHpzc at pH 6.2. This is much lower than the pHpzc found for the manganite material used in this thesis. Additionally, a small number of studies exist describing the location of pHpzc/iep of the manganite surface (Table 2). Table 2. Reported values for pHpzc or pHiep Weaver et al. 77 Shaughnessy et al. 5 Xyla et al. Matocha et al. 79 This work, paper 2 78 pHpzc/ iep method MnOOH origin 5.4 titration natural 7.4 titration natural 6.2 titration synthetic 6.3 electrophoretic mobility synthetic 8.2 electrophoretic mobility synthetic The previous data in Table 2 or the acid base model could not be supported by the new work presented in this thesis, perhaps due to surface contaminations (e.g. sulphate) and differences in origin of manganite material. However, the agreement between the different techniques used in paper 2 gives us reason to believe that pHiep is situated at pH 8.2. § A model assuming the deprotonation of the surface to occur in two steps with the surface groups=OH2+, =OH and =O-. 46 Dissolution at Low pH Below pH 6, the dissolution of manganite becomes significant (Figure 25). During the dissolution, manganite is proposed to disproportionate to form Mn(II) ions in solution and a Mn(IV) oxide in the solid phase41 according to eq. 12. 2MnOOH(s) + 2H+ MnO2(s) + Mn2+ + 2H2O (12) 3 2+ [Mn ] (mM) 2.5 2 1.5 1 0.5 0 3.5 4 4.5 5 5.5 6 pH Figure 25. The dissolution of manganite at low pH, illustrated by the aqueous concentration of Mn(II) after 24 h of reaction time (paper 1). In paper 1, the remaining solid after the dissolution process was examined using XPS. It was very difficult to distinguish Mn(IV) in spectra that contained large quantities of Mn(III). However, in a manganite suspension at pH 1.5 (three days reaction time), the presence of Mn(IV) could be detected. Mn(IV) was observed as a change in peak shape of the Mn 2p peak, as well as an altered ratio between the hydroxide and oxide components of the O 1s spectra. Comparing the XPS spectra at pH 1.5 with a mixture of MnOOH and β-MnO2 (~50/50 % by volume), a similar change is observed in Mn 2p and O 1s spectra (Figure 26 & 27), which supports the formation of a Mn(IV) phase in manganite suspensions at low pH. 47 Intensity (counts/s) 35 x 10 30 2 30 25 a) 25 2 b) 20 20 15 15 10 10 5 660 40 Intensity (counts/s) x 10 35 30 x 10 650 640 660 650 640 2 30 x 10 2 25 c) d) 20 25 15 20 15 10 10 5 660 650 660 640 Binding energy (eV) 650 640 Binding energy (eV) Figure 26. Mn 2p XPS spectra of a) manganite, b) pyrolusite, c) 50/50 mixture manganite/pyrolusite and d) manganite paste at pH 1.5, showing the changes in peak shape as a consequence of MnO2 formation (paper 1). 2 45 x 10 2 x 10 Intensity (counts/s) 30 a) 40 b) 35 25 30 20 25 20 15 15 10 10 5 5 544 540 536 532 528 544 540 Binding energy (eV) 48 536 532 528 2 50 x 10 45 2 x 10 30 c) Intensity (counts/s) 40 d) O 25 35 OH 20 30 25 15 20 15 H2O 10 10 5 5 544 540 536 532 544 528 540 536 532 528 Binding energy (eV) Figure 27. O 1s XPS spectra of a) dry manganite, b) pyrolusite, c) mixture 50/50 manganite/pyrolusite and d) manganite paste at pH 1.5, showing the altered OH/O ratio as a consequence of formation of MnO2 (paper 1). In paper 3, the dissolution of manganite at low pH was examined further. Using literature equilibrium constants11 and the WinSGW computer code75,76, the pH dependency of the dissolution of manganite was calculated (Figure 1b). It was found that for a suspension with 10 g of manganite/dm3, the phase transformation should be completed below pH 4.8. Experiments were performed in the pH range 5.0 – 6.8 and although the dissolution/disproportionation proved to be a slow process, equilibrium seemed to be reached after one year. From solution data an equilibrium constant for eq. 12 was obtained (Figure 28). 0 2+ log [Mn ] log [Mn2+] = 7.40 + 2log[H+] -2 -4 -6 -8 5 6 + 7 8 - log [H ] + Figure 28. Plot of log Mn vs –log[H ] showing the alignment of experimental data to a theoretical line with a slope of 2. The line crosses the x-axis at 7.40, which represents the log K for eq. 12 in 100 mM ionic strength (paper 3). 49 The formation of pyrolusite (β-MnO2) was after nine months observed in XRD patterns of a sample at pH = 2.6, which suggests that eq. 12 can be written as: 2 γ-MnOOH + 2H+ β-MnO2 + Mn2+ + 2H2O log K0 = 7.61 ± 0.10 (13) where log K has been recalculated to zero ionic strength using the Davies equation. With time, this phase transformation resulted in a colour change from brown to black, and as a consequence, a colour gradient could be seen in the suspensions. The gradient represented the ratio between the brown manganite and the black pyrolusite and appeared to be relative to the initial amount of H+ added. The morphological changes as the crystals dissolve were investigated using AFM. In real-time AFM measurements, it was observed that the dissolution starts at the short edges of the manganite crystals (Figure 29). Over a period of 2 hours, this results in a shortening of the manganite crystals while their width is maintained. Figure 29. Real-time AFM images of dissolving manganite crystals, showing the shortening of the crystals while their width is maintained. The left picture shows a manganite crystal just before the addition of acid, and the right picture after 120 minutes of reaction time (paper 3). XRD patterns of partly dissolved manganite samples between pH 3.8 and 1.6 (reaction times of one week) displayed a small peak at ~4.10 Å, suggesting the presence of ramsdellite (MnO2) (Figure 30a & c). In SEM images of corresponding samples, needle shaped crystals with smaller dimensions compared to the manganite crystals (Figure 30b & d) appeared, 50 which probably represent ramsdellite. In samples at pH < 1.6, the presence of pyrolusite was indicated by an increasing reflection at 3.10 Å (Figure 30c & e). #2 a) 10 Intensity (Counts) X 100 41- 1379 MANGANITE b) 8 6 4 4.10 Å 2 10 20 30 40 50 2-Theta Angle (deg) #5 41- 1379 MANGANITE c) d) 7 6 5 4 3.10 Å Intensity (Counts) X 100 8 3 2 1 10 10 20 30 40 50 2-Theta Angle (deg) #6 41- 1379 MANGANITE e) f) 4 4.40 Å 6 5.65 Å Intensity (Counts) X 100 8 2 10 20 30 40 50 2-Theta Angle (deg) Scale bar Figure 30. XRD patterns and SEM images of samples at pH=3.8 (a & b), pH=1.6 (c, d & f) and pH=0.9 (e). The presence of manganite is indicated with lines in all XRD patterns and seen as the larger needles in the SEM images. b) and d) show ramsdellite as small needles, f) shows manganese chloride in the form of hexagons. Scale bar: 1.20 µm in b), 600 nm in d) 1.50 µm and in f) (paper 3). 51 However, the pyrolusite could not be seen in SEM images. The only “new” phase at pH 1.6 had a morphology similar to that of MnCl2 × 4H2O crystals (Figure 30f). This phase was considered to be a precipitate that fell out of solution during the drying procedure. Small amounts of MnCl2 were also seen in the XRD patterns (5.65 and 4.40 Å) at pH 0.9 (Figure 30e), which supports this hypothesis. The difficulties in recognising a pyrolusite phase in the SEM images can suggest that the pyrolusite phase is formed through a rearrangement of the ramsdellite or manganite phases. This hypothesis is supported by the complete rearrangement into pyrolusite in the pH = 2.6 sample described above, and is in agreement with previous observations38,80. Surface Complexes with Glyphosate and Cd(II) At pH < pHpzc, the manganite surface displays a net positive charge. In this pH range, manganite has an increased possibility to attract negatively charged ligands such as oxalate5, ascorbic acid7, phenolic compounds81, aminocarboxylates82 and arsenate83. The research presented in papers 4 and 6 shows that glyphosate forms bonds to the manganite surface in this pH region. Manganite-Glyphosate N-(phosphonomethyl)glycine (PMG, H3L) with the trivial name glyphosate is one of the most commonly used organophosphorus herbicides in the world today due to its low toxicity to mammals, fast degradation, and inactivation in soils84,85,86,87. In solid phase and in solution, glyphosate molecules display a zwitterionic structure so that a proton is dislocated from the phosphonate group to the amine group88,89 (Figure 31). The presence of three donor groups; the carboxylate, the amine and the phosphonate group, leads to interesting complexation behaviour with metal ions and mineral surfaces. 52 pKa3 = 10.1 - + pKa2 = 5.44 pKa1 = 2.22 H3L Figure 31. Protonated glyphosate molecule showing the zwitterionic structure and the three 90 deprotonation constants (I=0.1M) . From the left the structure represents; HO3P-CH2-NH2-CH2-CO2H 91 (colour version on p 68, modified from Sheals et al ). The inactivation of glyphosate in soils is mainly through adsorption onto mineral surfaces. Although this adsorption decreases the toxicity to plants and the mobility of the herbicide, the interactions can also affect the degradation of the glyphosate92,93,84. This can result in accumulated amounts of glyphosate in soils and perhaps also in ground waters over time. Previous studies show that the glyphosate mainly adsorbs through the phosphonate group84,94,95. Due to the preferred interactions between hard acids and hard bases it is not surprising that high affinities for Fe-94,95 and Al-minerals96 have been observed, since phosphonate is a relatively hard base. As a consequence of the phosphonate interactions, phosphate ions compete with glyphosate for binding sites in phosphate rich soils and thereby increase the mobility of glyphosate 84,97. Very little research has been performed concerning the adsorption of glyphosate onto manganese (hydr)oxides, even though they are a common component in natural soils24. Mn2+ ions added to soil have been reported to increase the degradation of glyphosate, while Fe3+ and Al3+-ions added to soil decrease the degradation92. An explanation for this was suggested to be due to changes in the bioavailability of glyphosate to micro-organisms. The difference between these ions raises the question about possible differences in the reactivity of glyphosate with the surfaces of FeOOH, AlOOH and MnOOH minerals. In paper 4 and 6, the adsorption of glyphosate onto manganite was examined and compared to the adsorption onto goethite. It was found that glyphosate adsorbed to the manganite surface over the entire pH range studied (Figure 32) but that the concentration of glyphosate decreased at the surface with increasing pH (paper 4 & 6). The glyphosate molecule displays a net negative charge above pH ~2 (pKa(H3L) = 2.22) 90 and, thus, is expected to be more 53 attracted to a positively charged surface. The continued adsorption at high pH to the net negatively charged surface could be a result of formation of covalent bonds between the ligand and the surface, and/or electrostatic effects due to the zwitterionic properties of the Atomic ratio N/Mn molecule, which remain until pH~10 (pKa(HL2-) =10.1)90. 0.08 0.06 0.04 0.02 0 6 7 8 pH 9 10 Figure 32. Amount of adsorbed glyphosate on the surface of manganite. 2.4 µmol glyphosate were 2 added per m of manganite. Data points represent XPS N 1s spectra normalised against Mn 2p. From the infrared spectra of glyphosate at the surface of manganite (Figure 33), it can be seen that the carboxylate and amine groups do not form bonds to the surface since they display spectra similar to aqueous glyphosate complexes. Thus glyphosate is likely bonded to the manganite surface through the phosphonate group. For glyphosate adsorbed onto goethite, the main information about the mineral – glyphosate interaction was derived from the phosphonate bands 94. Unfortunately, these bands overlap with strong manganite peaks. Thus, no direct evidence of bond formation between the phosphonate group and the surface could be extracted from infrared measurements. However, the structure of glyphosate at the manganite surface is here assumed to be similar to the structure of glyphosate at the surface of goethite. 54 Intensity a) b) c) 1800 1400 1000 / cm -1 Figure 33. FTIR spectra of glyphosate; a) glyphosate at the surface of manganite, (pH 7.1 and 2 2[PMG]tot= 9.8 µmol/m manganite), b) aqueous glyphosate in the form of HL and c) fully deprotonated 3aqueous glyphosate (L ). The lines illustrate the presence and absence of carboxylate hydrogen -1 bonded to the amine group of the glyphosate. In spectrum a) the vibrations below ~1200 cm were excluded due to the strong overlapping manganite peaks. From XPS and FTIR measurements, the presence of both protonated and deprotonated glyphosate was observed at the manganite surface throughout the pH range (Figure 33 and 34) studied. However, compared with the goethite surface, a higher percentage of the glyphosate 0.05 0.04 0.04 0.03 0.03 0.02 0.02 0.01 0.01 Atomic ratio N/Fe Atomic ratio N/Mn was protonated when bound to manganite, as shown in Figure 34. 0 0 6 7 8 pH 9 3 10 4 5 6 7 pH 8 9 Figure 34. XPS data showing the amount of protonated ( ) and deprotonated ( ) glyphosate at the 2 surface of: a) manganite with 2.4 µmol glyphosate added per m of manganite, b) goethite with 2.3 2 94 µmol glyphosate added per m of goethite (paper 4) . Furthermore, the manganite surface did not adsorb glyphosate from the solution as efficiently as the goethite surface at similar surface loadings. At pH 7, only ~50 % of the glyphosate was 55 adsorbed at the surface in the manganite-glyphosate system (unpublished data) compared to ~65 % in the goethite-glyphosate system (paper 4)94. Manganite-Glyphosate-Cadmium(II) In order to understand the chemistry at the mineral-water interface, it is very important to also have an understanding of the processes occurring in the bulk solution. All the reactions occurring at the interface are coupled with those in solution. Thus, the chemistry of the surface species will be influenced by the chemistry in solution, and vice versa. The speciation and equilibria of different Cd(II)-glyphosate complexes in solution are presented in paper 5. Stability constants were determined and agree well with previous formation constants in other ionic media98. The equilibrium analysis showed that Cd(II) exists in solution in the form of 1:1 and 1:2 Cd(II)-glyphosate complexes at pH > 3.0 (Figure 35). Fi 1 CdL- Cd2+ 0.8 CdL24- 0.6 0.4 CdHL CdOHL2- 0.2 0 3 4 5 6 7 8 9 10 pH + Figure 35. Distribution of Cd(II) in the glyphosate-Cd(II)-H system (4 mM glyphosate, 2 mM Cd(II), paper 5). FTIR spectra of the two dominating complexes (CdL- & CdL24-) show an increased separation between the symmetric and asymmetric carboxylate stretching vibrations of the glyphosate, which has been shown to indicate monodentate coordination to a metal ion99 (Figure 36). Furthermore, no broadening was observed in the asymmetric carboxylate peak (at ~1600 cm-1), which is interpreted as a lack of hydrogen bonding between the carboxylate and amine groups (Figure 36a). The changes in the phosphonate bands also indicate altered bond structure, suggesting that the phosphonate group is coordinated to the Cd(II) ion. 56 a) Intensity b) c) d) 1800 1400 1000 /cm -1 2- 3- Figure 36. FTIR spectra of a) partly protonated glyphosate HL b) fully deprotonated glyphosate, L c) 4CdL d) CdL2 . Showing the increased split between carboxylate stretching vibrations and the changes in phosphonate vibration in the Cd(II)-glyphosate complexes (paper 5). EXAFS data show that the coordination environment around the Cd(II) ion is very similar for the two dominating aqueous complexes (Figure 37). EXAFS data for both complexes could be fitted with a model including single scattering paths from Cd-O/N, Cd-C and Cd-P together with a multiple scattering path from Cd-C-O. In this model, the first shell consists of five oxygen atoms and one nitrogen atom, the second shell consists of three Cd-C and one Cd-P, and the third shell consists of a multiple scattering path corresponding to Cd-C-Odistal (Figure 38). In the Cd(II):glyphosate 1:2 complex, the second shell grew in magnitude due to the presence of the two glyphosate molecules in the complex. Furthermore, an increased disorder was observed in this complex, which is most probably caused by the large number of isomers expected in these types of complexes100. 57 Cd-O/N FT magnitude Odistal Cd-C Cd-P C Cd(II) N C Cd-C-O P C 0 2 4 6 R/Å - Figure 38. CdL complex. (paper 5) version on p 68) Figure. 37. Fourier transforms of EXAFS data 4for CdL (grey) and CdL2 (black) complexes. The fit to the data is represented by the broken lines (paper 5). 101 . (colour Thus, it was found that the structure of the 1:1 complex displays two chelating rings formed between the Cd(II) and the carboxylate, amine, and phosphonate groups in the glyphosate (Figure 38). The 1:2 complex had a similar chelating structure, but with two glyphosate molecules per Cd(II). This chelating-ring structure is the most common way in which glyphosate coordinates to metal ions (Cu(II)91,102, Pt(II)103, Co(III)100, Cr(III)104, La(III), Ce(III), Nd(III), Er(III)105, Pt(IV)106, V(IV)107, U(VI)108). However, glyphosate coordinates to some metals through a bidentate chelating structure where one of the functional groups in glyphosate does not interact with the metal. This has been described for crystallised complexes with Fe(III)109, Ba (II), Sr(II)110 or Na(I)111 where the amine group is uncoordinated, in alkaline solutions with Pt(II) where the carboxylate group is not complexed, and in acidic solutions with Pt(II) where the phosphonate group is free103. A third coordination mode, is where the metal ions form bonds only through the phosphonate group, as for complexes with Fe(III) in solution109 or Ca(II)112,113 in the solid phase. Eu(III)114 and Ag(I)111 display polymeric structures with glyphosate, where glyphosate molecules form bridges between metal ions through the phosphonate group on one side and the carboxylate group on the other. 58 In paper 6, the co-adsorption of Cd(II) and glyphosate was studied at pH 6 -10 and compared to adsorption in the corresponding binary systems (Cd(II)-manganite and glyphosatemanganite). In the binary Cd(II)-manganite system, Cd(II) adsorption increases with pH, and an adsorption edge is found at pH 8 - 9 (Figure 39a). This agrees well with previous adsorption studies of metals such as, Pb(II)79, Cd(II) 115, Zn(II)116,117 and Pu(VI)78 onto manganite where the metal ions display an increased adsorption at neutral to alkaline pH values. The increased adsorption also corresponds well with the development of a net negative charge at the surface above the pHpzc = 8.2. In the glyphosate-manganite system, glyphosate is adsorbed throughout the pH range studied (as was discussed earlier), and this is shown in Figure 39a. µmol ads/m2 8 a) 8 Cd PMG 6 b) Cd Cd & PMG 6 4 2 2 2 1 7 8 pH 9 10 6 7 8 pH PMG & Cd PMG 3 4 6 c) 4 9 10 6 7 8 9 pH Figure 39. Adsorption of glyphosate and Cd(II) on the surface of manganite as a function of pH. 2 [PMG]tot and/or [Cd(II)]tot = 14.6 µmol per m of manganite. Note the different scales on the y-axis (paper 6). In the presence of glyphosate, Cd(II) adsorption was increased at pH < pHpzc (Figure 39b). This could be a consequence of several processes: i) an increased electrostatic attraction for the surface due to the presence of adsorbed negatively charged glyphosate, ii) the formation of ternary surface complexes and/or iii) the formation of a Cd(II) containing precipitate. In the region pH > pHpzc, the adsorption of Cd(II) in the ternary system was less than the adsorption observed in the binary system. This can be explained by the competing formation of negatively charged Cd(II):glyphosate complexes in solution that display a decreased affinity for the surface. The presence of Cd(II) increased the adsorption of glyphosate over the entire pH range studied (Figure 39c), which also can be explained by the corresponding processes i) - iii) described above. 59 10 Using FTIR and XPS, the speciation of the glyphosate at the surface of manganite could be studied with respect to protonation and Cd(II) interactions. FTIR data have shown that both the formation of aqueous Cd(II)-glyphosate complexes and the protonation of glyphosate affect the glyphosate carboxylate frequencies. In XPS, the nitrogen spectra can indicate whether the amine group is protonated or not. In the binary glyphosate-manganite system, XPS results showed that the main part of the glyphosate was protonated throughout the pH range investigated (Figure 34). Thus, the complete absence of such protonation of glyphosate at the manganite surface indirectly suggests a Cd(II) interaction. FTIR spectra of the Cd(II)-glyphosate-manganite system at high pH have a similar appearance to the spectra of aqueous CdL- complex, indicating that glyphosate is coordinated to Cd(II) in these surface species (Figure 40a & f). In the low pH samples, the carboxylate bands were broadened compared to these bands at high pH (Figure 40a & c), which was also observed in the spectra of the aqueous manganite-glyphosate complexes. Subtraction of the spectrum of the binary glyphosate-manganite system (d) from the spectrum of the ternary Cd(II)-glyphosate-manganite system at a similar pH (c) reveals the presence of a glyphosate species coordinated to Cd(II) (Figure 40e & a). These results suggest the presence of a ternary Cd(II)-glyphosate-manganite complex at high pH and two coexisting complexes at low pH, which are believed to be a ternary Cd(II)-glyphosate-manganite and a binary glyphosatemanganite complex. 60 Abs Ratio N/Mn 0.1 0.05 a 0 b 6 c e 1500 1400 1300 9 10 Figure 40. (left) FTIR spectra of a) Cd(II)-glyphosate-manganite at pH 10, b) Cd(II)-glyphosate-manganite at pH 8.0, c) Cd(II)-glyphosate-manganite at pH 6.7, d) glyphosate-manganite at pH 6.5. e) subtraction c-d, f) aqueous CdL (paper 5 & 6). f 1600 8 pH Figure 41. (above) The speciation of the amine group obtained from XPS analyses. The filled triangles represent the protonated amine group + (>NH2 ) and the empty triangles represent the deprotonated amine group (>NH). The atomic percentages for N are normalised to the atomic percentage of Mn (paper 6). d 1700 7 1200 / cm-1 An XPS study was conducted to determine the protonation of the amine group of surface bound glyphosate in the ternary Cd(II)-glyphosate-manganite system, and the results are shown in Figure 41. At high pH, XPS spectra only showed the presence of deprotonated amine, which indirectly indicates glyphosate coordination to Cd(II). At pH < 8 the protonated form is detectable in the XPS spectra. This indicates that a fraction (~30 - 40 %) of the glyphosate is present in the form of a protonated surface species at low pH but not at high pH. These results support the FTIR results discussed above. XPS and batch adsorption experiments were performed to quantify the amount of Cd(II) and glyphosate at the surface of the manganite. The total Cd(II) adsorbed is shown in Figure 42, calculated both from solution and from XPS atomic ratios at the surface. These results show that at pH 6.7 the ratio between Cd(II) and glyphosate adsorbed is 1:1, but with increasing pH this ratio increases to about 2:1. 61 3 0.08 2 0.04 1 0 Atomic ratio µmol/m2 adsorbed 0.12 0.00 6 7 8 pH 9 10 11 Figure 42. The concentration of Cd(II) and glyphosate at the surface of manganite. represent Cd concentration from XPS (normalised as Cd/Mn), × represent adsorbed Cd from solution data, represent N concentrations from XPS i.e. the glyphosate concentration (normalised as N/Mn) and represent adsorbed glyphosate from solution data. A error < 10% can exist in XPS atomic 49 2 concentrations . A total concentration 4.8 µmol of both glyphosate and Cd(II) was added per m of manganite, which represents 37 % of the crystallographic site density (paper 6). From batch adsorption, XPS and FTIR data, the presence of a Cd(II)-glyphosate complex was observed at the surface of manganite throughout the pH range studied. An additional protonated glyphosate complex was detected at low pH. Considering the 1:1 Cd(II)glyphosate ratio at low pH and the 2:1 ratio at high pH, the presence of a third surface complex must be assumed. At low pH, the concentration of this third complex would be similar to the concentration of the protonated glyphosate complex (~30 - 40 % of the Cd(II)) in order to generate the 1:1 Cd(II):glyphosate ratio at the surface. To resolve this, the speciation from the perspective of Cd(II) was studied using EXAFS spectroscopy. EXAFS data for the surface species in the ternary Cd(II)-glyphosate-manganite system were compared with EXAFS data for the aqueous Cd(II)-glyphosate complexes described above (and in paper 5). The general appearance of these datasets was very similar but with a significantly decreased second shell in the surface species (Figure 43). 62 a b c d 0 1 2 3 4 R/Å 5 6 7 8 Figure 43. Fourier transforms of a) Cd(II)-glyphosate-manganite at pH 6.7, b) Cd(II)-glyphosatemanganite at pH 8.0, c) Cd(II)-glyphosate-manganite at pH 10, and d) CdL (aq). The broken lines are the results from the R-space fits. The above hypothesis of the coexistence of Cd(II) containing ternary and binary surface complexes implies that on average less than one glyphosate per cadmium is available for coordination. This explains the decreased amplitude for the second shell compared to solution data. The structure of the complexes was modelled assuming six O/N in the first shell and a second shell with a combination of Cd-Mn back-scattering and Cd-C + Cd-P back-scattering. The best fit to data was obtained for a scenario where 40 % of the Cd(II) was in the form of a binary Cd(II)-manganite surface complex, and the remaining 60 % was in the form of a ternary Cd(II)-glyphosate-manganite complex in which the Cd(II) is not directly bound to the surface. The absence of a direct interaction between Cd(II) and the surface suggests that the ternary complex is of type B (surface-glyphosate-Cd(II)) rather than of type A (surfaceCd(II)-glyphosate). Furthermore, the presence of P in the second coordination shell indicates that the ternary complex displays a structure similar to the aqueous Cd(II)-glyphosate complexes in that the two chelating rings between Cd(II) and glyphosate are intact. This mode of surface coordination is also observed in Cu(II)-glyphosate-surface complexes on aluminium (hydr)oxides118. The EXAFS results were also in good agreement with XPS data, which indicated the presence of 30 - 40 % of the binary glyphosate-manganite complex. The absence of Cd-Cd distances rule out the possibility of surface precipitation and the Cd-Mn 63 distances obtained suggest edge sharing between Cd and Mn octahedral in the binary Cd(II)surface complex, similar to what has been observed in previous studies115. From the combined solution-, XPS-, FTIR- and EXAFS data, the presence of three different surface complexes was established. At low pH approximately 60 % of glyphosate and Cd(II) is present at the surface in the form of a ternary surface complex and 40 % in the form of binary surface complexes (Figure 44). Calculations of the solution composition showed that a 1:1 ratio between Cd(II) and glyphosate exists both at the surface and in solution in the lower pH range. c) a) b) Hydrogen bonded Figure 44. Possible structures for the identified surface complexes at the manganite surface (not to scale); a) type B Cd(II)-glyphosate-manganite ternary surface complex, b) binary Cd(II)-surface complex and c) binary glyphosate-surface complex (with a structure similar to the one described on 94 goethite by Sheals et al ). Hydrogen atoms have been omitted for clarity (paper 6). With increasing pH the binary glyphosate-manganite complex desorbs from the surface (complex c in Figure 44) while the concentration of e.g. the CdL24- complex in solution 64 increases. Apart from electrostatic effects between the complexes in solution and the surface, these solution complexes successfully compete with the surface for the glyphosate, resulting in a Cd(II):glyphosate ratio of 2:1 at the surface and 1:2 in solution. The structure of the ternary complex can be compared to results from previous studies, e.g. Cu(II)-glyphosate at the surface of goethite90, gibbsite and bayerite118. In the Cu(II)glyphosate-goethite system, the structure of the ternary complex changed with pH. At low pH, a type B ternary complex was formed and at high pH it changed into to a type A ternary complex90. The coordination around the metal ion was similar to the aqueous 1:1 Cu(II)glyphosate complex in that the bonds between the amine and carboxylate group were kept intact at the surface of goethite. However, the chelating ring formed by bonds to the amine and the phosphonate opened due to the phosphonate interaction with the surface. In the Cu(II)-glyphosate-gibbsite and Cu(II)-glyphosate-bayerite systems, the structure of the ternary complex was of type B and did not change with pH. Both of the chelating rings were kept intact despite the phosphonate interaction with the surface118 similar to Cd(II)-glyphosate complexes on manganite. Summary By combining XPS, potentiometry and electrophoretic mobility measurements, a simple 1-pKa acid-base model for the manganite surface was established that accounts for a weak Na+-surface interaction. Furthermore, this acid-base model explains the low proton buffering capacity of manganite at pH > 6, the net positive charge at pH < 8.2, and the net negative charge at pH > 8.2 (pHiep = 8.2). The dissolution of manganite at pH < 6 was studied at equilibrium (after one year) and in nonequilibrated samples. After one year of equilibration, stable [H+] readings were obtained. Together with XRD data, this enabled an equilibrium constant for the manganite disproportion into pyrolusite (β-MnO2) and Mn(II)-ions to be established. Non-equilibrium microscopy measurements indicated that the dissolution initially starts at the apices of the crystals. Furthermore, XRD measurements indicated that ramsdellite (MnO2) can be formed in non-equilibrium processes, but that it rearranges into pyrolusite with time. 65 Increased Cd(II) adsorption at a net negative manganite surface (pH > 8.2) was observed in wet chemical adsorption studies. EXAFS revealed that this adsorbed Cd(II) formed an inner sphere complex by edge sharing between Mn and Cd octahedra. The adsorption of glyphosate, which displays an overall negative charge at pH > 2, took place over the entire pH range although it decreased with increasing pH. This can be explained by the strong interaction between the phosphonate group and the surface, in combination with the zwitterionic properties of the molecule. The structure of this binary complex was explained using FTIR and XPS data. It was described as an inner sphere complex where the phosphonate group in glyphosate forms bonds to the surface. This shows a preference for an interaction between the surface and the phosphonate group, instead of the carboxylate or amine groups. From XPS data it was observed that the largest fraction of the glyphosate-manganite surface complexes displayed a protonated amine group up to pH 10. In the presence of both Cd(II) and glyphosate, the overall adsorption at the manganite surface was increased. At low pH, Cd(II) adsorption was increased in the presence of glyphosate. This was explained by the formation of a ternary surface complex and altered electrostatics at the mineral interface. At high pH, Cd(II) adsorption was decreased in the presence of glyphosate. This is thought to be due to electrostatic repulsion between negatively charged species in solution and a net negatively charged surface, as well as the formation of aqueous complexes in solution that compete with the surface for Cd(II) (and glyphosate). The presence of Cd(II) increased the concentration of surface-bound glyphosate throughout the pH range which was explained by formation of ternary surface complexes and altered electrostatics at the interface. By combining several techniques, the structure of the formed complexes was derived. XPS and FTIR were used to study changes in the glyphosate, and EXAFS to study the Cd(II) coordination environment. The data indicated the formation of a ternary type B surfaceglyphosate-Cd(II) complex throughout the pH range studied. EXAFS studies showed that the chelating structure between Cd(II) and glyphosate, as seen in aqueous Cd(II)-glyphosate complexes, was maintained at the surface. This structure involves the carboxylate, amine and phosphonate groups coordinating to the Cd(II)-ion. The Cd(II)-glyphosate complex was bound to the surface through the phosphonate group of the glyphosate molecule. This research shows the importance of understanding both the chemistry in solution and at the mineral surface in order to describe surface processes. Furthermore, it seems likely that not 66 only aluminium (hydr)oxides and iron (hydr)oxides, but also manganese (hydr)oxides have the possibility to influence glyphosate mobility and degradation. Future Research • Development of thermodynamic surface complexation models that give a more complete understanding of the interactions between the manganite surface, glyphosate and Cd(II). • Preliminary experiments have shown that although glyphosate inhibited Cu(II)hydroxide precipitation on the goethite surface, it did not do so at the manganite surface. The nature of this precipitate (polymerisation) would be interesting to study and compare with Cu(II)-glyphosate-goethite complexes. Also, can this precipitation be avoided under different experimental conditions? • Further development of the low temperature XPS measurements, with respect to sample pre-treatment, to avoid contributions from the solution in XPS spectra and to minimise evaporation. 67 - + Figure 31 (page 53 in the thesis). Protonated glyphosate molecule showing the zwitterionic structure. Red atoms are oxygen, blue nitrogen, black carbon, white hydrogen and pink phosphorous 91 (modified from Sheals et al ). Figure 6 (in paper 5). The molecular structure of Cd9(PMG)6(H2O)12 × 6H2O. The coordination geometries of the cadmium ions and the phosphonate groups are represented by grey octahedra and pink tetrahedra, respectively. Hydrogens have been omitted for clarity. The remaining atoms were colour coded; blue for N, red for O and black for C. - Figure 38 (page 58 in the thesis). CdL complex. Red atoms represent oxygen, blue nitrogen, black carbon, pink phosphorous and grey cadmium(II), hydrogens have been omitted (also Figure 4 in paper 5). 68 References 1 Stumm W, Morgan JJ, Aquatic Chemistry, 3rd Edition, John Wiley & Sons, New York, 1996 Banwarth, S. A.Aqueous Speciation at the Interface Between Geological Solids and Groundwaters, Chapter 7 in Modelling in Aquatic Chemistry, Nuclear Energy Agency, Organisation for Economic Co-operation and Development (OECD), Paris, France, 1997 3 Stumm, Werner. Chemistry of the solid-water interface – Processes at the mineral-water and particle-water interface in natural systems, John Wiley & Sons, 1992, 4 Brown, G. E. Jr. Henrich, V.E. Casey, W. H. Clark, D. L. Eggleston, C. Felmy, A. Goodman, D.W. Grätzel, M. Maciel, G. McCarthy, M. I. Nealson, K. H. Sverjensky, D. A. Toney, M. F. Zachara, J. M. Metal oxide surfaces and their interactions with aqueous solutions and microbial organisms, Chem. Rev. 1999, 99, 77-174 5 Suter, D. Banwart, S. Stumm, W. Dissolution of Hydrous Iron(III) Oxides by Reductive Mechanisms, Langmuir, 1991, 7, 809-813 6 Xyla AG, Sulzberger B, Luther GW III, Hering JG, Van Cappellen P, Stumm W, Reductive Dissolution of Manganese (III,IV) (Hydr)oxides by oxalate: The effect of pH and Light, Langmuir, 1992, 8, 95-103 7 Jun Y-S, Martin ST, Microscopic Observations of Reductive manganite Dissolution under Oxic Conditions, Environ. Sci. Technol., 2003,37,2363-2370 8 Varentsov IM, Grassely GY, Geology and Geochemistry of Manganese, Volume 1, E. Schweizerbart’she Verlagsbuchhandlung, Stuttgart, 1980 9 Pontér, C. Ingri, J. Boström, K. Geochemistry of manganese in the Kalix River, northern Sweden, Geochim. Cosmochim. Acta, 1992, 56, 1485-1494 10 Crerar. D. A. Cormick, R. K. Barnes, H. L. Geochemistry of manganese: an overview, in Geology and Geochemistry of Manganese, Volume 1, Ed Varentsov IM, Grassely GY, E. Schweizerbart’she Verlagsbuchhandlung, Stuttgart, 1980, 293-334 11 Bricker O, Some stability relations in the system Mn-O2-H2O at 25° and one atmosphere total pressure, Am. Mineral., 1965, 50, 1296-1354 12 Lindsay, W.L. Chemical Equilibria in Soils, John Wiley and Sons, New York, 1979 13 Biedermann, G. Palombari, R. On the hydrolysis of the Manganese(III) Ion, Acta Chem. Scand. 1978, A32, 381-390 14 Kostka, J. E. Luther, G. W. III: Nealson, K. H. Chemical and biological reduction of Mn(III)-pyrophosphate complexes: Potential importance of dissolved Mn(III) as an environmental oxidant, Geochim. Cosmochim. Acta, 1995, 59, 885-894 15 Klewicki, J. K. Morgan, J. J. Kinetic behaviour of Mn(III) complexes of pyrophosphate, EDTA, and Citrate, Environ. Sci. Technol. 1998, 32, 2916-2922 16 Hem JD, Lind CJ, Nonequilibrium models for predicting forms of precipitated manganese oxides, Geochim. Cosmochim. Acta, 1983, 47, 2037-2046 17 Hastings D, Emerson S, Oxidation of manganese by spores of a marine bacillus: Kinetic and thermodynamic considerations, Geochim. Comsmochim. Acta, 1986, 50, 1819-1824 18 Tipping E, Thompson DW, Davison W, Oxidation products of Mn(II) in lake waters, Chem. Geol. 1984, 44, 359-383 19 Murray JW, Dillard JG, Giovanoli R, Moers H, Stumm W, Oxidation of Mn(II): Initial mineralogy , oxidation state and ageing, 1985, Geochim. Cosmochim. Acta., 49, 463-470 20 Giovanoli R, On Natural and Synthetic Manganese Nodules, in Geology and Geochemistry of Manganese, Volume 1, Ed Varentsov IM, Grassely GY, E. Schweizerbart’she Verlagsbuchhandlung, Stuttgart, 1980, 159202 21 Kessick MA, Morgan JJ, Mechanisms of Autooxidation of manganese in Aqueous Soultion, Environ. Sci. Technol. 1975, 9(2) 157-159 22 Lind CJ, Hausmannite (Mn3O4) Conversion to Manganite (γ-MnOOH) in Dilute Oxalate Solution, Environ. Sci. Technol. 1988, 22, 62-70 23 Mandernack KW, Post J, Bradley MT, Manganese mineral formation by bacterial spores of the marine Bacillus, strain SG-1: Evidence for the direct oxidation of Mn(II) to Mn(IV), Geochimica. Cosmochim. Acta., 1995, 59(21) 4393-4408 24 Post JE, Manganese oxide minerals: Crystal structures and economic and environmental significance, 1999, Proc. Natl. Acad. Sci. USA, Vol 96, 3447-3453 25 De Vitre RR, Davison W, Manganese particles in freshwaters, Chapter 7 in Environmental Particles Vol 2, Ed Buffle J, van Leeuwen HP, 1993, Lewis Publishers, Boca Raton 2 69 26 Dong D, hua X, Li Y, Zhang J, Yan D, Cd adsorption properties of components in different freshwater surface coatings: The important Role of Ferromanganese Oxides, Environ. Sci. Technol., 2003, 37, 4106-4112 27 Taillefert M, MacGregor BJ, Gaillard JF, Lienemann CP, Perret D, Stahl DA, Evidence for a dynamic cycle between Mn and Co in the water column of a stratified lake, Environ. Sci. Technol., 2002, 36, 468-476 28 Pettersson, U. T. Ingri, J. The geochemistry of Co and Cu in the Kafue River as it drains the Copperbelt mining area, Zambia, Chem. Geol. 2001, 177, 399-414 29 Lehninger AL, Nelson DL, Cox MM, Principles of Biochemistry, 2nd Edition, Worth Publishers, New York, 1993 30 Diem D, Stumm W, Is dissolved Mn2+ being oxidised by O2 in the absence of Mn-bacteria or surface catalysts?, Geochim. Cosmochimica. Acta., 1984, 48, 1571-1573 31 Pontér, C. Ingri, J. Burman, J-O. Boström, K. Temporal variations in dissolved and suspended iron and manganese in the Kalix river, northern Sweden, Chem. Geol. 1990, 81, 121-131 32 Marklund, K. Nationalencyclopedin, Band 13, Bokförlaget Bra Böcker, Höganäs, Sweden, 1994, p40-42, 33 Dachs von H, Neutronen- und Röntgenundersuchungen am Manganit, MnOOH, Zeitschrift für Kristallographie, 1963, 118, 303-326 34 Buerger MJ, The symmetry and crystal structure of manganite, Mn(OH)O., Zeitschrift für Kristallographie, 1936, 95, 163-174 35 Chabre Y, Pannetier J, Structural and electrochemical properties of the proton/γ-MnO2 System, Prog. Solid St. Chem., 1995, 23, 1-130 36 Maslen, B. E. N. Streltsov, V. A. Streltsova, N. R. Electron Density and Optical Anisotropy in Rhombohedral Carbonates. III: Synchrotron X-ray Studies of CaCO3, MgCO3 and MnCO3, Acta Cryst, 1995, B51, 929-939 37 Stumm W, Giovanoli R, On the Nature of Particulate Manganese in Simulated Lake Waters, Chimica, 1976, 30(9) 423-425 38 Kohler T, Armbruster T, Libowitzky, E. Hydrogen Bonding and Jahn-Teller Distortion in groutite, αMnOOH, and manganite, γ-MnOOH, and Their Relation to the manganese Dioxides Ramsdellite and Pyrolusite, J. Solid State Chem. 1997, 133, 486-500 39 Ramstedt M, Shchukarev AV, Sjöberg S, Characterization of hydrous manganite (γ-MnOOH) surfaces – an XPS study, Surf. Interface Anal., 2002, 34, 632-636 40 Bochatay L, Persson P, Sjöberg S, Metal Ion Coordination at the Water-Manganite (γ-MnOOH) Interface. I. An EXAFS study of cadmium(II), J. Colloid Interface Sci., 2000, 229, 584-592 41 Klewicki, J. K. Morgan, J. J. Dissolution of β-MnOOH particles by ligands: Pyrophosphate, etylenediaminetetraacetate, and citrate, Geochim. Cosmochim. Acta, 1999, 63, 3017-3024 42 Hiemstra, T. Riemsdijk, van W. H. A Surface Structural Approach to Ion Adsorption: The Charge Distribution (CD) Model, J. Colloid. Interface Sci. 1996, 179, 488-508 43 Pauling L, The principles determining the structure of complex ionic crystals, J. Am. Chem. Soc., 1929, 51, 1010-1026 44 Rakovan, John. Becker, Udo. Hochella, Michael F Jr. 1999, Aspects of goethite surface microtopography, structure, chemistry, and reactivity, Am. Mineral. 84, 884-894 45 Berry, L. G. Mason, B. Mineralogy – Concepts, descriptions, determinations, 2nd Edition, Ed Dietrich, R.V. W. H. Freeman and Company, San Fransisco, 1983, 318-319 46 Giovanoli, R. Leuenberger, U. Über die Oxydation von Manganoxidhydroxid, Helv. Chim. Acta, 1969, 52, 2333-2347 47 Brunauer, S., Emmet, P.H., Teller, E. Adsorption of Gases in Multimolecular Layers , J. Am. Chem. Soc., 1938, 60, 309-319 48 Skog, D. A. Leary, J. J. Principles of Instrumental Analysis, 4th Ed, Hardcourt Brace College Publishers, Fort Worth, USA, 1992 49 Ratner, B. D. Castner, D. G. Electron Spectroscopy for Chemical Analysis, Chapter 3 in Surface Analyses – The Principal Techniques, Ed. Vickerman, Wiley & Sons, 1997 50 NIST Standard Reference Database 82, NIST Electron Effective-Attenuation-Length Database Version 1.0, 2001, U.S. Department of Commerce, Gaithersburg, Maryland, USA 51 Hochella, M. F. Jr. Auger Electron and X-ray Photoelectron Spectroscopies, Chapter 13 in Spectroscopic Methods in Mineralogy and Geology, Volume 18 in Reviews in Mineralogy, , Ed Hawthorne, F. C.,Mineralogical Society of America, Washington D.C. 1988, 573-637 52 Ilton, E. S. Moses, C. O. Veblen, D. R. Using X-ray photoelectron spectroscopy to discriminate among different sorption sites of mica: with implications for heterogeneous reduction of chromate at the mica-water interface, Geochim. Cosmochim. Acta, 2000, 64, 1437-1450 53 Paterson, E. Swaffield, R. X-ray photoelectron Spectroscopy, Chapter 6 in Clay mineralogy: Spectroscopic and Chemical Determinative Methods, Ed. Wilson, M. J. Chapman & Hall, London, 1994, 226-259 70 54 Mitchell, D.F. Clark, K.B. Bardwell, J.A. Lennard, W.N. Massoumi, G.R. Mitchell, I.V., Film Thickness Measurements of SiO2 by XPS, Surf. Interface Anal. 1994, 21, 44-50 55 Brown, G. E. Calas, G. Jr. Waychunas, G. A. Petiau, J. X-ray absorption spectroscopy and its application in mineralogy and geochemistry, Chapter 11 in Spectroscopic Methods in Mineralogy and Geology, Volume 18 in Reviews in mineralogy, , Ed Hawthorne, F. C.,Mineralogical Society of America, Washington D.C. 1988, 430512 56 Ankudinov, A.L., Ravel, B., Rehr, J.J., and Conradson, S.D. Real-space multiple-scattering calculation and interpretation of x-ray-absorption near-edge structure, Phys. Rev. B 1998, 58, 7565-7576. 57 Jalilehvand, F. Structure of Hydrated Ions and Cyanide Complexes by X-ray Absorption Spectroscopy, PhD Thesis, Inorganic Chemistry, Royal Institute of Technology, Stockholm, 2000, ISBN 91-7170-561-9 58 Giacovazzo, C. Fundamentals of Crystallography, Oxford Press, New York, 1992 59 Hammond, C. The basics of Crystallography and Diffraction, Oxford University Press, New York, 2001 60 Bish, D. L. Post, J. E. Modern Powder Diffraction, Volume 20 in Reviews of Mineralogy, Mineralogical Society of America, Washington D.C. 1989 61 Goldstein, J. I. Lyman, C. E. Newbury, D.E. Lifshin, E. Echlin, P, Sawyer, L. Joy, D. C. Michael, J. R. Scanning Electron Microscopy and X-ray Microanalysis, Kluwer Academic/Plenum Publishers, New York, 2003 62 Buseck, P. R. Minerals and Reactions at the atomic scale: transmission electron microscopy, Volume 27 in Reviews in Mineralogy, Mineralogical Society of America, Washington D.C. 1992 63 Russell, P. Batchelor, D. Thornton, J. SEM and AFM: Complementary Techniques for High Resolution Surface Investigations, Digital Instruments, Carlifornia, USA. 64 Öhman, L-O. Experimental determination of stability constants of aqueous complexes, Chem. Geol., 1998, 151, 41-50 65 Ginstrup, O. Experimental and Computational Methods for Studying Multicomponent Equilibria. II: An Automated System for Precision emf Titrations, Chem. Instr. 1973, 4, 141-155 66 Sjöberg, S. Lövgren, L. The application of potentiometric techniques to study complexation reactions at the mineral/water interface, Aq. Sci., 1993, 55, 324-335 67 Forsling, W. Hietanen, S. Sillén, L. G. Studies of the hydrolysis of metal ions. III. The hydrolysis of the mercury(I) ion, Hg22+, Acta Chem Scand. 1952, 6, 901-909 68 O’Brien, R.W. Cannon, D.W. Rowlands, W.N. Electroacoustic Determination of Particle Size and Zeta Potential, J. Colloid Interface Sci., 1995, 173, 406-418 69 Schindler, P. W. Gamsjäger, H. Acid-base reactions of the TiO2 (Anatase) – water interface and the point of zero charge of TiO2 suspensions, Kolloid-Z. U. Z. Polymere, 1972, 250, 759-765 70 Van Riemsdijk, W. H., Bolt, G. H., Koopal, L. K. and Blaakmeer, J Electrolyte Adsorption on Heterogeneous Surfaces – Adsorption Models J. Coll. Interf. Sci 1986, 109, 219-228 71 Bolt, G. H. Soil Chemistry, B. Physico-Chemical Models, Elsevier, Amsterdam, 1982 72 Lützenkirchen, J. Boily, J. F. Sjöberg, S. Current applications of and future requirements for surface complexation models, Current topics in Colloid and Interface Science, 2002, 5, 157-190 73 Ingri, N., Andersson, I., Pettersson, L., Yagasaki, A., Andersson L., and Holmström K. LAKE- A Program System for Equiliibrium Analytical Treatment of Multimethod Data, Especially Combined Potentiometry and Nuclear Magnetic Resonance Data, Acta Chem. Scand. 1996, 50, 717-734. 74 Westall, J. C. FITEQL: A Computer Program for Determination of Chemical Equilibrium Constants from Experimental Data, Version 2.0, Report 82-02, Department of Chemistry, Oregon State University, Corvallis, OR, USA, 1982 75 http://www.chem.umu.se/dep/inorgchem/samarbeta/WinSGW_eng.stm 76 Eriksson, G. An Algorithm for the Computation of Aqueous Multicomponent, Multiphase Equilibria, Anal. Chim. Acta 1979, 112, 375-383. 77 Weaver, R. M. Hochella, M. F. Ilton, E. S. Dynamic processes occurring at the CrIIIaq-manganite (γ-MnOOH) interface: Simultaneous adsorption, microprecipitation, oxidation/reduction, and dissolution Geochim. Cosmochim. Acta. 2002, 66(23), 4119-4132 78 Shaughnessy DA, Nitsche H, Booth CH, Shuh DK, Waychunas GA, Wilson RE, Gill H, Cantrell KJ, Serne RJ, Molecular Interfacial Reactions between Pu(VI) and Manganese Oxide Minerals Manganite and hausmannite, Environ. Sci. Technol., 2003, 37, 3367-3374 79 Matocha CJ, Elzinga EJ, Sparks DL, Reactivity of Pb(II) at the Mn(III,IV) (Oxyhydr)Oxide-Water Interface, Environ. Sci. Technol., 2001, 35, 2967-2972 80 Yamada, N. Ohmasa, M. 1986, Textures in natural pyrolusite, β-MnO2, examined by 1 MV HRTEM, Acta Cryst. B42, 58-61 81 Zhang H, Huang C, Oxidative Transformation of Triclosan and Chlorophene by Manganese Oxides, Environ. Sci. Technol., 2003, 37, 2421-2430 71 82 Mcardell CS, Stone AT, Tian J, Reaction of EDTA and realted aminocarboxylate chelating agents with CoIIIOOH (Heterogenite) and MnIIIOOH (Manganite), Environ. Sci. Technol. 1998, 32, 2923-2930 83 Chiu van Q, Hering JG, Arsenic Adsorption and Oxidation at Manganite Surfaces 1. Method for Simultaneous Determination of Adsorbed and Dissolved Arsenic Species, Environ. Sci. Technol., 2000, 34, 2029-2034 84 Gimsing, A. L. Borggaard, O. K. Sestoft, P. Modeling the Kinetics of the Competitive Adsorption and Desorption of Glyphosate and Phosphate on Goethite and Gibbsite and in soils, Environ. Sci. Technol. 2004, 38, 1718-1722 85 Carlisle SM, Trevors JT, Glyphosate in the Environment, Water, Air, and Soil Pollution, 1988, 39, 409-420 86 Sprankle P, Meggitt WF, Penner D, Adsorption, Mobility, and Microbial Degradation of Glyphosate in the Soil, Weed Sci., 1975, 23(3) 229-234 87 Sprankle P, Meggit WF, Penner D, Rapid Inactivation of Glyphosate in the Soil, Weed Sci., 1975, 23(3) 224228 88 Sheldrick von WS, Morr M, N-phosphonomethylglycine, Acta Cryst. 1981, B37, 733-734 89 Shkol’nikova LM, Porai-Koshits MA, Dyatlova NM, Yaroshenko GF, Rudomino MV, Kolova EK, X-ray structural study of organic ligands of the complexone type. III. Crystal and Molecular structure of phosphonomethylglycine and iminodiacetic-monomethylphosphonic acid, Zhurnal Strukturnoi Khimii, 1982, 23(5), 98-107, (English translation) 90 Sheals J, Granström M, Sjöberg S, Persson P, Coadsorption of Cu(II) and glyphosate at the water-goethite (αFeOOH) interface: molecular structures from FTIR and EXAFS measurements, J. Colloid Interface Sci. 2003, 262, 38-47 91 Sheals J, Persson P, Hedman B, IR and EXAFS Spectroscopic Studies of Glyphosate protonation and Copper(II) Complexes of Glyphosate in Aqueous Solution, Inorg. Chem. 2001, 40 (17) 4302-4309 92 Torstensson L, Behaviour of glyphosate in soils and its degradation, Chapter 9 in The Herbicide Glyphosate Ed, Grossbard E & Atkinson D, 1985, Butterworths, London 93 Rueppel ML, Brightwell BB, Schaefer J, Marvel JT, Metabolism and Degradation of Glyphosate in Soil and Water, J. Agric. Food Chem., 1977, 25, 517-528 (due to ads to soil) 94 Sheals J, Sjöberg S, Persson P, Adsorption of glyphosate in goethite: Molecular characterization of surface complexes, Environ. Sci. Technol. 2002, 36, 3090-3095 95 McBride M, Kung K, Complexation of glyphosate and related lligands with iron, Soil Sci. Soc. Am. J, 1989, 53, 1668-1673 96 Gimsing, A. L. Borggaard, O. K. Bang, M. Influence of soil composition on adsorption of glyphosate and phophate by contrasting Danish surface soils, Eu. J. Soil Sci., 2004, 55, 183-191 97 Jonge, H de. Jonge, L.W. de, Influence of pH and solution composition on the sorption of glyphosate and prochloraz to a sandy loam soil, Chemosphere, 1999, 39(5), 753-763 98 Motekaitis RJ, Martell AE, Metal chelate formation by N-phosphonomethylglycine and related ligands, J. Coord. Chem.,1985, 14, 139-149 99 Deacon, G.B. Philips, R.J. Relationships between the carbon-oxygen stretching frequencies of carboxylato complexes and the type of carboxylate coordination, Coord. Chem. Rev. 1980, 33, 227-250 100 Heineke D, Franklin SJ, Raymond KN, Coordination Chemistry of Glyphosate: Structural and Spectroscopic Characterization of Bis(glyphosate)metal(III) Complexes, Inorg. Chem. 1994, 33, 2413-2421 101 Ramstedt M, Norgren C, Sheals J, Sjöberg S, Persson P, Thermodynamic and Spectroscopic Studies of Cadmium(II)-N-(phosphonomethyl)glycine (PMG) Complexes, Inorg. Chim. Acta, 2004, 357, 1185-1192 102 Clarke ET, Rudolf PR, Martell AE, Clearfield A, Structural Investigation of the Cu(II) Chelate of Nphosphonomethylglycine. X-ray Crystal structure of Cu(II)[O2CCH2NHCH2PO3]·Na(H2O)3.5, Inorg. Chim. Acta, 1989, 164, 59-63 103 Appleton TG, Hall JR, McMahon IJ, NMR Spectra of Iminobis(methylenephosphonic acid), HN(CH2PO3H2)2, and related Ligands and their complexes with Platinum(II), Inorg. Chem. 1986, 25, 726-734 104 Rajendram CRA, Hoggard PE, The coordination of aminophosphonates to chromium(III), J. Coord. Chem. 1994, 33, 15-32 105 Ptaszy´nski B, Zwoli´nska A, Synthesis and Properties of Solid Complexes of Lanthanum, Cerium, Neodynium and Erbium with N-Phosphonomethylglycine, Polish J. Environ. Studies, 2001, 10(4), 257-262 106 Appleton TG, Byriel KA, Hall JR, Kennard CHL, Lynch DE, Sinkinson JA, Smith G, Reactions of Di- and Trimethylplatinum(IV) Complexes with N-(phosphonomethyl)glycine (glyphosate) and Iminobis(methylenephosphonic acid). Crystal structures of three dimethyl(IV) complexes with Nphosphonomethyl)glycine coordinated facially, Inorg. Chem. 1994, 33, 444-455 107 Sanna D, Bódi I, Bouhsina S, Micera G, Kiss T, Oxovanadium(IV) complexes of phosphonic derivatives of iminodiacetic and nitrilotriacetic acids, J. Chem. Soc. Dalton Trans, 1999, 3275-3282 108 Szabó Z, Structure, equilibrium and ligand exchange dynamics in the binary and ternary dioxouranium(VI)glyphosate-fluoride system. A multinuclear NMR study, J. Chem. Soc. Dalton Trans, 2002, 4242-4247 72 109 Barja BC, Herszage J, dos Santos Afonso M, Iron(III)-phosphonate complexes, 2001, Polyhedron, 20, 18211830 110 Sagatys DS, Dahlgren C, Smith G, Bott RC, Willis AC, Metal Complexes with N-(Phosphonomethyl)glycine (Glyphosate): the Preparation and Characterization of the Group 2 Metal Complexes with Glyphosate and the Crystal Structure of Barium Glyphosate Dihydrate, Aust. J. Chem., 2000, 53, 77-81 111 Sagatys DS, Dahlgren C, Smith G, Bott RC, White JM, The complex chemistry of N(phosphonomethyl)glycine (glyphosate):preparation and characterization of the ammonium, lithium, sodium (4 polymorphs) and silver(I) complexes, J. Soc. Dalton Trans, 2000, 3404-3410 112 Rudolf PR, Clarke ET, Martell AE, Clearfield A, Preparation and X-ray Structure of the Calcium(II) Complex of N-Phosphonomethylglycine Ca[O2CCH2NH2CH2PO3]·2H2O at pH 7.2, Acta Cryst. 1988, C44, 796799 113 Smith PH, Raymond KN, Solid-State and solution Chemistry of Calcium N-(Phosphonomethyl)glycine, Inorg. Chem. 1988, 27, 1056-1061 114 Galdecka E, Galdecki Z, Gawryszewska P, Legendziewicz J, Structure of a novel polynuclear europium compound with n-phosphonomethylglycine: heptaaquaperchlorate monohydrate, [Eu2(HO3PCH2NH2CH2CO2)2(H2O)7(ClO4)]·3ClO4·H2O, New J. Chem., 2000, 24, 387-391 115 Bochatay L, Persson P, Sjöberg S, Metal Ion Coordination at the Water-Manganite (γ-MnOOH) Interface. I. An EXAFS study of cadmium(II), J. Colloid Interface Sci., 2000, 229, 584-592 116 Bochatay L, Persson P, Metal Ion Coordination at the Water-Manganite (γ-MnOOH) Interface. II. An EXAFS study of zinc(II), ), J. Colloid Interface Sci., 2000, 229, 593-599 117 Pan, G. Qin, Y. Li, X. Hu, T. Wu, Z. Xie, Y., EXAFS studies on adsorption-desorption reversibility at manganese oxides-water interfaces I. Irreversible adsorption of zinc onto manganite (γ-MnOOH), J. Colloid. Interface Sci. 2004, 271, 28-34 118 Sheals J, Molecular Characterisation of glyphosate complexes in aqueous solution and at the solutionmineral interface, PhD thesis, Umeå University, Sweden, 2002, ISBN 91-7305-343-0 73 Acknowledgements A lot of absolutely wonderful people have surrounded me during these past years, and some of you I would like to give a special “Thank you!” here in the end of my thesis. To start with, I would like to thank my supervisor Staffan Sjöberg. You have many times showed me that there actually does exist a way through the “swamps” when I thought I was about to give up. I am very grateful for your support and guidance; you’ve been a great supervisor! I also would like to thank Per Persson, my assistant supervisor, for introducing me to many fascinating spectroscopic techniques, for helpful discussions and for being such a good colleague. I’ll miss your skånish “Tjena Madde!” Lars Lövgren, my other assistant supervisor, I am very grateful that you introduced me into this field of science and let me become MiMi-associated in some parts of my research. I have learned many interesting things from you; from how to find the best restaurant to the composition of mining waste. My unofficial assisting supervisor, Andrei Shchukarev, who have taught me almost everything I know about XPS. I am very grateful to you for this knowledge and that you’ve shared your ideas and chocolate with me. It’s been a lot of fun! Thank you! Caroline and Kristina, my office mates; we have shared many things! Johan, Leif, Frasse, Laurence, Jörgen, Maria, Agneta, Ingegärd – thank you so much for helping and coaching me during my first years. Dan; thanks for answering all my questions about structural chemistry. Kristina, Julia, John and Åsa for proof reading my texts. Unfortunately, I won’t be able to thank all of my colleagues and friends at the department of inorganic chemistry individually on this page. All the same, I am very grateful to all of you for making my time at the department as great as it has been. I’ve had so much fun and I’ve learnt so many things! I want to thank Bruce Johnson at La Trobe University for helping me arrange my time in Australia and giving me the opportunity to work in the lab in Bendigo, but also for making me feel at home in Australia from the very first beginning. Thanks to the rest of the colloid surface group in Bendigo for your kindness and teaching me about Australian football. Paul Pigram, Narelle Brack and the students in their group for really making me feel welcome in Melbourne and allowing me to use their nice XPS. Ewen Silvester, thanks for all your help at CSIRO Clayton during and after my visit, and for showing me a bit of your Australia. Patrick Hartley and co-workers, for allowing me to work in their group and learn about AFM, especially I would like to thank Celesta Fong for all her help. Gudrun Boström, min mentor och bollplank; jag har verkligen uppskattat våra samtal, de har gett mig mycket glädje och varit som vitamininjektioner, Tusen tack!! Min underbara familj, i flera generationer, som alltid stöttar och tror på mig. Särskilt vill jag förstås tacka mamma, pappa och Natanael. Tack för att ni finns där!!! Och sist men inte minst är jag så tacksam till alla mina vänner utanför jobbet. Ni ger mig perspektiv på tillvaron, tillför så mycket glädje och betyder oerhört mycket för mig. Tack allihopa!! 74