CRYSTAL STRUCTURE MODELING OF A HIGHLY DISORDERED POTASSIUM BIRNESSITE

Similar documents
Copyright SOIL STRUCTURE and CLAY MINERALS

Crystal structures of two partially dehydrated chlorites: The modified chlorite structure

FORMATION OF CRYSTAL STRUCTURES DURING ACTIVATED CARBON PRODUCTION FROM TURKISH ELBISTAN LIGNITE

The effect of isomorphous substitutions on the intensities of (OO1) reflections of mica- and chlorite-type structures.

Analysis of Clays and Soils by XRD

Cation Exchange Capacity, CEC

Time-Resolved Synchrotron X-ray Diffraction Study of the Dehydration Behavior. of Chalcophanite. Jeffrey E. Post 1* and Peter J.

Layered Compounds. Two-dimensional layers. Graphite. Clay Minerals. Layered Double Hydroxides (LDHs) Layered α-zirconium Phosphates and Phosphonates

Structure of synthetic monoclinic Na-rich birnessite and hexagonal birnessite: II. Results from chemical studies and EXAFS spectroscopy

Soil Colloidal Chemistry. Compiled and Edited by Dr. Syed Ismail, Marthwada Agril. University Parbhani,MS, India

Remote Asymmetric Induction in an Intramolecular Ionic Diels-Alder Reaction: Application to the Total Synthesis of (+)-Dihydrocompactin

X-ray Diffraction. Diffraction. X-ray Generation. X-ray Generation. X-ray Generation. X-ray Spectrum from Tube

Soil Mechanics Prof. B.V.S. Viswanadham Department of Civil Engineering Indian Institute of Technology, Bombay Lecture 3

Structure Report for J. Reibenspies

NOTE TOSUDITE CRYSTALLIZATION IN THE KAOLINIZED GRANITIC CUPOLA OF MONTEBRAS, CREUSE, FRANCE

RIETVELD REFINEMENT WITH XRD AND ND: ANALYSIS OF METASTABLE QANDILITE-LIKE STRUCTURES

The structure of liquids and glasses. The lattice and unit cell in 1D. The structure of crystalline materials. Describing condensed phase structures


WHY DOES HALLOYSITE ROLL?--A NEW MODEL

The Solid State. Phase diagrams Crystals and symmetry Unit cells and packing Types of solid

EXPERT SYSTEM FOR STRUCTURAL CHARACTERIZATION OF PHYLLOSILICATES: II. APPLICATION TO MIXED-LAYER MINERALS

Supporting Information

Clays and Clay Minerals

MINERALOGICAL ASSOCIATION OF CANADA CLAYS AND THE RESOURCE GEOLOGIST

Electronic Supplementary Information

Lecture 14: Cation Exchange and Surface Charging

REFINEMENT OF THE CRYSTAL STRUCTURE OF A MONOCLINIC FERROAN CLINOCHLORE

HETEROGENEITY IN MONTMORILLONITE. JAMES L. MCATEE, JR. Baroid Division, National Lead Co., Houston, Texas

A re-examination of the structure of ganophyllite

Crystal structure of mixed-layer minerals and their X-ray identification: New insights from X-ray diffraction profile modeling.

CHARACTERIZATION OF SMECTITES SYNTHESISED FROM ZEOLITES AND MECHANISM OF SMECTITE SYNTHESIS

ADVANCED APPLICATIONS OF SYNCHROTRON RADIATION IN CLAY SCIENCE

STEPWISE HYDRATION OF HIGH-QUALITY SYNTHETIC SMECTITE WITH VARIOUS CATIONS

CHARACTERIZATION OF INTERCALATED SMECTITES USING XRD PROFILE ANALYSIS IN THE LOW-ANGLE REGION

Prof. B V S Viswanadham, Department of Civil Engineering, IIT Bombay

CRYSTAL STRUCTURE AND RIETVELD REFINEMENT OF ZEOLITE A SYNTHESIZED FROM FINE-GRAINED PERLITE WASTE MATERIALS

CLAY MINERALS BULLETIN

Tikrit University. College of Engineering Civil engineering Department SOIL PROPERTES. Soil Mechanics. 3 rd Class Lecture notes Up Copyrights 2016

Setting The motor that rotates the sample about an axis normal to the diffraction plane is called (or ).

Supporting Information

THE USE OF PIPERIDINE AS AN AID TO CLAY-MINERAL IDENTIFICATION

The Reciprocal Lattice

First-principles study of the structure of stoichiometric and Mn-deficient MnO 2

New insights on the distribution of interlayer water in bi-hydrated smectite from X-ray diffraction profile modeling 00l reflections

Microporous Manganese Formate: A Simple Metal-Organic Porous Material with High Framework Stability and Highly Selective Gas Sorption Properties

ION EXCHANGE, THERMAL TRANSFORMATIONS, AND OXIDIZING PROPERTIES OF BIRNESSITE

RESOLUTION OF THE POLYTYPE STRUCTURE OF SOME ILLITIC CLAY MINERALS THAT APPEAR TO BE 1Md

DIFFRACTION PHYSICS THIRD REVISED EDITION JOHN M. COWLEY. Regents' Professor enzeritus Arizona State University

Synthesis of Lithiophorite in High Alkaline Conditions

MOLECULAR SIMULATIONS OF MONTMORILLONITE INTERCALATED WITH ALUMINUM COMPLEX CATIONS. PART II: INTERCALATION WITH AI(OH)3-FRAGMENT POLYMERS

Chapter 3. The structure of crystalline solids 3.1. Crystal structures

UNIT-1 SOLID STATE. Ans. Gallium (Ga) is a silvery white metal, liquid at room temp. It expands by 3.1% on solidifica-tion.

Lecture 6. Physical Properties. Solid Phase. Particle Composition

Nanoparticles in Soils

Crystals, X-rays and Proteins

Understanding the relationship between crystal structure, plasticity and compaction behavior of theophylline, methyl gallate and their 1:1 cocrystal

Possible chemical controls of illite/smectite composition during diagenesis

THE EFFECTS OF ADDING 1 M NaOH, KOH AND HCl SOLUTION TO THE FRAMEWORK STRUCTURE OF NATURAL ZEOLITE

Solids. properties & structure

Crystal Structure Species Crystallisation

Manganese-Calcium Clusters Supported by Calixarenes

The Controlled Evolution of a Polymer Single Crystal

Supplementary Information: Twinning of cubic diamond explains reported nanodiamond polymorphs

RAMAN MICROPROBE SPECTROSCOPY OF HALLOYSITE

muscovite PART 4 SHEET SILICATES

Supplementary Information

Arsenite Oxidation by a Poorly Crystalline Manganese-Oxide. 2. Results from X-ray Absorption Spectroscopy and X-ray Diffraction

Order, Disorder and Crystallinity in Layered. Hydroxides. P. Vishnu Kamath. Central College, Bangalore University. Department of Chemistry

CRYSTAL STRUCTURE OF TETRAMETHYLAMMONIUM-EXCHANGED VERMICULITE

Reaction Landscape of a Pentadentate N5-Ligated Mn II Complex with O 2

NOTE FIBROUS CLAY MINERAL COLLAPSE PRODUCED BY BEAM DAMAGE OF CARBON-COATED SAMPLES DURING SCANNING ELECTRON MICROSCOPY

The crystal structure of lithiophorite

Phase identification and structure determination from multiphasic crystalline powder samples by rotation electron diffraction

X-ray analysis. 1. Basic crystallography 2. Basic diffraction physics 3. Experimental methods

Iron Complexes of a Bidentate Picolyl NHC Ligand: Synthesis, Structure and Reactivity

EPSC501 Crystal Chemistry WEEK 5

Studying the Effect of Crystal Size on Adsorption Properties of Clay

Time-resolved structural analysis of K- and Ba-exchange reactions with synthetic Na-birnessite using synchrotron X-ray diffraction

Supplementary Figures

MINERAL CONTENT AND DISTRIBUTION AS INDEXES OF WEATHERING IN THE OMEGA AND AHMEEK SOILS OF NORTHERN WISCONSIN

Study on Magnetic Properties of Vermiculite Intercalation compounds

Courtesy of Ray Withers Electron Diffraction and the Structural Characterization of Modulated and Aperiodic Structures

Solid Type of solid Type of particle Al(s) aluminium MgCl2 Magnesium chloride S8(s) sulfur

Chapter 12 Solids and Modern Materials

Schematic representation of relation between disorder and scattering

Crystal structure of Ni-sorbed synthetic vernadite: A powder X-ray diffraction study

Supplementary Information. Surface Microstructure Engenders Unusual Hydrophobicity in. Phyllosilicates

CHEM-E5225 :Electron Microscopy. Diffraction 1

- A general combined symmetry operation, can be symbolized by β t. (SEITZ operator)

Minerals: Minerals: Building blocks of rocks. Atomic Structure of Matter. Building Blocks of Rocks Chapter 3 Outline

Supporting information

A. BEN HAJ AMARA. L.P.M., Facultd des Sciences de Bizerte, 7021 Zarzouna, B&erte, Tunisia

Jnrpnnv E. Posr. D.l.vro L. Brsn

Basic Crystallography Part 1. Theory and Practice of X-ray Crystal Structure Determination

metal-organic compounds

CHAPTER 4. Crystal Structure

Supplementary Figure 1: Crystal structure of CRCA viewed along the crystallographic b -direction.

SOLID STATE 18. Reciprocal Space

RESULTS AND DISCUSSION Characterization of pure CaO and Zr-TiO 2 /CaO nanocomposite

Ch. 4 - Clay Minerals, Rock Classification Page 1. Learning Objectives. Wednesday, January 26, 2011

Transcription:

Clays and Clay Minerals, Vol. 44, No. 6, 744-748, 1996. CRYSTAL STRUCTURE MODELING OF A HIGHLY DISORDERED POTASSIUM BIRNESSITE KERRY L. HOLLANDt AND JEFFREY R. WALKER Department of Geology and Geography, Vassar College, Poughkeepsie, New York 12601 Abstract--The structure of a highly disordered synthetic birnessite was studied by comparing powder X-ray diffraction (XRD) data with calculated patterns generated by BIRNDIF and WILDFIRE9 in an attempt to describe the nature of disorder and to estimate the size of the coherent diffracting domains. The material has a turbostratic stacking sequence and coherent diffracting domains that are 25 to 30 ]k on a side in the ab plane (N1 = 5 unit cells, N2 = 10 unit cells) and which average 2.5 unit cells thick parallel to c. Turbostratic stacking probably results because there are few constraints on the relationship between adjacent layers. Key Words--Birnessite, Crystal Structure Modeling, Turbostratic Stacking. INTRODUCTION Birnessite is a phyllomanganate, one of a family of Mn-oxide minerals with similar layer structures and different interlayer arrangements, which includes chalcophanite with interlayer Zn, birnessite with a variety of interlayer cations, buserite with hydrated interlayer cations, asbolane with interlayer cations octahedrally coordinated by OH and vernadite (Burns and Burns 1976). These phyllomanganates are of interest in sedimentary environments because they can be effective scavengers and oxidizers of metals (Oscarson etal. 1981; Moore et al. 1990). Results of an investigation of the oxidation of arsenic by bimessite suggest that the crystal structure of the mineral is important to its efficient participation in this reaction (Moore etal. 1990). However, due to their fine grain size and poor crystallinity the structures of the phyllomanganates are not well known. Most are seen as derivatives of the chalcophanite structure (Wadsley 1955), which is based on sheets of edge-sharing Mn4+-O octahedra. In chalcophanite, vacancies in the Mn-O octahedra create charge deficiencies that are balanced by Zn cations in the interlayers (Post and Appleman 1988). Giovanoli et al. (1970) and Burns and Burns (1976) suggested that birnessite has a similar structure in which Mn atoms occupy the Zn sites; however, a recent refinement of the birnessite structure (Post and Veblen 1990) was unable to resolve whether layer charge resulted from substitution of Mn 3+ for Mn 4+ or from the presence of vacancies. Post and Veblen (1990) also reported asymmetric broadening of (hk0) peaks, which they felt did not affect their refined atomic coordinates but did suggest the presence of some degree of structural disorder. t Present address: Colorado School of Mines, Boulder, Colorado. Naturally-occurring birnessite is often very poorly crystalline with only a few broad diffraction peaks. Most careful crystal structure work has been conducted on synthetic specimens that have a reduced variety of interlayer cation species and, since they are relatively well-crystallized, exhibit sufficient diffraction maxima to make structural studies possible. The relationship between the poorly crystalline natural material and the relatively well-crystallized synthetic material is not clear. It is the aim of this work to investigate the crystal structure of a poorly crystalline synthetic K-bimessite that might be thought of as an analog for natural specimens. EXPERIMENTAL METHODS Samples of synthetic K-birnessite were obtained from J. N. Moore, at University of Montana. The material was synthesized by reduction of KIVInO 4 with HC1. Details of synthesis and subsequent reaction experiments with As are given by Moore et al. (1990). The chemical composition of the samples at the time of synthesis was (Moore et al. 1990): K0.33Mn39~vOla.7H20 [1] The cell parameters and atom positions of Post and Veblen (1990) were used in the structural modeling with the exception of c, which was calculated from the position of the (002) peak in the experimental diffraction pattern (Table 1). XRD data were collected with a Siemens D5000 diffractometer using CuKet radiation (40 kv, 30 ma), a graphite monochromator, a l~ slit assembly and a 0.6-mm receiving slit. Samples were backpacked in an A1 mount and scanned from 10 to 90 ~ using a step of 0.05 ~ Between 10 and 55 ~ the samples were analyzed at 20 s/step and between 55 and 90 ~ at 60 s/step. A representative diffraction pattern is shown in Figure l a. Copyright 9 1996, The Clay Minerals Society 744

Vol. 44, No. 6, 1996 Highly disordered birnessite 745 Table 1. K-birnessite atomic and cell parameters (Post and Veblen 1990). Temp x y z occt factor Mn 0 0 0 2 0.5 O 0.365 0 0.136 4 1 K 0.723 0 0.522 0.5 0.5 H20 0.723 0 0.522 0.5 1.5 O 0 0 0.5 0.3 1 a = 5.149 b = 2.834 c = 7.176 A = 100.76 ~ t occ = occupancy factors in atoms/unit cell. Diffraction results were modeled using 2 programs, both modifications of existing software. The first program, BIRNDIE was modified from CLAYDIF (Reynolds 1985) by adding the X-ray scattering coefficients (Wright 1973) for Mn and by substituting the atomic coordinates of the birnessite Mn-O octahedral layer for one of the silicate layers in the program. BIRNDIF calculates the basal diffraction pattern of birnessite and can be used to determine the mean defect-free stacking distance (8) perpendicular to the layers by iterative modeling of selected basal peak shapes and widths. The second program, WILDFIRE9 (Reynolds 1993, 1994), was also modified to include X-ray scattering by Mn. Because WILDFIRE9 uses an input file of atomic coordinates, it did not have to be modified to include the birnessite structure. The monoclinic coordinates of Post and Veblen (1990) were transformed to orthogonal space as described by Reynolds (1993) prior to modeling. WILDFIRE9 calculates diffraction phenomena from non-basal atom planes within the crystal. It performs the calculation by sweeping a vector 1/d in length through reciprocal space and calculating the contribution to diffracted intensity at each point on the resulting surface. Because layered minerals can exhibit varying amounts of disorder in one dimension (perpendicular to the layers), diffraction spots in reciprocal space are stretched parallel to the direction of the disorder (c*) with maxima that vary in intensity depending upon the severity of the disorder and the change in the structure factor along c*. For the maximum / 20; 11 02;31 22;40 *1,I,I, I,l.l,l,f,l,I,I,l,l,l,l,l,f,l,l,l,l,l,lll,l,l,l,l.l,I, I,l,l,l,l,l,lxl,l,l,I i0 14 18 22 26 30 34 38 42 46 50 54 58 62 66 70 74 78 82 86 Figure I. X-ray powder diffraction patterns of potassium birnessite identifying the 2-dimensional diffraction bands, la) Experimental data. lb) Calculated pattern which is a combination of patterns calculated by BIRNDIF and WILDFIRE9

746 Holland and Walker Clays and Clay Minerals results were obtained by setting N1 = 5 unit cells and N2 = 10 unit cells, suggesting that the diffracting domains are equidimensional in the ab plane with sides 25 to 30,~ in length. 10 20 30 40 50 60 70 s0 90 ~ 20 (CuKt~) Figure 2. Comparison of calculated basal XRD pattern with experimental pattern. 2a) Experimental data (same as Figure la). 2b) Diffraction pattern calculated using BIRNDIE amount of disorder as represented by turbostratic stacking (that is, each layer is translated and/or rotated randomly in relation to adjacent layers), intensity varies smoothly along the axis of stretching and the spots become rodlike. Variables that were manipulated during modeling with WILDFIRE@ were the proportion of Mn to simulate vacancies or Mn 3+ substitution, and the size of coherent scattering domains in terms of the number of unit cells along the a = N1, b = N2 and c = N3 axes in WILDFIRE@. The final calculated pattern (Figure lb) was created by adding the basal diffraction pattern calculated by BIRNDIF to the non-basal pattern calculated by WILDFIRE@ (Reynolds 1993) for reasons discussed below. RESULTS The diffraction pattern from the K-birnessite used in this study contains 5 broad peaks, 3 of which are asymmetric. Initial investigations of this sample (Moore et al. 1990) reported data to 65 ~ (CuKa) and assumed that all peaks represented (00l) diffraction. However, modeling of the diffraction pattern with BIRNDIF indicates that only the 001 and 002, and possibly the 004, peaks are present at the resolution of the experimental pattern (Figure 2). Therefore, other peaks in evidence must be due to hkl diffraction. BIRNDIF modeling also determined that the mean defect-free distance (8) parallel to c is 2.5 unit cells. The character of the asymmetry of the maxima at 35, 66 and 78 ~ suggests that the layer stacking is turbostratic, in which case the asymmetric peaks represent the (20;11), (02;31) and (22;40) diffraction bands (Figure 1). To test this hypothesis, hk diffraction was modeled with WILDFIRE@ by setting N3 = 1 unit cell (i,e., no coherence along the c axis). Best DISCUSSION The most important results of this study are the turbostratic nature of the layer stacking of this birnessite sample and the small size, but nearly equidimensional ab shape, of its diffracting domains. The nature of this disorder is consistent with the findings of Post and Veblen (1990), who reported streaking in (110) electron diffraction peaks and broadening of (111) XRD peaks. Confusion may arise because the diffraction domains are said to be 2.5 unit cells along the c axis for (00/) reflections yet only 1 unit cell thick for 3-dimensional reflections. The mean defect-free distance (~) is the average number of parallel layers, so the reported distance implies that stacking sequences average only 2.5 layers without a stacking fault disrupting the parallelism of the layers. Rotations and translations of adjacent layers do not affect ~ as long as they do not disrupt that parallelism; however, non-basal atom planes are not coherent across the interface and N3 = 1. Because WILDFIRE@ cannot reconcile 8 = 2.5 and N3 = 1, the final calculated pattern is a composite of the 1-dimensional diffraction pattern calculated with BIRNDIF using the appropriate mean defect-free distance, and the 3-dimensional diffraction pattern calculated with WILDFIRE@ using the appropriate N3. Occurrence of turbostratic stacking sequences may be controlled by the nature of the basal atomic surface of the octahedral layers. The regular basal surface of the birnessite octahedral layers presented to the interlayer region is made up of O atoms with, possibly, some OH molecules. The regularity of this surface produces no holes into which large interlayer cations can fit. The situation is analogous to the smectites in which turbostratic stacking is common because the interlayer space is large enough to accommodate hydrated cations or molecules of ethylene glycol. Turbostratic sequences are not common in illite or micas where layers collapse around dehydrated interlayer cations that fit into the ditrigonal holes in the silicate layer base. The fact that the atoms of the basal plane of birnessite are predominantly O rather than OH groups implies that H-bonding is not common although it is an important bonding mechanism in many other layered structures. Turbostratic stacking is rare in chlorite, for instance, where H bonds between interlayer OH groups and silicate basal O atoms are strong enough to be the only force joining the layers (Bish and Giese 1981). Because sites of vacancies or Mn 3 in the octahedral layers appear to be disordered, the location of layer

Vol. 44, No. 6, 1996 Highly disordered birnessite 747 Figure 3. Schematic representation of the upper half of an a'c* reciprocal lattice plane demonstrating the method of calculation used by WILDFIRE9 for a 2-dimensional crystal (Reynolds 1993). The vertical columns labeled (lkl), (2kl), etc., are reciprocal lattice rods of a turbostratic crystal. The structure factor varies smoothly along the rods. Vectors (1/dl, l/d2, etc.) are swept through reciprocal space and a diffracted intensity is calculated wherever a vector intersects a rod. If diffracting domains are small, the diameter of the rods is large (lighter pattern) and the structure factor is more likely to change over the portion of the rod included in the calculation. In addition, rods of higher index, such as (3kl) and (4kl), are cut by the vector at a lower angle to their axes relative to rods of lower index, such as (lkl) and (2kl), so the area of the rod included by the calculation is larger, further increasing the chances that the structure factor will vary over the portion of the rod included. charge associated with these charge deficiencies is also disordered. The K atoms in the interlayer of K-bimessite, which are in the same crystallographic sites as the H20 molecules in the center of the interlayer (Post and Veblen 1990), can be coordinated by O atoms anywhere in the basal planes of the surrounding layers and are probably located near the disordered layer charge centers of the 2 adjacent octahedral layers in a configuration that maximizes local charge balance and minimizes cation repulsion. The regularity of the atomic surface, the probable lack of H-bonding and the disorder of layer charge sites indicate that there are few constraints on the way in which successive octahedral layers articulate around the interlayer. These factors further suggest that Nabirnessite, in which Na atoms occupy the same crystallographic plane as the water molecules in K-birnessite (Post and Veblen 1990), should also form turbostratic stacking sequences, whereas Mg-birnessite, in which Mg cations are located between the H20 layer and the basal O planes (Post and Veblen 1990), might form more ordered structures. The agreement between the observed and calculated patterns in this study is not perfect, and becomes worse at higher diffraction angles. There are several possible reasons for this discrepancy. WILDFIRE9 is optimized for the study of illite using diffraction phe- nomena below approximately 45 ~ (CuKc0. If the temperature factor (~Liso) used is too large, calculated intensities will become increasingly smaller than experimental intensities at higher diffraction angles. However, even when ixi~ o was decreased to ~ of the default value, no significant improvement in the agreement of high angle intensities was noted. WILDFIRE9 is also optimized to model diffracting domains that are large in the ab plane (300,~ on a side), making the assumption that the structure factor (F) is constant across the diffraction "rods" in reciprocal space (Reynolds 1993). As shown in Figure 3, this assumption may not be valid for diffracting domains that are as small as those modeled for this birnessite. For small diffracting domains, the diameter of the diffraction "rods" in reciprocal space is proportionately larger, increasing the possibility that F is not constant in the region of the calculation. Additionally, at increasingly higher indices, the diameter (l/d) of the hemisphere over which intensity is integrated becomes larger, so the calculation sweeps through the rod at a lower angle to the axis of the rod and incorporates a larger portion of reciprocal space. Because intensity varies smoothly along the rods, there is a greater chance that the structure factor will change in the calculation region. The 2 higher angle diffraction bands contain diffraction phenomena from the (311) and (401)

748 Holland and Walker Clays and Clay Minerals classes of reflections, respectively, and so may be affected by changes in the structure factor across the (31) and (40) diffraction rods. It was not possible with the data and analytical method used here to distinguish between the 2 possible layer-charge producing mechanisms, vacancies or Mn 3 substitution. The samples used in this study were prepared under reducing conditions so some Mn 3+ may be present in the octahedral layers. The proportion of vacancies or trivalent substitution is small (Post and Veblen 1990) and their presence is probably not recognizable given the structural disorder evidenced by the data. It has been noted that birnessite prepared under oxidizing conditions tends to form more well-ordered crystals than that prepared under reducing conditions (Alain Manceau, personal communication, 1994) suggesting that future studies might address the structural differences between birnessites synthesized by the 2 different methods. CONCLUSION Although the birnessite studied here is synthetic, it has an XRD pattern that suggests crystalline disorder analogous to that found in naturally-occurring sampies. The mineral has a turbostratic stacking sequence and coherent diffracting domains that are 25 to 30,~ on a side in the ab plane and average 2.5 unit cells thick parallel to c. Turbostratic stacking probably results because there are few constraints on the articulation of adjacent layers. The experimental and calculated XRD patterns are not in perfect agreement, especially at higher diffraction angles, possibly because the assumption made by WILDFIRE9 that the structure factor does not vary across the diffraction "rods" in reciprocal space may not be valid for such small diffracting domains and/or reflections with such high indices. ACKNOWLEDGMENTS We would like to thank J. N. Moore for introducing us to birnessite and for providing us with these samples. The mann- script benefited from reviews by R. C. Reynolds, Jr, J. E, Post and D. L. Bish. Address for A. Manceau personal communication: LGIT-IRIGW, BP53X, Grenoble 38041, France. REFERENCES Bish DL, Giese R. 1981. Interlayer bonding in IIb chlorite. Am Mineral 66:1216-1220. Burns RG, Burns VM. 1976. Mineralogy of ferromanganese nodules. In: GP Glasby, editor. Marine manganese deposits. Amsterdam: Elsevier Science 554 p. Giovanoli R, St~hl E, Feitknecht W. 1970. f.)ber Oxihydroxides des vierwertigen Mangans mit Schichtengitter, 1. Mitteilung: Natiummangan(II,III)manganat(iV). Helv Chim Acta 53:209-220. Moore JN, Walker JR, Hayes TH. 1990. Reaction scheme for the oxidation of As(lII) to As(V) by birnessite. Clays Clay Miner 38:549 555. Oscarson DW, Huang PM, Liaw WK. 1981. The role of manganese in the oxidation of arsenite by freshwater lake sediments. Clays Clay Miner 29:219-225. Post JE, Appleman DE. 1988. Chalcophanite, ZnMn307.3H20: New crystal-structure determinations. Am Mineral 73: 1401-1404. Post JE, Veblen DR. 1990. Crystal structure determinations of synthetic sodium, magnesium, and potassium birnessite using TEM and the Rietveld method. Am Mineral 75:477-489. Reynolds RC, Jr. 1985. CLAYDIF: A computer program for the calculation of one-dimensional diffraction patterns of pure clay minerals. Hanover, NH: RC Reynolds, Jr, 8 Brook Rd. Reynolds RC, Jr. 1993. Three-dimensional X-ray powder diffraction from disordered illite: simulation and interpretation of the diffraction patterns. In: Reynolds RC, Jr, Walker JR, editors. CMS workshop lectures, vol 5, Computer applications to X-ray powder diffraction analysis of clay minerals. Boulder, CO: The Clay Minerals Society. p 43-78. Reynolds RC, Jr. 1994. WILDFIRE9 A computer program for the calculation of three-dimensional X-ray diffraction patterns for mica polytypes and their disordered variations. Hanover, NH: RC Reynolds, Jr, 8 Brook Rd. Wadsley AD. 1955. The crystal structure of chalcophanite, ZnMn3Ov.3H20. Acta Crystallogr 8:165-172. Wright AC. 1973. A compact representation for atomic scattering factors. Clays Clay Miner 21:489-490. (Received 26 January 1994; accepted 23 January 1996; Ms. 2460)