You are on page 1of 11

Clay Minerals (1984) 19, 865-875

MAGNESIA

FROM

SEAWATER:

A REVIEW

A. S. B H A T T I ,

D. D O L L I M O R E

ANt) A. D Y E R *

University of Toledo, Department of Chemistry, Toledo, Ohio 43606, USA, and *University of Salford, Department of Chemistry & Applied Chemistry, Salford M5 4 WT, UK (Received 7 November 1983; revised 18 May 1984)

ABSTRACT: The process developed in the UK to produce magnesium hydroxide from seawater is described, together with the heat treatment that the hydroxide receivesto produce the active oxide. Some of the characteristics required of the dolomite used in the process are also discussed. Impuritiesintroduced by the seawater are noted and the means by which they can be reduced explained. Magnesium oxide has long been an important industrial material. The major tonnage outlet for magnesia has been the refractories industry in which it is employed as 'dead-burned' grain magnesite for the construction and maintenance of the hearths or bottoms of open-hearth steel furnaces. Magnesia also is used in the form of calcined refractory magnesite brick and as a chemical addition in the manufacture of chrome, olivine and other refractories. The industrial importance of magnesia is due to its chemical behaviour and this is largely determined by the position of magnesium in the periodic table. Magnesium oxide is a weak base, particularly at high temperatures, and therefore magnesium oxide is barely attacked by alkalis but readily by acids. All strong mineral acids dissolve or corrode magnesium oxides, even at room temperatures. Only phosphoric, boric and, to a certain degree, silicic acids have little effect on magnesium oxides, because salts with these acids are insoluble in water and stable substances. For the same reason, hydrofluoric acid is not particularly aggressive towards magnesium oxide, and the insoluble magnesium fluoride causes a certain degree of protection of the magnesium oxide surface against further attack. Commercial magnesium oxides are generally used in the form of fine grains and powders (an exception is hot-pressed MgO) and, as such, are susceptible to attack by atmospheric moisture and carbon dioxide, forming hydrates and carbonates respectively. Larger crystals of calcined oxides, especially after hot-pressing, are resistant for all practical purposes to water vapour and to carbon dioxide. Magnesium oxide can only be melted at 2800~ in a fully oxidizing atmosphere, and there is very little volatilization below 3200~ A slightly reducing atmosphere suffices at these temperatures, however, to reduce it to the metal which is then oxidized by air in the upper part of the furnace to produce thick clouds of magnesium oxide, the density of which increases from 3.0 to 3.58 g/cm 3 for the sintered product. DIFFERENT TYPES OF MAGNESIAS

In general, only two classifications of calcined magnesia are in common useage: 'caustic/calcined' magnesia, calcined at low temperatures (350-450~ and 'dead-burned' 9 1984 The Mineralogical Society

866

A. S. Bhatti et al.

magnesia, calcined at much higher temperatures (>1300~ Temperature, rate of calcination and impurities materially influence the properties of magnesia obtained by heating the carbonate, hydroxide and other compounds. The quantitative effects of these factors on ultimate crystal size and reaction rates have not always been completely evaluated in commercial products. Magnesium oxide does not occur naturally to any great extent. It is made by calcining magnesium salts and there are two main types available. These are: 1. calcined natural magnesite (MgCO3) or brucite (Mg(OH)2); 2. calcined precipitated magnesium hydroxide, utilizing seawater or natural brines as the main source of Mg 2+ ions. The first type has industrial applications based on its relatively non-reactive properties, e.g. as a filler in rubber or polyester resins. The second type is widely used as a filler where more reactivity is required; it is this type that will be discussed here. Fig. 1 shows the calcination of magnesium hydroxide and the type of relationship between surface area and magnesium oxide content of the products. By controlling the calcination conditions, a porous magnesium oxide with large surface area can be formed. The surface area affects its rate of reaction with acids. Active magnesium oxides will react rapidly with water in the air to form the hydroxide and with carbon dioxide to form the carbonate. Magnesium is present as a soluble salt in seawater which contains about 1.3 10-3 kg/litre Mg 2+ ions, associated with chloride and sulphate ions (Hall & Spencer, 1973). This is equivalent to 2 10 -3 kg/litre magnesium oxide and, in principle, can be easily extracted by the addition of a suitable alkali. In the early 1930s small-scale processes designed to extract chemical magnesia from concentrated brine and seawater were investigated. Earlier attempts to operate a small plant had not been successful as the gelatinous magnesium hydroxide precipitates proved too slow in settling and filtering. In 1938 the first large-scale commercial plant for producing refractory magnesia was built at Hartlepool, by the

Mg(OH) z

h~, ~' Mc:JO~(OH)2 ~y + yH20

70

810 % MgO

9'0

FIG. 1. Schematic representation of variation in surface area vs. M g O content of heated Mg(OH)2.

Magnesia from seawater

867

Steetley Company, using a then unique process of extracting magnesia using seawater and dolomite. Because of their consistent composition and quality, magnesias from seawater often are preferred to natural magnesias. The trace element content of the product is controlled by careful raw material selection and, in some cases, certain trace elements can be added to the magnesium hydroxide precipitates, before calcination, to improve the sintering characteristics of the magnesias. By controlling the preparation and precipitation conditions, the amounts of chloride, sulphate and other ions can be controlled. These, in turn, affect the sintering characteristics of the magnesia (Green, 1969; Guilliat & Brett, 1971; Banerjee, 1969; Riddell, 1969; Laminy, 1969)--e.g. a very low content of C1 ions would indicate that the particular magnesia will calcine more quickly than one with a higher concentration. In a review of the production and properties of seawater magnesia, Hall & Spencer (1973) outlined the physical and chemical characteristics required of magnesia powders for use in oxygen steel-making vessels. These include: (i) ahigh density (3.35-3-40 103 kg/m3); (ii) a low boron content (B203 < 0.05%); (iii) The correct CaO/SiO 2 molar ratio to give the refractory phase dicalcium silicate; (iv) a CaO content less than the solubility limit in MgO. Magnesia, like other solids, especially oxides, has physical properties which are determined by its previous history and heat treatment. The physical properties of such solids, as distinct from their chemical properties, may be affected profoundly by the pretreatment that they have received. The term loosely applied in this connection is 'activity'. Although the chemical constitution remains the same, the more 'active' a solid is, the more rapidly it takes part in chemical reaction. An early detailed investigation of the factors governing the activity of solids was made by Huttig (1941, 1942), with particular reference to the changes produced on heating.

The preparation of active magnesia


The usual way of preparing active solids is to utilize thermal decomposition or calcination (Sing, 1949; Gregg & Sing, 1951); the best known active solids prepared by this method are lime or plaster of paris. Consider a chemical reaction of the type Solid A --, Solid B + Gas C in which, providing the correct conditions of temperature and time of heating are chosen, the solid B is obtained in a very active form. The essential requirement for solid B to be more 'active' than A is that A and B should have different structures, i.e. the above equation represents a true chemical decomposition. The mechanism of activation has been discussed by Gregg & Sing (1951), Hill (1951) and Sing (1949): the decomposition of solid A first gives rise to product B, which is in the form of a pseudo-structure of A in which the metal and oxygen ions are still occupying much the same positions as they held in the original structure. RecrystaUization of the pseudo-structure produces the stable solid B. Care must be taken when preparing active solids that the conditions are not too vigorous, for then an inactive solid is obtained due to sintering; the active sample may be produced temporarily, but the stringent environment can cause the small crystallites to adhere with a subsequent reduction in internal area. An investigation by Huttig (1942) into the sintering of metals and metal oxides demonstrated the importance of

868

A. S. Bhatti et al.

calcination temperature in decreasing the activity of the solid, but in this investigation the solids were always morphologically stable over the temperature range used. Huttig's metal oxides were preheated to a high temperature and then finely ground; his metals were also in a finely-divided form with an active solid prepared by a decomposition of the type shown in the above equation, In equilibrium, however, the sintering process is altered by the occurrence of a polymorphic transition, since according to Hedvall (1937) such a transformation should give rise to additional activation. Sing (1949) and Gregg & Sing (1951) studied aluminium hydroxide and hydrated ferric oxide, two systems showing different activation-temperature curves. With ferric oxide activation was greatest with the unheated sample, and subsequent heating caused a steady, unbroken fall in activity; when aluminium hydroxide was examined, unheated samples were not very active and alumina (with a surface area of ~300 mZ/g) was produced only when the sample was heated to 300-500~ Magnesium hydroxide can readily be obtained as a stable, crystalline solid, and on heating it loses water to give magnesium oxide; the only solid components in this system are magnesium hydroxide (brucite) and magnesium oxide which does not undergo any subsequent changes up to its melting point (2800~ Some workers have studied the changes that occur on heating magnesium hydroxide while others have examined the closely allied system of magnesium carbonate. Thus Budnikov (1930) used the rate of dissolution in water as a criterion for the activity of magnesium carbonate at various temperatures, and discovered that the most 'active' sample was that obtained at 700~ Lamb & West (1946) studied the changes in adsorption of nitrous oxide on samples prepared by removing different amounts of water from magnesium hydroxide. However, this work did not restrict the time and temperature o f heating within definite limits and, in addition, did not extend outside the range 300-500~ the rate of sintering cannot therefore be estimated from their data. Again, Fricke & Luke (1935) and Fricke (1936) showed from the heats of solution in acids and other measurements that the total energy content of magnesium hydroxide and magnesium oxide varied considerably with the method and temperature of preparation; a variation of 11.3 k J/mole was noted between the energy contents of the most active and least active sample. Similar studies by later workers (Ginque, 1949; Torgoson & Schamer, 1948; Taylor & Wells, 1938) confirmed this effect. Taylor & Wells (1938) noted no change in crystal structure, so they attributed the variations in heat content mainly to differences in surface properties. Adsorption on solid magnesium oxide has been used as a basis for measurements of its surface area. Adsorption of dyes and other substances from the liquid phase is fraught with difficulties due to possible preferential adsorption of the solvent, as emphasized by McBain & Dunn (1948). Zettlemayer & Walker (1947) compared the adsorption of iodine from carbon tetrachloride solution with the results obtained by nitrogen adsorption on a number of active magnesium oxide samples. Agreement between the two methods was not good, iodine adsorption in all cases being lower. The active magnesium samples used by these workers were obtained from a large-scale plant and the experiments they performed were aimed at characterizing the particular samples rather than finding the exact changes in activity with varying treatments. Some calcination studies were carried out on magnesium oxide by Eubank (1951) who reached similar conclusions to those of other investigators, i.e. the change from the brucite to the periclase structure occurs rapidly above 350~ and that magnesium oxide only exists in one of these forms. However, he did notice a slight distortion of the unit cell of the periclase structure at lower temperatures.

Magnesia from seawater

869

More detailed research (Glasson, 1963) has shown that the unit-cell dimension of the periclase varies from 4.24 A to a limiting value of 4.21/~ as the calcination temperature is increased from 400 to 1000~ the final density of the periclase being 3.58. The contraction of the unit cell is accompanied by expulsion of traces of hydroxyl water. Magnesium oxide from basic magnesium carbonate gave similar results. Changes in surface area during calcination of magnesium hydroxide and magnesium carbonate (magnesite) have been followed by a gravimetric nitrogen gas sorption technique (Glasson, 1956). Magnesium oxide of maximum surface area is obtained at about 80-90% hydroxide decomposition and 90-100% carbonate decomposition. Similar studies on Steetley dolomite (Glasson, 1964b) gave analogous results. In the dolimes formed, the average crystallite sizes (from X-ray line-broadening) of the magnesium oxide were correspondingly much smaller than those of the calcium oxide. Steetly and Coxhoe dolomites gave products of correspondingly similar activities at lower calcination temperatures, but above 700~ the Coxhoe dolimes were less active, as impurities such as ferric oxide promoted sintering. Further studies have been made on the hydration of calcium and magnesium oxides (Glasson, 1958, 1960, 1961, 1963)and dolomitic lime (Glasson, 1964a), and their reactions with seawater have been studied by surface area, optical- and electronmicrographic (Glasson et al., 1968) and radio-isotopic (Glasson et al., 1968) techniques. Nearly every commercial process that has been developed for the production of magnesium metal or magnesia requires either a high-calcium or dolomitic lime as the basic raw material (Boynton, 1966). Thus the Dow Seawater Process uses hydrated time, MgC12 (from seawater) + Ca(OH) 2 --, Mg(OH) 2 + CaC12, followed by either (A) Mg(OH)2 + 2HC1 --, MgC12 + 2H20 and MgC12 + electrolysis -, Mg + C12
or

(B) Mg(OH) z + heat --, MgO + H20 (Seawater Chemical Magnesian Process) The Dow Natural Brine Process uses dolomitic lime, CaC12, MgC12 + CaO.MgO + 2CO z --, 2CaCO 3 + 2MgC12, followed by electrolysis of MgC12. The Steetley Process also uses dolomitic lime and is described now in detail.

PROCESS

FOR

PRODUCING

MAGNESIA

FROM

SEAWATER

The chemistry of the process is relatively simple. First, dolomite is decomposed in a rotary kiln at 1350-1400~ The calcination reaction is
CaMg(CO3) 2 ~

dolomite The dolime is then hydrated MgO. CaO

MgO. CaO + CO2 dolime

+ H20 ~

MgO. Ca(OH) 2

870

A. S. Bhatti et al.

and finally reacted with the seawater to precipitate magnesium hydroxide MgO. 2Ca(OH) 2 + MgSO 4 CaSO 4 MgC12 --, M g O . 2 M g ( O H ) 2 + CaC1 z magnesium salts spent seawater seawater

Some of the MgO from the dolomite will hydrate during the process, but most of it comes through unchanged, because when the dolomite is fired to 1350-1400 ~ C the MgO forms a deadburnt layer around itself and becomes almost inert. The production process The calcined dolomite is slaked to a fine dry powder in pay hydrators. The hydrated product is made into a slurry to facilitate handling by mixing with seawater and is classified to remove impurities, particularly fuel ash and any non-calcined or over-calcined, unreactive materials. The seawater is first pretreated to remove calcium bicarbonate-achieved by adding a small proportion of sulphuric acid to change the pH to 3.4-4.0 to the centre of the tank by rotating r a k e s - - a n d then pumped out as a thick slurry. This slurry is washed with fresh water to reduce the sodium chloride content. The main problem with this process has always been that the precipitates settle badly and are very difficult to filter. It has been found that magnesium hydroxide derived from

Fe2+ ~ 13% CaCOa.MgCO3

Fe2~ < 0 1% CaCO3. MgCO~

l oo~
CaO.MgO ,l HzO Ca(OH)2. MgO I H20 + Mg2+ 2Mg(OH)2 + Ca2+

t
CaO I H2o Ca(OH)2 I MgZ+ Mg(OH)2 + Caz{

Rotary vacuum filters

Paste of 50% Mg(OH)z + 50% H20

Multiple hearth furnace

MgO

FIG. 2. Two methods of producing magnesiumoxides.

Magnesia from seawater

871

the dolomite acts as nuclei for precipitated magnesium hydroxide and improves the settling rate. The settled slurry still contains 0.1 to 0.12 kg/litre of solids and, even after steam-heating to 90 ~ C before filtering, only produces a filter cake of 30-35% solid content (Hedvall, 1937). A process of seeding has been developed (Budnikov, 1930) where up to 85% of the product is fed back to the reaction tanks. These recycled magnesium hydroxide particles act as nuclei for the precipitation of further magnesium hydroxide and the resulting precipitate settles down to an increased solids concentration of 0.3-0.4 kg/litre; this can be filtered more rapidly, and to a higher solids concentration (50%), without heating. In 1955, organic flocculants became available and they are effective at very low concentrations (1-10 p.p.m.), thus making their use economically viable. Those based on polyacrylamide are particularly effective in enhancing the settling rate of magnesium hydroxide precipitates but do not result in a filter cake with a lower free water content. The slurry is filtered on rotary vacuum disc filters, and additions of lime, silica, and iron are made in order to control the ratios of the impurities in the magnesia. The filter cake, containing 50% solids, is then either fed directly into rotary kilns or calcined in a multi-hearth Herreshoff furnace and formed into pellets by passing through pelletizing rolls. The rotary kilns 'dead-burn' the product at temperatures in excess of 1600~ to produce refractory grade magnesia. Magnesias prepared using dolomite contain ~ 1.3% Fe203; magnesias containing <0.1% F%O 3 are prepared using calcium carbonate. The two methods are shown schematically in Fig. 2.

The Herreshofffurnaee
The paste from the rotary vacuum filters is fed to the top of the furnace and it slowly travels along the hearths until lightly calcined magnesia is discharged from the bottom. To produce magnesium oxide of high purity, high density and large grain size, a carefully controlled temperature profile is required. The upper hearths effect the first stage, i.e. removal of free water, the middle hearths remove combined moisture and volatiles, and the lower hearths condition (calcine and sinter) the oxide. The temperature profile ranges from 400~ at the top to 900~ at the bottom of the furnace. The exhaust gases carry with them some of the oxide (20%) as dust which is drawn from the furnace and cleaned in a bag filter unit which strips out the oxide and allows the clean gases to pass to atmosphere, the collected dust being returned to the middle section of the furnace.

Impurities from dolomite


Careful control of raw dolomite quality must be maintained because the impurities in the calcined dolomite are precipitated in the magnesium hydroxide. In order to achieve this control, it is sometimes necessary to blend rock from different areas of the quarry, and the crushed dolomite is washed to remove clays, and hence reduce potential AlzO 3 and SiO 2 contamination. Hydrocycloning removes coarse grit, which has a high silica content (Hall & Spencer, 1973). Dolime produced initially from dolomite must be of good reactivity, as any unreacted dolime will appear as a lime contamination in the product; the calcination conditions of the

872

A. S. Bhatti et al.

dolomite must therefore be such that all carbon dioxide is removed, while avoiding overburning which would result in poor slakeability. Gilpin & Heasman (1952)listed some of the compounds formed during calcination of the dolomite, with their stabilities in seawater. Many of the impurities are difficult to remove because they are very finely distributed, and inevitably the dolomite introduces SiO 2, Fe203, A120 3 and CaO impurities, with perhaps lower levels of others such as MnO and B203. Various methods of chemically purifying the magnesia are being investigated and these could result in extremely pure grades (Gilpin & Spencer, 1972).

Impurities from seawater Lime. Seawater contains calcium bicarbonate which reacts with calcium hydroxide to precipitate calcium carbonate. Ca(HCO3) 2 + Ca(OH) 2 --, 2CaCO 3 + 2H20 This, being insoluble, appears as a lime impurity in the product at a concentration of about 3% (Gilpin & Heasman, 1963). Earlier methods of minimizing this impurity have been abandoned in favour of the newer acid pretreatment of seawater whereby sulphuric acid is injected to lower the pH of the seawater from about 8.5 to 3.5-4.0. The sulphuric acid reacts with the calcium bicarbonate to form calcium sulphate and carbon dioxide. The carbon dioxide is removed in a wooden desorption tower. Here, the water is broken into small droplets and the carbon dioxide escapes; with this method the lime content of the product can be reduced by 0 . 4 - 0 . 5 % (Gilpin & Heasman, 1963). Silica. Most of the silica impurity comes from sand suspended in seawater. Although seawater inlets are sited in sheltered (Hall & Spencer, 1973) areas, some very fine sand is inevitably drawn into the system and in heavy storms silica contamination rises. A way of reducing the silica contamination would be to absorb the silica onto a precipitate of either magnesium hydroxide or calcium carbonate in the hydro-treaters prior to the main reaction. Boron. Boron is present in small amounts in seawater as boric acid (Svendup & Johnson, 1942). The precipitates of magnesium hydroxides have a high capacity for absorbing boron and the final concentration in the oxide can be as high as 0.4% (seawater contains 15 p.p.m, expressed as B203). This boron contamination can be reduced by using an excess of lime in the precipitation but this also increases the lime contamination; however, it has been used without increasing the lime contents beyond acceptable limits. Alternative methods have been studied, and one commonly used is to remove the boron at the dead-burning stage by volatilization. The rate can be accelerated by the addition of certain alkali metal salts, e.g. potassium hydroxide. The use of such methods may reduce the boron content of seawater magnesias to 0.03% (Gilpin & Spencer, 1972). Calcium sulphate. As the precipitation of magnesium hydroxide proceeds, the concentration of Ca 2+ increases until after ~75% of the magnesium hydroxide has been removed; the theoretical solubility limit of calcium sulphate is then exceeded and hence the possibility arises of gross calcium sulphate contamination. Fortunately, calcium sulphate forms stable super-saturated solutions and this prevents the calcium sulphate contamination problems.

Magnesia from seawater Calcination--effect of time

873

The effect of the time of heating on magnesium hydroxide can be demonstrated by quoting some of our own results. Adsorption-desorption measurements of nitrogen at 77.4 K were carried on the parent as well as on the thermal decomposition products of the hydroxide, which were formed by heating at 350~ for 1, 2, 3, 4, 5, 6 and 7 h in both wet and dry nitrogen. The BET surface areas of the products are shown as a function of the calcination time in Fig. 3. The most striking result is that obtained by heating in dry nitrogen for 6 h; the area increases about 13-fold over that of the parent material. Heating at 350~ for longer than 6 h causes a gradual decrease in surface area, obviously due to sintering forces (Laminy, 1969). Calcination of the magnesium hydroxide is also affected by the presence of a constant water vapour pressure. Fig. 3 shows that the specific surface areas of the various products obtained by decomposition in the presence of water vapour are appreciably lower than the corresponding areas obtained in dry nitrogen. However, the general trend of variation of area with respect to the calcination time is approximately the same in dry and wet nitrogen. In both cases, the area increases with time of calcination to reach a maximum. For decomposition in the presence of water vapour, the maximum is

400 9 dry nitrogen 0 wet nitrogen

300

to ~u 03

200

1 O0

I 1

l 2

I 3

I 4

I 5

I 6

I 7

C a l c i n a t i o n t i m e (h)

F]G. 3, The development of surface area in dry and 'wet' nitrogen at 350~ The 'wet' nitrogen was obtained by passing it through a series of Dreschel bottles filled with water and glass rods.

874

A. S. Bhatti et al.

shifted to a longer calcination time. The efficiency of removal of the water vapour formed on decomposition of Mg(OH) 2 affects the energy of activation which therefore varies appreciably from sample to sample of Mg(OH) 2 with different surface areas and porosity. For the decomposition of Mg(OH)2 in wet nitrogen the water vapour concentration gradients along the capillary passages in the product will be smaller than in dry nitrogen, giving lower rates of nucleation of the newly formed magnesium oxide crystallites. Differences of similar type are found for magnesium and calcium hydroxides decomposed in air with vacuum conditions (Glasson, 1956, 1963).

Calcination--effect of temperature
Our own results can be used also to show the effect of temperatures on the activity of magnesium oxide produced. Results of experiments on the adsorption of nitrogen at 77.4 K on samples of magnesium hydroxide calcined in air for 5 h at various temperatures are shown in Fig. 4. The appearance of a maximum in the specific area-temperature of dehydration curves has been frequently found in the decomposition of many crystalline hydroxides used as parent materials in the preparation of active solids. This maximum generally is attributed to the existence of more than one process during thermal treatment, namely, decomposition, recrystallization and sintering. It is known that when dehydration occurs at lower temperatures the former two processes predominate; whereas above these temperatures sintering becomes predominant. At temperatures near that producing the maximum surface area, the processes of activation and sintering overlap (McBain & Dunn, 1948).

300

200

1 O0

I 200

I 400

I 600

Calcination t e m p e r a t u r e (~

F1G. 4. The effect of calcination on the adsorption properties of magnesium oxide.

Magnesia from seawater


REFERENCES

875

ANDERSON P.J., HORLOCK R.F. & AVERY R.G. (1965) Proc. Brit. Ceram. Soc. 3, 33. BANEaJEE M. (1969) PhD Thesis, University of Sheffield. BUDNIKOVP. (1930) Z. Anorg. Allgem. Phem. 190, 79. BOYNTON R.S. (1966) Chemistry and Technology of Lime and Limestone, p. 352. Interscience, J. Wiley, London and New York. EUBANKW.J. (1951) J. Am. Ceram. Soc. 34, 225. FRICKE R. & LUKE J. (1935) Z. Electrochem. 41,174. FRICKER. (1936) Z. Electrochem. 42, 881. GILPIN W.C. & HEASMANN. (1952) Refractory J. July, 302. GILPIN W.C. HEASMANN. (1963)RefractoryJ. June, 214. GILPIN W.C. & SPENCER D.R.F. (1972) Refractory J. April, 4. GINQUE W. (1949) J. A m. Chem. Soc. 71, 3192. GLASSON D.R. (1956) J. Chem. Soc., 1506. GLASSON D.R. (1958) J. Appl. Chem., Lond. 8, 798. GLASSON D.R. (1960) J. AppL Chem., Lond. 10, 38. GLASSON D.R. (1961) J. AppL Chem., Lond. l 1, 24. GLASSON D.R. (1963a) J. Appl. Chem., Lond. 13, 111. GLASSON D.R. (1963b) J. Appl. Chem., Lond. 13, 119. GLASSON D.R. (1964a) J. Appl. Chem., Lond. 14, 121. GLASSON D.R. (1964b) J. Appl. Chem., Lond. 14, 125. GLASSON D.R., O'NEILL P. 8~, SHEPPARD M.A. (1968a) J. Appl. Chem., Lond. 18, 198. GLASSON D.R., HUFF E.G., JONES J.A. & SHEPPARD M.A. (1968b) J. AppL Chem., Lond. 18, 204. GREEN J. (1969) PhD Thesis, University of Sheffield, UK. GREGG S. & SING K.S.W. (1951) J. Phys. Colloid Chem. 55, 592. GUILLIAT F. & BRETTN.H. (1971) Phil. Mag. 23, 647. HALL R.J. & SPENCER D.R.F. (1973) Interceram 22, 212. HEDVALLJ. ( 1937) Reaktionfunhigkeit Fester Staffe, Leipzig. HILL K.J. (1951) PhD Thesis, University of London, UK. HUTX~GG.F. ( 1941) Kolloid Z. 97, 281. HVTTIG G.F. (1942)KolloidZ,98, 263. LAMB A. & WEST C. (1946)J. A m. Chem. Soc. 62, 3 176. LAMINY J.V. (1969) PhD Thesis, University of Sheffield. McBAIN J. & DUNN R. (1948) J. Colloid Sci. 3, 303. RIDDELL D.J. (1969) PhD Thesis, University of Sheffield. SING K.S.W. (1949) PhD Thesis, University of London, UK. SVENDUP H.V.M. & JOHNSON F. (1942) The Oceans. Prentice Hall Inc., New York. TAYLOR K. ~/. WELLS L. (1938)J. Research, NationalBureau of Standards 21, 1334. TORGESON E. 8~. SCHAMERZ. (1948) J. Am. Chem, Soc. 70, 2156. ZETTLEMAYERA.C. & WALKER W. (1947) Ind. Eng. Chem. 39, 69.

You might also like