Elsevier

Ceramics International

Volume 41, Issue 10, Part B, December 2015, Pages 13998-14007
Ceramics International

A systemic investigation on the hydroxylation behavior of caustic magnesia and magnesia sinter

https://doi.org/10.1016/j.ceramint.2015.07.012Get rights and content

Abstract

Caustic magnesia (CM) and magnesia sinter (MS or hard-burnt magnesia) are the main sources of MgO employed in industry. Both are obtained from the calcination of Mg(OH)2 or MgCO3, however, due to differences in their manufacture, they display significant distinct physical–chemical properties. MS is produced at temperatures above 1500 °C and shows large crystal size and low inner micro–meso pores content, whereas CM is attained at lower temperatures (600–900 °C) and has high specific surface area, micro–meso pores content and chemical reactivity. Their hydroxylation behavior is also different, as it takes MS several hours-days to react completely with water, while the CM reaction achieves total conversion after few minutes. Because of their technological importance, a deeper understanding of the processes involved in hydroxylation reactions and how particles׳ characteristics affect them is required. This study relates the characterization of MS and CM particles with their hydroxylation reactions in aqueous suspensions through thermogravimetry, X-ray diffraction, nitrogen adsorption, and scanning electron microscopy.

Introduction

Magnesium oxide (MgO) or magnesia, is one of the most important raw materials for industrial processes and technological applications. It can be produced by two main commercial routes [1], namely calcination of brucite (Mg(OH)2), produced from seawater or underground brine, and calcination of natural brucite or magnesite (MgCO3).

Seawater naturally contains a 1–3.5 wt% portion of dissolved ionic species, such as Cl, (SO4)2−, Br, Na+, K+, Ca2+ and Mg2+ and by means of water evaporation, this salt concentration can be raised to above 50 wt%. After the selective precipitation of less soluble salts, such as CaCO3 (0.0004 g L−1) and CaSO4 (2.5 g L−1), the resulting solution shows a higher content of highly soluble magnesium salts (MgCl2=540 g L−1 and MgSO4=350 g L−1) [1], [2]. A similar system can be obtained through the injection of plain water into underground deposits of salt-rocks and brine recovering [1], [2]. The next step is to treat Mg2+-rich brine with an alkaline compound, as NaOH or CaO to precipitate Mg(OH)2 (or brucite) according to the expression [1], [2], [3]Mg2++2Cl+Na++OH→Mg(OH)2(Precipitate)+NaCl(Dissolved)After washing, filtration and drying steps, Mg(OH)2 is recovered and calcined in rotary kilns to produce MgO [4], [5], [6], [7], [8]Mg(OH)2260400oCMgO+H2O(Gas)

Magnesite is a general name employed to describe an abundant family of magnesium carbonates and hydroxycarbonates (magnesite=MgCO3; artinite=MgCO3·Mg(OH)2·3H2O; nesquehonite=MgCO3·3H2O; hydromagnesite=4MgCO3·Mg(OH)2·4H2O) found in many countries [1]. Due to its sedimentary origin, magnesite is usually found in open pits or low-depth deposits. Depending on the geological activity of the area, the magnesite ore may contain impurities (dolomite, talc, calcite, clay, and sand), which can be removed by filtration and separation standard processes. After milling and granulometric classification, magnesite is calcined in rotary kilns (700–1700 °C). During this thermal treatment, it undergoes decarbonation and produces MgO, according to the general expression [1], [9], [10]MgCO3500750oCMgO+CO2(Gas)

After these operations, grades of MgO content above 98–99% are used in the production of refractory bricks or castables, whereas those of lower MgO content are employed as pH correctors in soil and water treatments and as mineral supplement in animal and human feeding.

The final step of routes 1.1 and 1.2 (expressions (2), (3)) is a calcination process, through which the hydroxylated (brucite) or carbonated (magnesite) precursors are transformed into MgO. Depending on the conditions employed (final temperature, residence time, atmosphere, and impurities), different types of polycrystalline materials can be obtained [1], [4], [5], [6], [7], [8], [9]. The MgO grade known as caustic magnesia (CM) is produced through the rapid calcination of brucite or magnesite at temperatures of up to 800–900 °C in rotary and Herreshoff-type kilns [8], [9], [11]. During calcination, such precursors undergo a series of topotactic changes before their total conversion into MgO (Fig. 1) [4], [5], [6], [7], [12], [13]. The loss of OH and (CO3)2− ions (as H2O and CO2) causes a significant volumetric shrinkage in their crystalline lattices (particularly at the c axis), which increases density, and the withdrawal of gases generates microcraked porous surfaces. The external morphology of the freshly produced CM particles is similar to the shape of the precursor ones and, because of their high specific surface area (usually above 10 m2 g−1) and chemical reactivity, they can be employed in several chemical processes [8], [9].

On the other hand, CM characteristics are unsuitable for uses that involve sintering for the production of dense parts, and a further processing is required. Therefore, CM particles are compacted into 10–30 mm beads (agglutinants, as carboximetil-cellulose, and sintering aids, B2O3 and Fe2O3, are often added previously [14], [15]) and sintered at temperatures of 1500–1700 °C, for up to 1–3 h. After cooling, the compacts produced (whose relative density is above 90%) are crushed and classified in the desired granulometric range. As they show very low porosity and chemical reactivity, they are called hard-burnt magnesia or magnesia sinter (MS) and are almost exclusively used in the production of refractory bricks and castables [1], [14], [16].

CM and MS have been the focus of many investigations for over half a century and an aspect that has drawn plenty of attention is their hydroxylation (or hydration) behavior [17], [18], [19], [20], [21], [22], [23], [24], [25], [26]. Depending on the application aimed at, MgO hydroxylation can be either valuable or undesired and, in both cases, the knowledge of how this process occurs can be highly useful. The hydroxylation process must be efficient for the production of fertilizers [1], [18], animal and human nutrition [1], [27], water-treatment input [1], [28], medicaments [1], [27], [28], fast-hardening Sorel and MgO-based cements [1], [29], [30], [31] and anti-flame agents for polymer composites [32], [33], [34] and generates particles of controllable size distribution, specific surface area and shape at competitive costs. On the other hand, it must be avoided or at least minimized in applications to refractory [10], [13], [14], [16], [35], [36], [37], [38], [39], [40] and abrasive [1] materials and optic and electronic devices.

MgO hydroxylation is a dissolution–precipitation reaction (Eqs. ((1), (2), (3)) [20], [21], [22], [23], [24], [27]. In the presence of water, MgO surfaces initially become protonated and positively charged due to their strong ionic character, according to the following equation [23], [24], [37]:MgO+H2O→MgOH++OHThe presence of positive charges enables the dissociation of the intermediate compound MgOH+MgOH++OH→Mg2++2OHThe low solubility of Mg(OH)2 in pH levels that range from neutral to highly alkaline favors the precipitation of Mg(OH)2 on the nearest MgO surfaces [3], [22], [27], [28], [41]Mg2++2OH→Mg(OH)2

The characteristics of the reacting medium (temperature and solid load) and the MgO source (single crystal or polycrystalline grains) affect the rates of steps 4–6 differently [14], [16], [18], [22], [25], [26]. When the dissolution steps (expressions (4), (5)) occur gradually (for instance, in very diluted suspensions of monocrystalline particles at temperatures below 20 °C), the precipitation (expression (6)) necessarily slows down and Mg(OH)2 is formed orderly as a dense thin layer on the surface of the MgO particles [16], [22], [25], [26], [37], [42]. This layer behaves as protective coating preventing further MgO dissolution and halting the process. Therefore, under such conditions, hydroxylation occurs along several days or weeks [30]. On the other hand, for high solid load suspensions of polycrystalline particles at temperatures above 20 °C, dissolution occurs more rapidly. Consequently, the massive and exothermic precipitation of Mg(OH)2 particles generate a thick and porous brucite layer [11], [12], [16], [25], [43]. Because of the large density mismatch (ρMgO=3.5 g cm−3 and ρMg(OH)2=2.4 g cm−3), this extra volume generated cannot be suitably roomed and induces compression stresses that crack the Mg(OH)2 layer and the MgO beneath it. The exposure of unreacted surfaces and the heat released restart and speed up the hydroxylation process up to the total consumption of MgO after few hours of reaction [16], [30].

The mechanisms described above enable the MgO conversion to Mg(OH)2 to be maximized by (i) the replacement of particles of monocrystalline magnesia by polycrystalline ones [14], [15] and (ii) use of thermally isolated reactors (autoclaves, for instance) for keeping the heat released during the reaction and using it to accelerate the process [16], [25], [26]. Alternatively, to prevent hydroxylation, several studies have proposed anti-hydroxylation (or anti-hydration) additives particularly useful to refractory castables. In general, they reduce the dissolution rate or inhibit further Mg(OH)2 precipitation by forming an insoluble coating on the surface of the MgO particles [14], [15], [31], [35], [36], [38], [44].

This manuscript addresses an investigation on the hydroxylation behavior of different MgO sources (CM and MS) and the way the previously mentioned kinetic aspects interact with their different physical–chemical characteristics (composition, density, specific surface area, pore content and average size). The results of hydroxylation tests conducted in aqueous suspensions were evaluated by thermogravimetry, X-ray diffraction and scanning electron microscopy.

Section snippets

Characterization of the as-received magnesia sources

The magnesia sources tested (Table 1) were magnesia sinter (MS) (M30 High purity, Magnesita Refratários S.A., Brazil, Fig. 2a and b) and caustic magnesia (CM) (Q-MAG-AR200, Magnesita Refratários S.A., Brazil, Fig. 2c and d).

Characterization of SM and CM particles (after 2 h at 200 °C under vacuum): (a) solid density (ρ, Helium pycnometer, Ultrapyc 1200e, Quantachrome Instruments, USA); (b) specific surface area (SSA, BET method, N2 adsorption, P/P0 ranging from 0.05 up to 0.3), total pore volume

As-received magnesia sources

Both as-received magnesia sources were dense (relative densities above 98.5%) and comprised by periclase (Fig. 4a and bi) and showed MgO contents higher than 98.5%, similar CaO/SiO2 ratios and average particle sizes (Table 1). Despite their chemical similarity, the particles morphologies were different (Fig. 2).

MS is usually produced by crushing and ball-milling pellets of calcined magnesite (originally of 10–50 mm diameter) previously sintered at 1700 °C for 3–5 h [1]. This process rebuilds the

Final remarks

The different MgO sources tested (sinter, MS, and caustic, CM) have showed distinct hydroxylation mechanisms. For MS, hydroxylation occurs near the surface of the MgO particles, where Mg(OH)2 forms precipitates. The large volumetric expansion occurred in this region begins to compress and crack the surface and exposes unreacted areas and restarts the process. Due to particles׳ low specific surface area and pore content, the reaction occurs very slowly at temperatures of up to 50 °C. CM, on the

Acknowledgments

The authors would like to acknowledge Brazilian Research Foundations FAPESP (2010-19274-5) and CNPq (470981/2011-3 and 306036/2011-8) for supporting this research and Magnesita Refratários S.A. (Brazil) for kindly supplying the samples of raw materials. They are also indebted to Wagner R. Correr MSc. (Electron Microscopy Laboratory of Center for Technology of Hybrid Materials, CTMH) for the SEM images.

References (48)

  • R. Salomão et al.

    Citric acid as anti-hydration additive for magnesia containing refractory castables

    Ceram. Int.

    (2011)
  • T.M. Souza et al.

    Systemic analysis of MgO hydration effects on alumina-magnesia refractory castable

    Ceram. Int.

    (2012)
  • J.P. Liu et al.

    Modeling hydration process of magnesia based on nucleation and growth theory: the isothermal calorimetry study

    Thermochim. Acta

    (2012)
  • S. Chatterji

    Mechanism of expansion of concrete due to the presence of dear-burnt CaO and MgO

    Cem. Concr. Res.

    (1995)
  • M. Sutcu et al.

    A microstructural study of surface hydration on a magnesia refractory

    Ceram. Int.

    (2010)
  • G. Ye et al.

    Hydration of hydratable alumina in the presence of various forms of MgO

    Ceram. Int.

    (2006)
  • R. Salomão et al.

    Hydrotalcite synthesis via co-precipitation reaction using MgO and Al(OH)3 precursors

    Ceram. Int.

    (2011)
  • M.A. Shand

    The Chemistry and Technology of Magnesia

    (2006)
  • G.E. Seil

    Study of literature on separation of magnesia from lime in dolomite and similar materials

    J. Am. Ceram. Soc.

    (1943)
  • M.F. Goudge

    Brucite magnesia

    J. Eur. Ceram. Soc.

    (1944)
  • W.R. Eubank

    Calcination studies of magnesium oxides

    J. Am. Ceram. Soc.

    (1951)
  • P.J. Anderson et al.

    Thermal decomposition of magnesium hydroxide

    Trans. Faraday Soc.

    (1962)
  • V.A. Phillips et al.

    Relations among particle size, shape and surface area of Mg(OH)2 and its calcination product

    J. Am. Ceram. Soc.

    (1978)
  • J. Green

    Review: calcination of precipitated Mg(OH)2 to active MgO in production of refractory and chemical grade MgO

    J. Mater. Sci.

    (1983)
  • Cited by (13)

    View all citing articles on Scopus
    View full text