Quick viewing(Text Mode)

Hydration Mechanisms of Magnesia-Based Refractory

Hydration Mechanisms of Magnesia-Based Refractory

HYDRATION MECHANISMS OF MAGNESIA-BASED

REFRACTORY

by

SHUXIN ZHOU

B. Sc., Tianjin University, China 1986 M. Sc., Shanghai Institute of , China 1989

A THESIS SUBMITTED IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF APPLIED SCIENCE

m

THE FACULTY OF GRADUATE STUDIES

(Metals & Materials Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA

December 2004

Shuxin Zhou, 2004 ABSTRACT

Hydration of magnesia-based refractory bricks could occur during storage, during drying

after installation, or in service, and the hydration would cause damage to refractory bricks

and furnace linings. In order to understand the hydration mechanisms of magnesia-based

refractories, three types of bricks were chosen: magnesia, magnesia-spinel and magnesia-

chrome bricks, and hydration tests were performed at 60 to 130°C in 98% relative

humidity, water, and steam. The variation of the modulus of elasticity (MOE),

determined by the impulse excitation technique (IET), as well as apparent porosity, air permeability, pore size and pore size distribution were correlated with the hydration data.

The phase compositions and microstructure modifications were also studied on selected specimens by XRD and SEM/EDS. Based on the experimental results, a hydration model of "cylindrical pore model" was established and a hydration mechanism was suggested.

The hydration takes place in three stages. In the first stage, which is controlled by chemical reaction, a film of brucite forms and MOE quickly increases. During the second stage, which is controlled by diffusion, the MOE gradually reaches a maximum value followed by a slow, decrease due to the formation of cracks on the film and weakening of grain boundaries until the MOE reaches the initial value. At this point, the third stage, corresponding to "dusting", starts to take place until the disintegrates. The results also indicate that the hydration rate increases with rising temperature and the CaO/Si02 ratio. The variations in permeability and porosity are opposite to that in MOE. A nondestructive method - IET to assess the hydration degree of magnesia-based refractory bricks was proposed.

ii TABLE OF CONTENTS

ABSTRACT .' ii

TABLE OF CONTENTS iii

LIST OF TABLES v

LIST OF FIGURES vi

LIST OF ABBREVIATIONS AND SYMBOLS ix

ACKNOWLEDGMENTS xiv

1. INTRODUCTION 1

2. LITERATURE REVIEW 4

2.1. Hydration of Magnesia-Based Refractories 6 2.2. Hydration of Polycrystalline Magnesia 7 2.3. Hydration of Pure Magnesia in Liquid Water 9 2.4. Hydration of Magnesia in Water Vapor 10 2.5. Kinetics of Hydration 12 2.6. Hydration of Refractory Grade Magnesia 17 2.6.1. Effect of the Physical Properties of Magnesia 17 2.6.2. Effect of Impurities 18

2.6.2.1. CaO 18

2.6.2.2. Si02 19

2.6.2.3. CaO/Si02 Ratio 20

2.6.2.4. A1203 20 2.6.2.5. 20 2.6.2.6. B2O3 22

2.6.2.7. Cr203 22 2.6.2.8. Other Oxides 22

3. OBJECTIVES 23

4. EXPERIMENT PROCEDURES AND METHODOLOGY 25

4.1. Experimental Materials 25 4.2. Samples Preparation Procedures 27

iii 4.3. Experimental Methods ..28 4.4. Hydration Studies 29 4.4.1. Hydration Test Conditions 29 4.4.2. Hydration Equipment and Procedures 29 4.4.3. Calculations 30

4.5. Modulus of Elasticity 33 4.6. Permeability 36 4.7. Pore Size and Pore Size Distribution 38 4.8. Mineralogy Studies 40 4.9. Microstructure Studies 41 5. RESULTS AND DISCUSSION 42

5.1. Degree of Hydration under Different Testing Conditions 42

5.1.1. Hydration of Magnesia-Based Bricks 42 5.1.2. Magnesia-Chrome Bricks 54 5.2. Modulus of Elasticity (MOE) versus Hydration Degree 61 5.3. Influence of Brucite on MOE Variation 68 5.4. Influence of Hydration on Pore Size Distribution 76 5.5. Influence of Hydration on Permeability 78 5.6. A Simplified Kinetic Analysis 80 5.7. The Mechanism of Hydration of Magnesia-Based Refractory Bricks 90

6. CONCLUSIONS 92

7. FUTURE WORK 95

8. REFERENCES 96

iv LIST OF TABLES

Table 4-1. Properties of the experimental magnesia and magnesia- spinel bricks 26

Table 4-2. The characteristics of experimental magnesia-chrome bricks 26

Table 4-3. A summary of hydration tests 31

Table 4-4. Relations between porosity P and Young's modulus E [41] 35

Table 5-1-1. The compositions of binding phases for sample MS2, Ml, and M2 from EDS analysis •. 51

Table 5-1-2. EDS results of the binding phase in the MCI and MC2 samples, wt% 58

Table 5-3-1. The compositions of different grains in MC3 shown in Fig. 5-3-2 70

Table 5-3-2. Apparent porosity of MC3 after hydration in steam at 109°C 76

Table 5-6-1. The reaction rate constants, K, of the Ml samples 87

Table 5-6-2. The reaction rate constants, K*\0, of samples MSI, MS2 and M2 in steam and humidity at different temperatures 88

Table 5-6-3. The rate constants, K*\0, for the magnesia-chrome bricks in steam 88

Table 5-6-4. Activation energies, E (kJ/mol), for the hydration of all the experimental bricks in steam and humidity 90

Table 5-6-5. Activation energy for the hydration of magnesia 90

v LIST OF FIGURES

Figurel-l. Hydrated magnesia brick 2

Figure 2-1. Crystal structure of periclase 5

Figure 2-2. Crystal structure of brucite 6

Figure 2-3. Hydration curves for the magnesia single crystals, where a is the hydration degree of magnesia 8

Figure 2-4. Hydration model for polycrystalline magnesia 8

Figure 2-5. Brucite participated at 60°C in (a) fast nucleation process at pH > 13 and (b) slow nucleation process at pH ~ 10 10

Figure 2-6. Phase diagram of system CaO-MgO 19

Figure 2-7. Phase diagram of system FeO-MgO 21

Figure 4-1. Schematic representation of the specimen preparation from a brick, a) for bars used in hydration studies and MOE measurement and (b) for cylinders for permeability and pore size and pore size distribution 26

Figure 4-2. The impulse excitation technique apparatus 34

Figure 4-3. The VacuPerm Decay Permeameter (a) schematic diagram and (b) instrument 37

Figure 4-4. The operation principle for measuring pore diameters 40

Figure 5-1-1. Weight increases versus time for hydration in steam at 121°C 44

Figure 5-1-2. Weight increases versus time for hydration in steam at 109°C 44

Figure 5-1-3. Weight increases versus time for hydration in steam at 100°C 45

Figure 5-1-4. Weight increases versus time for hydration in 98% humidity at 80°C 45

Figure 5-1-5. Weight increases versus time for hydration in 98% humidity at 60°C 46

Figure 5-1-6. Weight increases versus time for hydration in water at 100°C 46

Figure 5-1-7. Weight increases versus time for hydration in water at 80°C 47

vi Figure 5-1-8. Weight increases versus time for hydration in water at 60°C 47

Figure 5-1-9. Backscatter electron micrograph of the MSI sample 48

Figure 5-1-10. EDS results of the bonding phase in the MSI matrix 49

Figure 5-1-1.1. The XRD diffractograms for the unused MSI and MS2 samples (P = pericalse and S = spinel) 50

Figure 5-1-12. Backscatter electron micrograph of the MS2 sample 51

Figure 5-1-13. Backscatter electron micrograph of brick Ml, binding phase (light gray) and periclase (dark gray) 52

Figure 5-1-14. The XRD diffractograms for the unused Ml and M2 (P = pericalse) 53

Figure 5-1-15. Backscatter electron micrograph of the M2 brick structure 53

Figure 5-1-16. Weight increases versus time for hydration in steam at 109°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 55

Figure 5-1-17. Weight increases versus time for hydration in steam at 121°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 56

Figure 5-1-18. Weight increases versus time for hydration in steam at 130°C 57

Figure 5-1-19. XRD diffractogra for MCI- the unused sample. P = pericalse (MgO) and

C = complex spinel (Mg, Fe)(Cr, Al, Fe)204 58

Figure 5-1-20. Hydrated MgO versus CaO/Si02 ratio for the hydration of the magnesia- spinel and magnesia bricks in water at 60°C 60

Figure 5-1-21. Hydrated MgO versus CaO/Si02 ratio for the hydration of the magnesia- chrome bricks in steam at 109°C 60

Figure 5-2-1. MOE variation versus time for hydration in steam at 121°C for magnesia- spinel and magnesia bricks 61

Figure 5-2-2. MOE variation versus time for hydration in steam at 109°C for magnesia- spinel and magnesia bricks 62

Figure 5-2-3. MOE variation versus time for hydration in steam at 100°C for magnesia- spinel and magnesia bricks 63

Figure 5-2-4. MOE variation versus time for hydration in 98% humidity at 80°C for magnesia-spinel and magnesia bricks 64 Figure 5-2-5. MOE variation versus time for hydration in 98% humidity at 60°C for magnesia-spinel and magnesia bricks 64

Figure 5-2-6. MOE variation versus time for hydration in steam at 109°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 65

Figure 5-2-7. MOE variation versus time for hydration in steam at 121°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 66

Figure 5-2-8. MOE variation versus time for hydration in steam at 130°C for the magnesia-chrome bricks MCI, MC2, MC3, MC5 and MC6 : 67

Figure 5-3-1. Weight increase and MOE variation versus the time for hydration in steam at 109°C for brick MC3 69

Figure 5-3-2. SEM image for the fracture section of the non-hydrated MC3 sample....70

Figure 5-3-3. The film of brucite on the wall of a pore after hydration for 8 hours 71

Figure 5-3-4. The fractograph of the MC3 sample after hydration for 24 hours 74

Figure 5-3-5. SEM images of the sample MC3 after hydration for 38 hours in steam at 109°C, (a) the fracture surface and (b) the polished sample 75

Figure 5-4-1. The change in the PSD for different hydration periods in steam at 109°C for brick MC3. (a) smaller than 1 um pores and (b) larger than 1 um pores 77

Figure 5-5-1. Permeability versus time after hydration of brick MSI, MS2, Ml, and M2 in 98% relative humidity at 80°C 79

Figure 5-5-2. Permeability versus time after hydration of brick MCI, MC3, MC4 and MC6 in steam at 109°C 80

Figure 5-6-1. Hydration model-"cylinder pore model" 81

Figure 5-6-2. Hydration rate equations, R)(a) and R2(ot), versus time, t, for Ml samples, (a) the first stage hydration in steam at 121, 109, and 100°C, (b) the first stage hydration in humidity at 100, 80, and 60°C, (c) the second stage hydration in steam at 121, 109, and 100°C, (d) the second stage hydration in humidity at 100, 80, and 60°C 86

Figure 5-6-3. The plots of ln K versus 1/T for sample Ml 89

vin LIST OF ABBREVIATIONS AND SYMBOLS

Abbreviations

ASTM American Society for Testing and Materials

BSE Backscatter Electron

C2S Di- Silicate

EDS Energy Dispersive Spectroscopy

FCC Face Centered Cubic

IET Impulse Excitation Technique

MOE Modulus of Elasticity

PSD Pore Size Distribution

SEM Scanning Electron Microscopy

XRD X-Ray Diffraction Symbols

a fraction of reacted magnesia

A cross-section area

A frequency factor

b width of a bar specimen

CMgo chemical content of MgO in a brick

D diameter of a pore

d0 median pore diameter dD range of pore sizes

DT diffusion coefficient at temperature T

E activation energy

E Young's modulus

Eo modulus of elasticity before hydration

modulus of elasticity after hydration for n hours fd flow rates through dry samples

ff fundamental resonant frequency in flexure fw flow rates through wet samples h thickness of a brucite film

K rate constant k ratio of reaction constant to particle density in Eq.(2 k permeability k = pbrick M^gc/pbrucite MH20

K' constant in Eq. (2-9) ko constant in Eq. (2-7)

k ' = P brick

K2 'MMg(OH) = PMSQ 2

P brucite ' M MgO

L length of a bar specimen

L length of a sample in the macroscopic flow direction in Eq. (4-5)

Lo original length of cylindrical pores

m mass of a bar specimen

m reacted MgO

M/-/20 molecular mass of H2O

MMg(OH)2 molar mass of Mg(OH)2

MMgo molecular mass of MgO

n constant in Eq. (2-10)

N constant in Eq. (2-12)

n number of hours for hydration

N number of open pores in a sample p water vapor pressure

P apparent porosity p* 0.3 times saturation pressure po saturation vapor pressure

ps saturation vapor pressure

Q volumetric flow rate

xi R gas constant r cylindrical pore radius

r0 mean original radius of magnesia particles in Eq. (2-12)

r0 original radius of cylindrical pores

R,(a) = (l+kia),/2 -1

,/2 2 2 R2(a) = [(1 +k,a) - ((1 +k,a)- (l-k2) + k2f ]

S reaction area t time

T absolute temperature t thickness of a bar specimen

Ti correction factor in Eq. (4-3) u reaction rate per unit area v reaction rate

2 V0 = Ttr0 L0

V Sampie sample volume in Eq. (5-8) w average MgO mass around each pore

W0 sample weight before hydration

W„ sample weight after hydration for n hours y surface tension of wetting liquid

8 thickness of brucite layer

AE variation of modulus of elasticity

AP pressure difference between both ends of a sample

AP differential pressure required to remove wetting liquid from a pore

xii A W weight increase rj fluid viscosity

6 contact angle of a wetting liquid on a sample material pbrick brick density

Pbncite brucite density

PMgO magnesia density ACKNOWLEDGEMENTS

This thesis could not have been written without my supervisors, Dr. Tom Troczynski and

Dr. George Oprea. Throughout my study and the research work, they provided invaluable guidance and support. I would like to express my sincere gratitude to them.

I would also like to thank the members of UBCeram group of UBC: Carmen Oprea,

Guotian Ye, and Dr. Ahmad Monshi for their assistance. I want to convey my special thank to Ms. Mary Mager for her assistance with SEM/EDS analysis. I am grateful to all those who have provided help during the period of my study and the research.

I would like to thank all the industrial sponsors of the research consortium on refractories at UBCeram and the Natural Sciences and Engineering Research Council of Canada for their financial support.

Last but not least, I would like to thank my parents for their support and encouragement and my wife and daughter for their love and sacrifice through my study in UBC.

xiv 1. INTRODUCTION

Magnesia ( ) is an important raw material for the refractory industry because magnesia possesses a high corrosion resistance to basic and molten metals.

Magnesia-based refractory bricks have been widely used in metallurgical and cement furnaces, however, as a main component of the magnesia-based refractory bricks, magnesia reacts with water (both liquid and vapor) very easily. This reaction of hydration produces magnesium hydroxide (brucite) according to the following reaction:

MgO + H20 (liquid or vapor) = Mg(OH)2 (1 -1)

Due to the difference in density between magnesia (3.58 g/cm3) and brucite (2.36 g/cm3), when magnesia converts to brucite, the volume increases by factor of 2.2, with an accompanying 45% weight increase. This expansion will degrade the refractory bricks and destroy furnace linings; therefore, the hydration of magnesia-based refractory bricks has been of great concern for users of these materials, where the risk of this reaction taking place existed.

The hydration of magnesia-based refractories can occur during storage, during drying after installation, and during service, and generally in any environment where water exists or is generated. For example, in non-ferrous furnaces, water-cooling assists the performance of bricks in service, and water condensation around the cooling pipes or leakage from the cooling equipment often occurs and produces an environment with high water vapor pressure inside of the high temperature refractory linings. This can cause a serious destruction of the refractory linings due to hydration.

1 Figure 1-1 shows a hydrated burned brick, which was in service in an electrical smelting furnace only for one month. Due to the hydration of magnesia, extensive cracking occurred in the bricks [1].

Fig. 1-1. Hydrated magnesite brick [1].

There are qualitative and quantitative methods to assess the hydration of magnesia-based refractories. The American Society for Testing and Materials (ASTM) standard method for the Hydration Resistance of Basic Bricks and Shapes is a qualitative method to evaluate the hydration resistance of basic bricks by visual inspection after autoclave

2 treatment, using a scale of four rates. In addition to the visual inspection, weight increase is a quantitative method.

The hydration of pure has been extensively studied, but these studies mainly focused on pure caustic magnesias obtained by between 400 and

1200°C of Mg(OH)2 or MgCCb and main topics were: thermodynamic calculations for

Mg(OH)2 and MgO, influences of the preparation process of magnesium oxide on hydration, and the hydration kinetics. Hydration of magnesia clinkers was also studied, especially the effects of different chemical compositions on the hydration rate. However, there is scarce literature about hydration and its effects on complex magnesia-based refractory bricks. The present work focuses on the study of the hydration behavior for magnesia-based refractory bricks, in correlation with other properties, such as pore size, pore size distribution, permeability, and modulus of elasticity measured by the impulse excitation technique. Furthermore, the hydration kinetics is discussed and the hydration mechanisms are suggested.

3 2. LITERATURE REVIEW

It is known (Eq.(l-l)) that magnesium oxide can react with water (liquid and vapor) to produce magnesium hydroxide (Mg(OH)2).

At 25°C and under atmospheric pressure, the enthalpy and Gibbs free energy of the reaction (1-1) are AH0 = -9.09 kcal/mol and AG°= -6.68 kcal/mol, respectively. When water is in the gaseous state at 100°C and one atmosphere pressure, AH° = -19.43 kcal/mol and AG° = -6.01 kcal/mol [2].

Periclase, which is the only crystalline form of magnesium oxide, has a NaCI type (cubic) structure (Fig. 2-1), consisting of two interpenetrating face-centered cubic (FCC) lattices of magnesium and . Both cations and anions are in octahedral sites, with a coordination of six. The unit cell lattice of the cubic structure is a = 4.21 lA. The theoretical density and Young's modulus are 3.585 g/cm and 162.7 GPa, respectively

[3].

Brucite, which is the crystalline form of magnesium hydroxide, is isostructural with the

layered compound Cdl2 (hexagonal). In brucite, the hydroxyl ions form a hexagonal close-packed stack. The Mg ions are in octahedral sites between alternate pairs of hydroxyl planes (Fig. 2-2). The cell parameters are a = b = 3.142 A and c = 4.766 A; the theoretical density is 2.377 g/cm3 [4].

According to the thermodynamic data [2], brucite begins to decompose into MgO and

H2O above 246°C. However, at this temperature the decomposition is very slow. Even at

4 300°C under vacuum, only about 95% of the magnesium hydroxide is decomposed after ten days [5], and above this temperature the dissociation rate increases rapidly [6].

Fig. 2-1. Crystal structure of periclase [3].

5 Fig. 2-2. Crystal structure of brucite [7].

2.1 Hydration of Magnesia-Based Refractories

Only one published technical paper [8] related to the hydration of magnesia-based refractory bricks could be found. The authors reported that the commercial magnesia- chrome bricks that were heated to 500°C underwent a significant decrease in the modulus-of-rupture value. Such strength loss is due to the decomposition of the

hydroxide in the brick. The thermal decomposition of Mg(OH)2 in magnesia-chrome bricks has the following steps: (i) recrystallization of MgO from a defective layer of hexagonal structure to cubic structure, (ii) the precursor (brucite) crystal cracking into small fragments, (iii) gradual desorption of water, (iv) shrinkage of the MgO crystal lattice. As a result, the produced MgO on the surface of periclase grains has a very high apparent porosity. About 80% strength loss in a magnesia-chrome brick of 59% MgO was due to the hydration of 5.8% of the original content of MgO during storage [8].

6 Due to scarce literature on hydration of magnesia-based refractory bricks, the hydration of polycrystalline magnesia and pure magnesia is introduced.

2. 2 Hydration of Polycrystalline Magnesia

The hydration process for pure polycrystalline magnesia clinkers under saturated water vapor at temperature range from 135°C to 200°C are different from that for single crystals [9]. For single crystal magnesia, there is a slow stage in the early hydration, especially at the low temperature and with hydration, a Mg(OH)2 layer is gradually formed. For polycrystalline magnesia, at the beginning of the hydration, there is an accelerated period, and the cracks caused by the brucite produced at grain boundaries gradually lead to the destruction of grain boundaries. For the magnesia polycrystals (2.00

- 3.36 mm in diameter), when the hydration degree is 15-20%, the grains start disintegrating until single crystals (about 50 um) are formed. According to Kitamura [9], this phenomenon is called "dusting". During the "dusting" process the hydration rate becomes higher, due to the formation of new surfaces available for hydration. This model of hydration for polycrystalline magnesias is shown in Fig. 2-4.

7 0 10 20 30 40 50 60 70 80 90 100 t(h) t(h)

Fig. 2-3. Hydration curves for the magnesia single crystals, where a is the hydration degree of magnesia [9].

Fig. 2-4. Hydration model for polycrystalline magnesia [9]. 2.3. Hydration of Pure Magnesia in Liquid Water

Hydration of pure magnesia in liquid water is a process of dissolution of magnesia particles followed by Mg(OH)2 precipitation [6,7,10,11]. This process is controlled by the dissolution of magnesia particles, beginning with the formation of a surface layer of

MgOH+ around a magnesia particle, via the reaction:

+ MgOsurrace + H -> MgOH surface (2-1)

In acidic solution (pH < 5), the dissolution of the MgOH+ surface occurs through the following reaction:

+ + 2+ MgOH surface + H ^ Mg + H20 (2-2)

At a pH = 5, the dissolution of MgO depends on the diffusion and pH in the solution.

When the pH value is greater than 7, a higher local pH value is produced on the surface of the magnesia particles than that in the bulk solution. When the pH reaches saturation

point (>10.7 at 28°C) [12] during the dissolution of magnesia, the Mg(OH)2 is precipitated:

+ + 2+ MgOH surface + OH" ^ MgOH 'OH"surface Mg + 20H" (2-3)

2+ Mg + 2 OH" -> Mg(OH)2 (2-4)

Various particle shapes of the brucite are obtained in different solutions [13]. The high concentration of OH" caused in the solution (pH > 13) results in a fast nucleation process, generating not-well-defined nuclei, which form globular "cauliflower-like"

9 agglomerations (Fig. 2-5a). In a lower concentration of OH" (pH ~ 10), a platelet-like morphology is attained, resulting from the intergrowth of the hydrate particles (Fig. 2-

5b).

a) b)

Fig. 2-5. Brucite participated at 60°C in (a) fast nucleation process at pH > 13 and (b)

slow nucleation process at pH ~ 10 [13].

2.4. Hydration of Magnesia in Water Vapor

According to literature [14], the hydration of magnesia at room temperature with water vapor proceeds in four steps:

(i) Formation of a layer of chemically and physically adsorbed water.

(ii) Diffusion of Mg2+ and OH" ions in the layer of physically adsorbed water.

(iii) Nucleation of brucite.

(iv) Crystal growth of Mg(OH)2.

10 The process of the hydration of polycrystalline forms of magnesia depends on the raw material from which it was prepared and the preparation conditions. The hydration rate is also influenced by the specific surface and pore volume of the oxide particles [14].

In studying the hydration of magnesias (specific surface 10 - 180 m2g"') with water vapor at 22°C [15], it was found that before the chemical reaction takes place to produce magnesium hydroxide, the physical adsorption of water vapor on magnesia surface occurs. For magnesia powders with high specific surface areas, the hydration rate is higher than for those with low specific surface areas. The magnesia with a specific

2 1 surface area of 10.5 m g" would hydrate slowly and the large original crystallites would split into a number of small single crystallites. At 35°C little or no adsorption of water vapor on the magnesias with low surface areas occurs until the water vapor pressure, p, reaches 30-60% po ipo is saturation vapor pressure). Therefore for a low specific surface area, the required p/po value is high. The work shown in [16] also indicated that a vapor pressure greater than 0.3 po is necessary for hydration, because the water vapor above this pressure condenses as a film on magnesia particles to complete the hydration reaction.

The Van der Waals forces cause the adsorption of water on the particle surface in the initial stage of hydration, followed by the formation of a water film and the slow reaction

2+ between Mg and OH". The thin platelets of Mg(OH)2 grow on the cubic face of periclase grains in different directions .and produce a layer [14].

Hydration studies on caustic magnesias in [17, 18] showed that the capillary condensation of water in the pores of the magnesia particles occurred in the hydration reaction. The process of hydration of caustic magnesias starts with the water

11 condensation in the small pores, followed by the reaction on the pore walls to form hydroxide.

2.5. Kinetics of Hydration

The hydration reaction of magnesium oxide with liquid water or water vapor is a heterogeneous reaction and very dependent on the characteristics and physical properties of magnesia, such as reactivity and surface area. These properties are strongly influenced by the preparation of magnesia, such as the chemical form from which MgO is prepared

(Mg(OH)2, MgC03, etc.), the calcination temperature and holding time, and the preparation atmosphere (vacuum, air, etc.). Both the temperatures and the vapor pressures affect the hydration rates for vapor phase hydration [16].

For uniformly sized spherical particles the classic "shrinking core model" gives the reaction rate equation for processes controlled by chemical reactions [19].

l-(l-a)1/3 =K>t - (2-5)

where a - the fraction of reacted magnesia,

K = rate constant, sec"1,

/ = reaction time, sec.

For a heterogeneous non-catalytic reaction, the reaction rate is commonly either diffusion controlled or chemical reaction controlled [12]. The two processes can be distinguished by the activation energies. For example, the activation energy for the diffusion in solution

12 reaction is of the order of 5 kcal/mole (21 kJ/mole) and the chemical reaction controlled is usually between 10 and 25 kcal/mole (42 to 105 kJ/mole) [20].

The hydration of magnesia powders (specific surface: 12 to 80 m2/g) and particles

(particle size: 45 to 75 [im) under turbulent conditions at different temperatures was investigated [12, 21]. Equation (2-5) could not be applied, due to the non-uniform sizes of the magnesia powders and particles, while equation (2-6) was found suitable. For each particle size fraction of the powders, the individual rate curves can meet equation (2-5), and equation (2-6) would be the weighted average of the rate curves for each individual fraction.

- \n (\ - a) = K*t (2-6) where a = fraction of reacted MgO,

K = apparent rate constant, directly proportional to the weighted average of the constants for the single particle fractions, sec"1.

Under these conditions, the obtained hydration rates are directly proportional to the surface area at the surface of the particles, including pores. The calculated activation energies are 14.1 ± 0.2 kcal /mole (59.0 ± 0.8 kJ/mol) and 45 ± 0.5 kJ/mole (10.7 ± 0.1 kcal/mol). The hydration was controlled by the chemical reaction, because in the stirred solution the layer of brucite is not a stable surface layer and subsequently dissolves allowing further hydration of the magnesia [12, 21].

The hydration reaction of magnesia powders exposed to water vapor pressures of 260 -

570 mm Hg (0.342 - 0.75 atm.) and temperatures of 78 - 98°C is governed by the

13 chemical reaction at the interface [16]. At a constant temperature, the reaction rate varies with the vapor pressure almost linearly with an excess pressure over p*, which is 0.3 times saturation pressure. The activation energy of the reaction E = 16,100 cal/mol. The overall equation can be expressed by:

v = k0[(p/p *) - 1 ] exp(-E/RT) (2-7)

where v = the reaction rate, g'cm'^sec"1,

p = water vapor pressure, mmHg,

p* « 0.3 times saturation pressure, mmHg,

E = activation energy, cal'mol"1,

R = gas constant, 1.986 cal'K'^mol"1,

. T= absolute temperature, K,

2 1 kg = a constant, g*cm~ »sec~ .

The term [(p/p*) - 1] expresses the excess pressure over p* and governs the formation of the water film on the magnesia surface and affects the rate of the hydration [16].

After studying this system, it was found [22] that when the relative humidity is below about 60%, the nucleation of magnesium hydroxide is slow, and this retards the hydration reaction. At a low humidity, the reaction rate is greatly reduced, probably because a layer of water on the surface of magnesia particles forms only partially. At a high relative humidity, the reaction is as follows:

MgO(s) + H20(g) = MgO(H20)adsorbed -*Mg(OH)2 (s) (2-8)

14 The reaction process is governed by the interface reaction. The activation energy is 17.6 kcal/mole. The rate equation is [22]:

{P Ps \-(\-ay =K-t = K<- exp(- SL) • ' \• t (2-9)

RT \-(p/ps) where a = fraction of reacted MgO,

ps = saturation vapor pressure at temperature T, mmHg,

K' = a constant, sec"1,

K = rate constant and K = K'— -jLjLL sec-' !-(/>//>,)

t = reaction time, sec.

Jerofeev's equation (2-10) was used to describe the hydration process of caustic magnesium oxide in aqueous suspensions [23]. This is an empirically established kinetic equation for isothermal heterogeneous processes.

ln [1/(1 -a)] = {Kt)n (2-10)

where a = fraction of reacted MgO,

t - time, sec,

,K = rate constant, sec"1,

n = a constant of the empirical equation.

By defining the half-time to.5 of the given process as the time when the reacted MgO

attains the value of 50%, Eq. (2-10) is changed to the following form:

15 (2-11)

For caustic magnesia prepared by calcining magnesium carbonate in air at 900°C for 60 min, the constant, n, of equation (2-10), has the value of 2/3 and the activation energy is

E = 54.7 kJ/mole [23].

By incorporating a term that represents the influence of the hydration product deposited on magnesia particles, a modified shrinkage core model was developed [9], where the hydration reaction is controlled by the chemical reaction process. The reaction equation for vapor phase hydration is expressed by the following equation:

l/3 logr0[l -(1 -a) ] = (l/A01og(?-^) + (l/A01ogM (2-12) where ro = the mean original radius of magnesia particles, mm,

a = fraction of MgO reacted,

t = the reaction time, hour,

to = the induction time, hour,

k = the ratio of the reaction constant to the particle density, mm/hour,

N = a constant, expressing the contribution of the thickness of reaction layer to the diffusion. (N = 1, the reaction is controlled by chemical reaction, N = 2, the reaction is controlled by diffusion.).

For the vapor phase hydration of single crystals of magnesia at temperatures between 138 and 200°C, Eq. (2-12) fits the reaction well. The activation energy at the beginning of hydration was 65.5 kJ/mole (-15.7 kcal/mole). The values of N are between 1.5 and 1.6,

16 indicating that the reaction is not fully controlled by diffusion or chemical reaction, but by a combination of both [9].

2.6. Hydration of Refractory Grade Magnesia

Generally, magnesia used in refractory is either of dead burned or fused type, which is produced from natural magnesite (MgCO^) or extracted from seawater or inland brines

[24]. Refractory grade magnesia, typically dead burned around 1650°C or higher, is of high density (3.2-3.4 g/cm3) and low porosity (5-11%) and may contain different level of

impurities, such as Al203, SiC>2, Fe203, CaO, B2O3 [24]. These compositions can exist as

crystalline phases, like spinel (MgAl204), forsterite (Mg2Si04), monticellite (CaMgSi04),

etc. A complex spinel (Mg, Fe)(Cr, Al, Fe)204 may also form when MgO, A1203, Fe203,

FeO and Q2O3 exist [25]. The CaO, Si02 and B2O3 may form a glass phase. Due to the existence of these phases, the physical and chemical properties of magnesia-based refractories are more complicated than those of pure magnesias, and the hydration is likely affected by them.

2.6.1. Effect of Physical Properties of Magnesia on Hydration

Since hydration starts from the surface, the specific surface area of magnesia is particularly important for the hydration process. Fine magnesia particles (i.e. with large specific surface area) have high hydration rates. The surface properties of magnesia, such as specific surface area and reactivity, are determined by the temperature of the calcination process and starting materials [15, 26, 27]. The surface area reaches a maximum and then decreases with increasing the calcination temperature. For example,

17 the surface area of magnesia produced by dehydrating Mg(OH)2 changes from about 320

2 2 m /g at 350°C to 70 m /g at 700°C [27]; however, the surface area of magnesia obtained by decomposition of magnesite (MgCCh) increases with the temperature up to 900°C

[28].

The open porosity of magnesia particles is another important property that influences hydration. MgO reacts slowly with water vapor [29]. For porous magnesia, the reaction proceeds between MgO and water condensed in the micro-pores and absorbed on the surfaces of magnesia particles from water vapor. A sintered magnesia clinker has a higher hydration rate than a fused clinker, because typically the sintered magnesia is higher in porosity (8-11%) than the fused magnesia (porosity 3-5%) [24]. Also, the fused magnesia has larger crystals, which give a lower surface area; consequently the resistance to hydration is higher than for sintered magnesia. In other words, well-fused magnesia of high density, low porosity, and large crystal size has high hydration resistance.

2.6.2. Effect of Impurities

2.6.2.1. CaO

From the phase diagram of the CaO-MgO system (Fig. 2-6. [30]), up to 5 wt% of CaO can be soluble in MgO. The CaO content in the MgO solid solution decreases as temperature is reduced. During cooling after the calcination of magnesia, the CaO of the high content region is separated from the high temperature solid solution of CaO-MgO grains and distributed in and around the periclase (MgO) grains. Below the solubility limit, CaO has been shown to preferentially segregate into the grain boundary [31, 32].

18 The hydration activity of calcium oxide is greater than that of magnesium oxide by approximately 8 times [33].

MgO CaO

Fig. 2-6. Phase diagram of system CaO-MgO [30].

2.6.2.2. SiO?

Silica can form compounds with MgO. Magnesia forms forsterite (Mg2Si04) at elevated temperatures, which remains along the grain boundaries of periclase crystals and works as a coating to restrict contact with water, improving the hydration resistance of periclase

[34]. Due to the different crystal structures between forsterite and periclase, the coating effect of silica is inferior to AI2O3, but the silica is easily to form silicate glass phase along periclase grain boundaries and the silicate glass is an effective hydration inhibitor.

19 2.6.2.3. CaO/SiO? Ratio

Calcium oxide and silica can form a secondary phase around magnesia grains. With an

increase in the CaO/Si02 ratio, the secondary phases may be monticellite (CaMgSiO^), and mervinite (Ca3MgSi20s), dicalcium silicate (Ca2SiC>4), or a CaO-rich silicate glass.

Phases with a higher content of CaO, like dicalcium silicate (Ca2Si04), are more sensitive to hydration than periclase, resulting in an increase of the hydration rate of magnesia

[33]. Therefore, magnesia refractories with a high CaO/Si02 are more sensitive to

hydration. A low CaO/Si02 ratio (< 0.5) results in a high hydration resistance, due to the existence of a silicate glass or forsterite (Mg2Si04) along the magnesia grain boundaries.

2.6.2.4. AI2O3

Alumina may form a solid solution in MgO and magnesia-spinel (MgA^O^ with MgO at high temperatures. When alumina is present in the magnesia during burning, spinel forms and is then precipitated along magnesia grain boundaries during cooling. The spinel has the effect of a coating on periclase grains and enhances the hydration resistance of magnesia. This coating effect is most evident when the content of AI2O3 is between 2 and

4 wt% [34]. When the content of A1203 exceeds 4 wt%, the magnesia resistance to hydration is lower, due to the increase in the porosity of the magnesia clinker, caused by the volume expansion of the spinel.

2.6.2.5. Iron Oxides

According to the FeO-MgO phase diagram (Fig. 2-7. [35]), ferrous oxide is completely soluble in magnesia above 1200°C. The difference of ionic radii of Mg2+ (0.65A) and

20 Fe2+ (0.80A) causes the large expansion and weakening of the ionic bond in magnesia when Mg2+ is replaced by Fe2+, therefore, the hydration rate of MgO-FeO solid solution is higher than that of magnesia. However, due to the partial oxidation of Fe2+ to Fe3+, the hydration rate of magnesia with 5 atom% Fe2+ is lower than that of magnesia [36]. The ionic radius of Fe3+ (0.64A) is essentially the same as of Mg2+(0.65A). A 0.5 atom%

3+ 2+ substitution of Fe (in the form of Fe203) for Mg in magnesia produces cation vacancies, which increases the surface energy and promote sinterability. In effect, the density of magnesia increases and the hydration rate decreases [37]. The upper limit of the amount of Fe3+ to improve sintering is 3 atom% [38].

?600 1 1 1 1 1 I I I

ZAOO Liquid / y / y y y y y y y y y y eooo / y y y / y / y y y 1600 Magnesio-wu5tite /

1200 I I I I I I 1 1 L 0 20 40 60 80 100 FeO MgO

Fig. 2-7. Phase diagram of system FeO-MgO [35].

2.6.2.6. B9O3

Small amounts of B2O3 usually improve the hydration resistance of magnesia in the

presence of CaO and Si02 [39], by forming a glass phase along the magnesia grain

21 boundaries. The B2O3 apparently increases the solubility of dicalcium silicate (C2S) in the glass phase, even at concentrations of B2O3 as low as 0.1%. Because it reduces the hot strength and corrosion resistance of MgO refractories, therefore, the amount of B2O3 in magnesia-based refractories must be restricted.

2.6.2.7. Cr?Q3

The effect of Cr203 on the magnesia sintering and hydration behavior is similar to Fe203

[38] [40]. There is also a 0.1 atom % limit for promoting sintering. The excess Cr203 forms MgO-Cr203 spinel that does not hydrate.

2.6.2.8. Other Oxides

With the addition of oxides such as Zr02 and Ti02 to magnesium oxide, the magnesia resistance to hydration is improved, owing to high density and low porosity. and titanium ions, both of 4+ valences, produce vacancies in periclase crystals [37] and the increased surface energy. These oxides promote the sintering of magneisa and produce magnesia of low porosity and large crystal sizes.

22 3. OBJECTIVES

The present work is a part of an ongoing collaborative research project - "Refractories for Non-ferrous Metals Smelting", in which one of the objectives is to enhance the fundamental knowledge in hydration of magnesia-based refractories and to offer practical solutions that would enable non-ferrous metal producers to extend the life in service of such refractories and make a proper selection of bricks to be used for their specific industrial environments.

The objectives of this work are to establish the hydration mechanisms of magnesia-based refractory bricks and to find correlations between the hydration behaviors of various composition bricks and their physical, chemical and micro structural characteristics.

In order to achieve these objectives, the following research work has been conducted:

1. Hydration studies in different hydrating conditions.

Various brands of magnesia-based bricks were studied: magnesite, magnesia-spinel, and magnesia-chrome bricks. The hydration conditions were slightly different for the various bricks, according to their mineralogical compositions:

a) Magnesia bricks, spinel-added and spinel-bonded magnesia bricks: in water at 60,

80, and 100°C, in water vapor at 98% relative humidity at 60 and 80°C, and in

steam at 100, 109, and 121°C.

b) Magnesia-chrome bricks: in steam at 109, 121, and 130°C.

23 2. Investigation of the correlations of the hydration behaviors and the modulus of elasticity by a non-destructive method.

Modulus of elasticity (MOE) was measured by the Impulse Excitation Technique (IET).

The hydration of a magnesia brick alters its internal structure, which triggers changes in

MOE. If a reliable correlation between variation of the MOE and the hydration degree could be established, the degradation caused by hydration in a refractory brick could be evaluated by IET.

3. Investigation of correlation of the hydration and texture-related properties such as pore size, pore size distribution, permeability, mineralogical phases and their distribution. The following parameters are of particular interest:

a) Pore size and pore size distribution, using a Porometer apparatus and kerosene as

an impregnating medium.

b) Permeability, using a VacuPerm Decay Permeameter.

c) Mineralogical compositions, using the X-ray Diffractometry method.

d) Microstructure, using a Scanning Electron Microscope (SEM) with an Energy

Dispersive Spectrometer (EDS).

The results from this work will serve the collaborative research project to develop a selection method for hydration resistant refractories.

24 4. EXPERIMENTAL PROCEDURES AND METHODOLOGY

4.1. Experimental Materials

Magnesia-based refractory bricks include magnesite, magnesia-spinel, and magnesia- chrome bricks. In terms of raw materials, there are two types of magnesia-spinel bricks

(magnesia-spinel and spinel bonded magnesia bricks)* and three types of magnesia- chrome bricks (burned, direct bonded and rebonded fused grain bricks)**. The raw materials for the manufacturing of the first two types of magnesia-chrome bricks are magnesia and chrome ore (complex solid solution of spinels ((Mg,Fe)(Cr,Al)204)), while the starting materials for the third type are fused grains. In the present work, different types of experimental refractory materials, with various chemical compositions and ratios of CaO/Si02 were chosen, in order to study how these factors would affect the hydration behavior.

All three groups of magnesia-based refractory bricks were selected for the experimental work: magnesite with MgO greater than 87%, magnesia-spinel with over 93% MgO

(chemical and physical properties of these refractories are presented in Table 4-1) and magnesia-chrome, "rebonded fused grains", "direct bonded" and "burned" bricks (their properties are presented in Table 4-2).

* Magnesia-spinel refractory brick - a fired refractory made predominantly from a mix of refractory-grade magnesia and refractory-grade spinel.

Spinel bonded magnesia brick - a fired magnesia brick with MgOAl203 spinel as the matrix phase. ** Burned magnesia-chrome brick - a fired brick made from a mix of refractoy-grade burned chrome ore and refractory-grade burned magnesia. Direct bonded basic brick - a fired refractory in which the grains are joined predominantly by a solid state diffusion mechanism. Rebonded fused grain brick - a fired brick made predominantly or entirely from fused grains.

25 Table 4-1. Properties of the experimental magnesia and magnesia- spinel bricks***

Experimental Bricks Ml M2 MSI MS2 Chemical composition, %

Si02 2.40 0.85 0.20 0.60

A1203 0.60 0.15 10.6 6.00

Fe203 1.40 0.30 0.50 0.20 CaO 1.60 2.15 1.40 1.00 MgO 93.0 96.5 87.3 92.2

B203 0.50 - -

CaO:Si02 0.71 2.71 7.50 1.79 Apparent Porosity, % 14.0 14.0 15.0 15.5 Bulk Density, g/cm3 2.92 2.96 3.00 2.96 Modulus of Rupture, MPa 26.2 17.2 5.60 9.00 Cold Crushing Strength, MPa 103 60.0 60.0 _ Modulus of Elasticity, GPa 111 61.4 9.80 16.8 Type B B B,M+S B, M-S *B = burned, M+S = magnesite-spinel brick, M-S = spinel bonded magnesite brick.

Table 4-2. Characteristics of the experimental magnesia-chrome bricks***

Experimental Bricks MCI MC2 MC3 MC4 MC5 MC6 Chemical composition, %

Si02 4.90 0.90 0.60 1.00 2.00 1.60

A1203 21.5 9.80 6.30 6.00 11.3 7.10

Fe203 11.8 16.3 12.0 11.0 8.00 12.0 CaO 0.80 0.60 0.80 1.00 0.70 1.10 MgO 34.6 42.0 62.0 63.0 62.5 60.0

Cr203 26.0 30.4 18.0 17.0 15.5 19.0

CaO:Si02 0.17 0.71 1.43 1.07 0.37 0.74 Apparent Porosity, % 20.6 16.4 13.0 14.5 18.0 18.0 3 Bulk Density, g/cm 3.06 3.33 3.30 3.25 3.06 3.14 Modulus of Rupture, MPa 7.00 7.00 10.3 10.0 7.60 7.00 Cold Crushing Strength, MPa 26.0 24.0 82.7 90.0 49.0 50.0 Modulus of Elasticity, GPa 31.5 24.4 35.3 36.1 12.7 6.20 Type B B RFG RFG DB DB B = "burned", DB = "direct bonded", RFG = "rebonded fused grain".

The data were provided by the brick producers, except for the modulus of elasticity. 4.2. Sample Preparation Procedures

The brick samples were first cross-sectioned by dry saw-cutting into 25 mm thick slices across the length and along their original pressing direction and then further cut into 75 x

25 x 25 mm bars, also along their original pressing direction for the hydration tests and the MOE measurements (Fig. 4-la). For determining pore size and permeability, specimens of 45 mm in diameter and 35 mm in height were core-drilled from bricks along their original pressing direction by using a drill machine that was refrigerated with kerosene (Fig.4-lb). After cutting and drilling, all specimens were dried for 20 hours at

110°C, and stored in desiccators to avoid the contact with moisture. Before the hydration test, they were fired at 800°C for 1 hour, to eliminate the effect of any potential organic hydration inhibitors, and then returned to desiccators and kept there until testing time.

27 45mm in diameter

A

The original pressing direction

b)

Fig. 4-1. Schematic representation of the specimen preparation from a brick, (a) for bars

used in hydration studies and MOE measurement and (b)-for cylinders used to measure

permeability and pore size and pore size distribution.

4.3. Experimental Methods

The specimens as prepared above were subjected to hydration cycles under different conditions. After each cycle, they were dried and visually inspected, and the weight, modulus of elasticity (MOE) by impulse excitation technique (IET), and gas permeability were measured. The hydration cycling continued until disintegration occurred or no readings for MOE measurements could be taken. The mineralogical and microstructural studies by XRD and SEM studies were also performed on the as-received specimens and after hydration test; a representative material was studied by SEM after certain hydration cycles.

28 4.4. Hydration Studies

In this work the weight increase incurred by specimens during the hydration tests was used for assessing the extent of hydration.

4.4.1. Hydration Test Conditions

For magnesite and magnesite-spinel bricks:

• In steam at 100, 109 and 121 °C

• In humid air (98% relative humidity) at 60 and 80°C

• In water at 60, 80 and 100°C

At temperatures higher than 100°C, the samples were exposed to steam between two

and 12-hour cycles. At 60°C and 80°C, the first cycle of hydration was for 6 hours

and increased to 3 days at the end of test because of low hydration rates.

For magnesia-chrome bricks:

• In steam at 109°C (34.5 kPa), 121°C (103.4 kPa) and 130°C (172.4 kPa).

The specimens were exposed to steam for 2-hour cycles at 121 °C and 130°C and

up to 12 hours at 109°C, because of expected low hydration rates in these

conditions.

4.4.2. Hydration Equipment and Procedures

For pressures above 25 psi (172.4 kPa), a Cenco-Menzel autoclave (by Central Scientific

Co.) was used, as recommended in ASTM (the American Society for Testing and

29 Materials) C 456-93 "Standard Test Method for Hydration Resistance of Basic Bricks and Shapes." For the low steam pressures (< 103.4 kPa), a model No. 915 low-pressure vessel (by Wisconsin Aluminum Foundry Co. Inc.) was used.

At temperatures lower than 100°C, a humidity chamber was used and specimens were individually set in beakers. For hydration in water, the beakers with specimens were immersed in water. For the hydration in relative humidity values at different temperatures, the beakers were set on a sample holder in the chamber and kept away from the water. The hydration time was recorded after the testing temperature was reached. A digital hygrometer was used to measure the relative humidity inside the chamber.

The specimens were weighed with ±0.00lg accuracy and placed in glass beakers covered with aluminum foil to protect them against the dripping water from condensation. After treatment in an autoclave, the specimens were dried for 4 hours at 65°C followed by drying at 110°C for 16 hours in a dryer with forced air circulation. They were immediately placed in desiccators to cool down to room temperature and returned to the next hydration cycle after measuring the weight and MOE. The weight increase and MOE were determined on three specimens with the average value taken.

4.4.3. Calculations

The extent of hydration for each sample was calculated as (1) weight increase and (2) fraction of hydrated MgO. The following equations were used:

(1) Weight increase, A W, %:

30 (4-1)

where W„ = the sample weight after hydration for n hours, g,

Wo = the sample weight before hydration, g,

n = the hydration time, hours.

(2) Fraction of reacted MgO, a:

M AW 40.3 AW AW a = M0 = • =2.24

M H20 CMG0 18 CMGQ C-MgO

= where MM8O the molecular mass of MgO,

MH20 ~ the molecular mass of H2O,

CMSO ~ the chemical content of MgO in the brick.

A summary of the hydration tests is presented in Table 4-3.

Table 4-3. A summary of hydration tests

Hydration Total Hours Hours (Cycles) of Brick Testing (Cycles) for Hours Samples Sample Conditions AW,% MOE (Cycles) Keeping Measurement Integrity Ml Steam, 121°C 0.20 ±0.04 24(12) 24(12) > 24 (12) Steam, 109°C 0.23 ± 0.02 48(11) 48 (11) >48 (11) Steam, 100°C 0.21 ±0.03 60(12) 60(12) > 60 (12) Humidity, 80°C 0.37 ±0.02 96(13) 96(13) > 96 (13) Humidity, 60°C. 0.13 ±0.01 192 (9) 192 (9) > 192(9) Water, 100°C 0.16 ±0.06 54(13) - -

31 Table 4-3 cont'd Water, 80°C 0.14 ±0.02 84(12) - Water, 60°C 0.07 ±0.01 192 (9) - - M2 Steam, 121°C 2.04 ±0.15 4(4) 3(3) 3(3) Steam, 109°C 1.49 ±0.18 6(3) 6(3) 6(3) Steam, 100°C 1.57 ± 0.12 16(6) 12(5) 12(5) Humidity, 80°C 2.35 ±0.12 24 (3) 12(2) 12(2) Humidity, 60°C 1.23 ±0.06 144 (8) 144 (8) 144 (8) Water, 100°C 3.34 ±0.06 16(7) - - Water, 80°C 2.89 ±0.10 32(4) - - Water, 60°C 3.23 ±0.05 192 (9) - - MSI Steam, 121°C 1.35 ±0.11 10(7) 10(7) 10(7) Steam, 109°C 1.76 ±0.09 26(8) 20 (7) 20 (7) Steam, 100°C 1.49 ±0.14 36(10) 30(9) 30(9) Humidity, 80°C 0.90 ±0.05 48 (7) 36(5) 48 (7) Humidity, 60°C 0.30 ±0.07 192 (9) 192 (9) > 192(9) Water, 100°C 1.00 ±0.02 32(10) - - Water, 80°C 0.98 ±0.09 48 (7) - - Water, 60°C 0.77 ±0.08 192 (9) - - MS2 Steam, 121°C 1.04 ±0.08 2(2) KD . 1(1) Steam, 109°C 1.13 ±0.11 4(2) 2(1) 2(1) Steam, 100°C 1.45 ±0.11 6(3) 6(3) 6(3) Humidity, 80°C 1.91 ±0.10 24 (3) 12(2) 12(2) Humidity, 60°C 0.38 ±0.04 . 72 (6) 48 (5) 48 (5) Water, 100°C • 1.61 ±0.08 8(4) - Water, 80°C 0.67 ±0.14 12(2) - - Water, 60°C 0.95 ±0.15 96 (7) - MCI Steam, 130°C 0.14 ±0.01 8(4) 8(4) 8(4) Steam, 121°C 0.37 ±0.02 18(9) 10(5) 16(8) Steam, 109°C 0.19 ±0.01 84 (22) 84 (22) > 84 (22) MC2 Steam, 130°C 0.14 ± 0.01 8(4) 6(3) 6(3) Steam, 121°C 0.22 ±0.01 16(8) 14(7) 14(7) Steam, 109°C 0.27 ±0.01 84 (22) 72 (21) >72 (21) MC3 Steam, 130°C 0.59 ±0.03 12(6) 12(6) > 12 (6) Steam, 121°C 1.28 ±0.06 18(9) 14(7) 16(8) Steam, 109°C 1.14 ±0.05 84 (22) 72 (21) 72 (21) MC4 Steam, 130°C - - - - Steam, 121°C 0.72 ±0.04 18(9) 16(8) 18(9) Steam, 109°C 0.50 ±0.02 84 (22) 84 (22) > 84 (22) MC5 Steam, 130°C 0.49 ±0.02 16(8) 16(6) > 16(8) Steam, 121°C 0.52 ±0.03 18(9) 18(9) >18 (9) Steam, 109°C 0.40 ± 0.02 84 (22) 84 (22) > 84 (22) MC6 Steam, 130°C 0.46 ±0.02 8(4) 8(4) 8(4) Steam, 121°C 0.56 ±0.03 8(4) 8(4) >8(4) Steam, 109°C 0.59 ±0.03 84 (22) 84 (22) > 84 (22)

32 4.5. Modulus of Elasticity

The modulus of elasticity (MOE) was determined on as-received samples (not subjected to hydration testing) and after each hydration cycle, using a GrindoSonic MK5

"Industrial" instrument (J. W. Lemmans Inc.), according to the ASTM C1259 -01 the standard test method for "Dynamic Young's Modulus (E), Shear Modulus (G), and

Poisson's Ratio for Advanced Ceramics by Impulse Excitation of Vibration."

The impulse excitation technique (IET) is a non-destructive method for determining the elastic constant. The instrument consists of an impulser, a transducer, an electronic system (signal amplifier and signal analyzer), a frequency read-out device, and supports as shown in Fig. 4-2.

The instrument measures the fundamental resonant frequency of the specimens of suitable geometry, by exciting them mechanically with a light strike of the impulser (a small steel ball attached to an elastic handle). A transducer picks up the vibrations and transforms them into electric signals, processed by an analyzer, which provides a numerical reading that is either the frequency or the period of the specimen's vibration.

The appropriate fundamental resonant frequencies and the dimensions and mass of the specimen are used to calculate the dynamic Young's modulus, the dynamic shear modulus, and Poisson's ratio.

Soft foam supports, which were used to allow for a free vibration of the specimens, were located at the fundamental nodal points (0.224 of the length of the specimen from each end). The transducer was placed in contact with the test specimen to pick up the vibration

33 and the reading was recorded for the calculation of Young's modulus by using provided

software.

Signal amplifier Frequency analyzer

Read out device

Specimen

Impulser Transducer Supports

Fig. 4-2. The impulse excitation technique apparatus.

Young's modulus was expressed the following equation [41] [42]:

2 mf f j} E = 0.9465 • b t (4-3)

where E = Young's modulus, Pa,

m = mass of the bar specimen, g,

34 b, L, t- width, length, and thickness of the bar specimen, respectively, mm,

ff= measured fundamental resonant frequency in flexure, Hz,

Ti = correction factor for fundamental flexural mode to account for finite thickness of the sample bar specimen.

Fro a single crystal, the intrinsic elastic extension corresponds to extending the atom separation uniformly. The stronger the atomic bond the greater the stress required to increase the interatomic spacing and the greater the value of E. The characteristic elastic properties of materials depend primarily on composition, porosity, cracking, and mineralogical phases. The shape and distribution of pores and the interconnections between pores also affect the elastic constants. Many equations have been proposed for relating E to porosity and there is no universal relation. A selection of equations is listed in Table 4-4 [43]

Table 4-4 Relations between porosity P and Young's modulus E [43],

E = E0(l-aP)

2 E = E0(1 -aP + bP )

b E = E0{l-a P)

E = E0\1 +aP/(l-(a+l)P)]

E = E0 exp (-a P)

2 E = E0 exp \-(a P + b P )]

where Eo = the elastic modulus of a fully dense ceramic, and

P = porosity,

a, b = constants.

The variations of the modulus of elasticity for hydrated samples were calculated using the following equation:

35 E„ - E,o AE = (4-4) E, xlOO%

Where EN = MOE after hydration for n hours,

EQ = MOE before hydration,

n = the hydration time, hours.

4.6. Permeability

Permeability measurements were conducted using a VacuPerm Decay Permeameter

(University of Missouri-Rolla, USA) and the permeating fluid was air. Figure 4-3 shows the schematic and instrument of the VacuPerm Decay Permeameter. By vacuuming one side of a sealed sample, the apparatus produces a difference of pressures between the two sides of the sample and when reaching 75% vacuum, it allows air to flow through the sample. The air permeability is determined by measuring the vacuum decay versus time.

The permeability, k, is defined by Darcy's law [44], For sufficiently slow, unidirectional and steady flow:

Q = (lcA/n)(AP/L) (4-5)

Where k = permeability, Darcy,

Q = the volumetric flow rate, cm3 sec- i

A = the cross-section area of the sample, cm2,

L = the length of the sample in the macroscopic flow direction, cm,

AP = ¥\ - ?2, pressure difference between both ends of the sample, atm,

rj = the fluid viscosity, cP.

36 Specimen

Computer

a)

37 A practical unit of permeability is Darcy. A porous material has permeability equal to 1

Darcy if a pressure difference of 1 atm will produce a flow rate of 1 cm3/s of a fluid with

1 cP viscosity through a 1 cm cube, [44]. Thus,

s cP 1 Darcy = ^ ' >^ ) = 0M7^ (4.6) \{cm ) • \{atm I cm)

For low permeability materials, the unit milliDarcy = 0.001 Darcy is used.

After the cylindrical sample was dried and cooled down at the end of each hydration cycle, air permeability was measured and the value was the average of three measurements. After that the sample was subjected to the next hydration cycle.

4.7. Pore Size and Pore Size Distribution

Pore size and pore size distribution were measured with a Capillary Flow Porometer

(Porous Materials Inc.). The apparatus consists of a regulator, valves, barometers, flow meters, and a sample chamber. When compressed air flows through the regulator and valves, which are controlled by a computer, to reach the sample chamber with a sealed sample, pressure differences between both ends of samples and flow rates are recorded, and pore size and pore size distribution can be calculated.

When the pores of a sample are spontaneously filled with a wetting liquid by displacing the gas present in the pores, application of differential pressure of a non-reacting gas on the sample is required to remove the wetting liquid from pores (Fig. 4-4). The differential pressure, AP, required to remove the wetting liquid from a pore of diameter D, is given by [45]:

38 AP = 4y cosd/D (4-7)

Where AP = differential pressure required to remove a wetting liquid from a pore, Pa,

y = surface tension of the wetting liquid, J/m ,

D = diameter of a pore, m,

6 = contact angle of the wetting liquid on the sample material, degree.

The sample filled with a wetting liquid is installed in a sample holder, and the gas pressure is increased on one side of the sample. At a certain pressure (expressed by

Eq.(4-8)), the largest pore is emptied and the gas flow starts. With an increase of pressure, air goes through smaller pores and flow rate increases. From the flow rates through wet and dry samples at the different gas pressures, the pore diameters and pore distribution are calculated using equations (4-7) and (4-8), respectively [46].

f=[-d(fJfd)/dD] (4-8)

where fw and fd are flow rates through wet and dry samples respectively, nrVsec, and dD is the range of pore sizes corresponding to two differential pressures (APs), m.

Kerosene (surface tension = 28 dynes/cm i.e. 0.0028 N/m) was used as the impregnating fluid. Pore size distribution was measured for all of the brick samples before the hydration tests. One selected sample - MC3 was tested only after every four- cycle hydration and the specimen was not returned back to the hydration cycles.

39 Fig. 4-4. The operation principle for measuring pore diameters [47].

4.8. Mineralogy Studies

Mineralogical compositions were analyzed by an X-ray diffractometer (XRD) (Siemens

D5000 X-ray Powder Diffractometer) using CuKa (X = 0.154 nm) radiation. Samples were prepared as powder and the diffraction patterns were taken between 15 and 70 degree two-theta. The analyses on phase compositions were done for all as-received samples and after-hydration specimens. Because there are low content of brucite (less than 2%) in specimens during hydration cycles, brucite is difficult to be detected for the samples during the hydration.

40 4.9. Microstructure Studies

Microstructures were observed with SEM (Hitachi S-3000 Scanning Electron

Microscope) on as received specimens and the selected sample MC3 after each four cycles hydration. To avoid hydration of magnesia during preparing specimens, kerosene was used for lubricant with a Buehler polishing machine. Attempt was made to use polished specimens for after-hydration bricks, but it did not succeed because brucite is very soft and it is difficult to keep it on specimens during polishing, even using low viscosity resin. In order to be able to observe brucite in specimens, some SEM/EDS analyses were performed on fresh fractures of samples.

41 5. RESULTS AND DISCUSSION

5.1. Degree of Hydration under Different Testing Conditions

5.1.1. Magnesite and Magnesia-Spinel Bricks

For these two types of bricks, hydration tests were conducted under the following

conditions: in steam at 121, 109 and 100°C, in 98% relative humidity at 80 and 60°C, and

in water at 100, 80 and 60°C. The weight increased for each sample with the hydration

time. Figures 5-1-1 to 5-1-8 show the results of the hydration tests. Under the same

conditions, different brick samples hydrated differently. The MSI brick had a moderate

hydration rate, lower than MS2 and M2, but greater than Ml. For the hydration tests

conducted above 100°C, the curves of weight increase for MSI showed three stages

(Figures 5-1-1 to 5-1-3, 5-1-6): i) a fast increase in weight followed by ii) a slow increase and finally iii) another fast increase, until the sample disintegrated. The last stage was not clear for the hydration of MSI in water or humidity at 80 and 60°C (Figures 5-1-4, 5-1-5 and 5-1-7, 5-1-8).

Figures 5-1-1 to 5-1-8 show that brick MS2 is the most vulnerable to hydration. It took the shortest time for its samples to fail, even though the weight increase of the MS2 sample was lower than MSI and M2 in most cases. The hydration was accelerated after the first few cycles.

The Ml specimens showed the highest hydration resistance: the weight increase for all hydration conditions was very low (less than 0.3%), no disintegration occurred in any tests and no cracking was observed in specimens.

42 The M2 samples were less affected by hydration than the MS2 samples. The weight increase, in some cases, was higher than that of MS2, but the time for the M2 samples to fail was longer than that for MS2. The plots of weight increase versus time for M2 showed three stages, similar to the MSI specimens.

For all experimental bricks, the weight increase was higher at higher temperatures. If disintegration time is the total hydration time until a specimen disintegrates, then the disintegration time was longer at low temperatures than at high temperatures. This indicated that high temperature accelerates the hydration of magnesia bricks.

The MSI brick, containing 10.6% AI2O3, is a burned magnesia-spinel brick. Spinel was added as aggregate and fine powders, so MSI sample contains large aggregates of magnesia and spinel and the matrix also contains both spinel and periclase. A bonding spinel phase around fine magnesia particles in the matrix was found (Fig. 5-1-9). CaO exists in the binding phase by EDS (Fig. 5-1-10). The XRD diffractogram shows that the

mineralogical phases in MSI are periclase and spinel (MgO»Al203) (Fig. 5-1-11). The

content of MgO was only about 87.3%, while the 10.6% Al203 combines about 4.2%

MgO in MgO»Al203 spinel. This was the lowest content of MgO in the four experimental bricks. One of the reasons why the MSI sample had a relative low hydration rate was probably that the spinel binding phase around the periclase grains worked as a protecting layer against hydration.

43 0 4 8 12 16 20 24 Hydration Time, hrs

Fig. 5-1-1. Weight increase versus time for hydration in steam at 121°C.

Hydration Time, hrs

Fig. 5-1-2. Weight increase versus time for hydration in steam at 109°C. 2 -i

Hydration Time, hrs

Fig. 5-1-3. Weight increase versus time for hydration in steam at 100°C.

2 -i -•-MSI 1.6 - -•-MS2 -A-Ml 1.2 - ^-M2 'eas e

0.8 - igh t I i

0.4 -

, t- -*—-»- -*—* * -*—*— V ™ i 0 1 i 1 1 1 i i 0 12 24 36 48 60 72 84 Hydration Time, hrs

Fig. 5-1-4. Weight increase versus time for hydration in 98% relative humidity at 80°C.

45 Fig. 5-1-5. Weight increase versus time for hydration in 98% relative humidity at 60°C.

+-MS1 *-MS2

0 10 20 30 40 50 60 Hydration Time, hrs

Fig. 5-1-6. Weight increase versus time for hydration in water at 100°C.

46 Fig. 5-1-7. Weight increase versus time for hydration in water at 80°C.

MSI *-MS2

0 40 80 120 160 20(J Hydration Time, hrs

Fig. 5-1-8. Weight increase versus time for hydration in water at 60°C. Fig. 5-1-9. Backscatter electron micrograph of the as-received MSI sample.

48 Al

O Ca

Fe Fe —i— keV

Oxides MgO AI2O3 Si02 CaO Fe203 wt% 18.45 59.12 2.70 18.27 1.45

Fig. 5-1-10. EDS results of the bonding phase in the MSI matrix.

The MS2 brick is a spinel bonded magnesite brick, in which the aggregates of coarse and intermediate size are magnesia and the matrix contains a spinel phase as a bond (Fig. 5-1-

12). The XRD diffractogram (Fig. 5-1-11) shows that there are two phases in MS2:

periclase and MgO»Al203 spinel. Comparing the XRD pattern, the intensity of periclase phase in MS2 is higher than in MSI, i.e. the content of periclase phase is higher in MS2

49 than in MSI. The spinel phase assembles together and does not form a protecting film on periclase (Fig. 5-1-12). Comparing Figures 5-1-12 and 5-1-9, it can be seen that the magnesia particles in MSI and MS2 are different. The magnesia particles in the'MSI sample have irregular shapes, and there are more pores in the MS2 sample than in the

MSI sample. The periclase phase in the MS2 sample contains more grain boundaries and their sizes are smaller than in sample MSI. These kinds of periclase grains can more easily produce new surfaces for hydration and tend to disintegrate faster. This was confirmed by the weight increase of the MS2 sample, which was lower at the beginning of hydration, but later increased faster and had a shorter disintegration time than the MSI sample.

MSI

S S^P } v S

B 4> MS2 . P S s,p S 1 s

15 20 25 30 35 40 45 50 55 60 65 70

2 Theta, Deg.

Fig. 5-1-11. XRD diffractograms for the unused MSI and MS2 samples (P = periclase

and S = spinel)

50 Fig. 5-1-12. Backscatter electron micrograph of the MS2 sample.

Table 5-1-1. The compositions of binding phases for sample MS2, Ml, and M2 from EDS analysis.

wt% MgO A1203 Si02 CaO Fe203

MS2 28.12 71.54 0.34 - _ Ml 27.03 0.20 38.12 34.36 0.29 M2 98.32 0.21 0.42 0.79 0.25

The Ml sample is a burned magnesite brick with a relatively compact texture. The

binding phase around the magnesia particles has the ratio of CaO : MgO : Si02 of 1:1:1

51 (Table 5-1-1), which is equivalent to that of monticellite (CaMgSi04). The fact that XRD did not detect any silicate compounds in the Ml sample indicates that the binding phase is a glass phase (Fig. 5-1-13 to 5-1-14). The silicate glass around periclase and the lowest ratio of CaO/SiCh (= 0.7) of the sample would be the reasons why the hydration rate was the lowest out of the four chrome-free bricks. The graphs of the weight increase of the

Ml samples showed that there was a fast increase in weight during the first few cycles of hydration, but the second stage of the hydration seemed not complete.

Fig. 5-1-13. Backscatter electron micrograph of brick Ml: binding phase (light gray) and

periclase (dark gray).

52 p

Ml P

C

M2 1 15 20 25 30 35 40 45 50 55 60 65 70 2 Theta, Deg.

Fig..5-1-14. The XRD diffractograms for the unused Ml and M2 samples (P = periclase).

Fig. 5-1-15. Backscatter electron micrograph of the M2 brick structure.

53 Sample M2 is also a burned magnesite brick containing the highest content of MgO and

has the most compact texture of all four bricks (Fig. 5-1-15). There are some cracks on

the magnesia particles, but there was no second phase around magnesia particles (Table

5-1-1). The EDS results showed that there were very small amounts of impurities. From

Figures 5-1-1 to 5-1-8, the hydration rate for the M2 samples is higher than for MSI and

Ml in terms of the weight increase, and the disintegration time is shorter than for these

two samples. Microcracks in the magnesia particles are assumed to have a greater

contribution to the high hydration rates. The microcracks probably formed during the

cooling process of the brick and that increased surfaces to react with water.

5.1.2. Magnesia-Chrome Bricks

For magnesia-chrome bricks, hydration tests were conducted under the following conditions: in steam at 130°C, 121°C and 109°C. The hydration results are shown in

Figures 5-1-16 to 5-1-18. The increase of hydration rates with the temperature indicates that the rising temperature accelerates the hydration process. Similar to the hydration of magnesia-spinel and magnesia bricks, the samples were disintegrated in a shorter time for hydration at a higher temperature. At 109°C, all six magnesia-chrome samples maintained their shapes after 84 hours of treatment, but at 121°C the hydration time before disintegration was less than 20 hours. At 130°C, the shortest disintegration time was 8 hours, i. e. MCI and MC2 (Fig. 5-1-18). At 121°C, the three stages of the hydration processes were clearly shown in Fig. 5-1-17 for all six MC bricks, but at 109°C the last stage did not appear for most samples because of lower hydration rates.

54 A MC2

0 12 24 36 48 60 72 84 Hydration Time, hrs

a)

b)

Fig. 5-1-16. Weight increases versus time for hydration in steam at 109°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6.

55 0.8 -,

0 4 8 12 16 20 Hydration Time, hrs

b)

Fig. 5-1-17. Weight increases versus time for hydration in steam at 121°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6.

56 • MCI -*-MC2 -*-MC3 0.8 -•-MC5 ->—MC6 i 0.6 ^ u B

0.4 DD

I 0.2

8 12 16 20 Hydration Time, hrs

Fig. 5-1-18. Weight increases versus time for hydration in steam at 130°C.

In terms of weight increase, the MCI sample had the lowest hydration rate of the six magnesia-chrome bricks under all experimental conditions. Brick MCI is of "burned" type brick. The XRD analysis indicates the main phases in MCI are complex spinel (Mg,

Fe)(Cr, Al, Fe)204 and periclase (Fig. 5-1-19). Generally, silicate-bonded magnesia-

chrome bricks contain -3-10 wt% Si02, whereas "direct bonded" bricks contain < 3

wt% Si02 [48]. Brick MCI, containing 4.9% Si02, is a silicate-bonded magnesia-chrome brick with a silicate glass binding phase, which are not detected by XRD analysis. Brick

MCI is of the lowest content of MgO (34.6%) and the lowest ratio of CaO/Si02 (= 0.2) of all six magnesia-chrome bricks. The characteristics of the structures and compositions could be the explanation of the lowest hydration rate for the MCI sample.

57 c c

pi

15 20 25 30 35 40 45 50 55 60 65 70

Fig. 5-1-19. XRD diffractogram for MCI- the unused sample. P = periclase (MgO) and

C = complex spinel (Mg, Fe)(Cr, Al, Fe)204.

Table 5-1-2. EDS results of the binding phasee in the MCI and MC2 samples, wt%.

Oxide MgO A1203 Si02 CaO ; Cr203 Fe203 MCI 40.86 8.44 13.66 2.57 9.60 24.86 MC2 39.40 11.29 - - 13.15 : 36.15

MC2 is of the same type as MCI and of similar mineralogical composition, i.e. complex

spinel (Mg, Fe) (Cr, Al, Fe)20.4 and periclase (MgO), but the grain boundary does not indicate the presence of silicate-containing phases (Table 5-1.-2). MC2, with the higher

MgO (42.0%) and CaO/Si02 ratio (0.7) than MCI, showed higher hydration rates than sample MCI.

58 Both the MC3 and MC4 samples are of "rebonded fused grain" type and have similar properties. Both bricks had higher hydration rates than MCI and MC2. These two bricks also showed that they had higher contents of MgO (62.0% and 63.0%> respectively) and

higher CaO/Si02 ratio (1.4 and 1.1 respectively) than those of MCI and MC2.

Comparing MC3 and MC4, they have almost the same chemical compositions and

physical properties, except for a higher CaO/Si02 ratio for MC3. The higher hydration

rates are assumed to relate to the higher CaO/Si02 ratio. Brick MC5 and MC6 are of

"direct bonded" type, and their MgO contents, CaO/Si02 ratios and hydration rates are between the groups of "burned" and "rebonded fused grain". The MC6 brick showed

higher hydration rates and higher CaO/Si02 ratio than brick MC5.

From the discussion above, the CaO/Si02 ratio plays an important role in the hydration of

magnesia-based bricks. The phases containing CaO and Si02 with a low CaO/Si02 ratio

(less than 1) is not easy to react with water and can protect the magnesia from water if

they surround the pericalse grains. With the increase in the CaO/Si02 ratio, especially

when greater than 2, the phases containing CaO and Si02 can be 2CaO- Si02 (C2S), and other compounds with high content of CaO. These CaO-rich phases are very sensitive to hydration. Wherever they are distributed in bricks, they will increase the hydration rate of

magnesia-based bricks. For the experimental bricks, the relationship between CaO/Si02 ratio and hydrated MgO is shown in Figures 5-1-20 and 5-1-21. For all experimental

bricks, the amount of hydrated MgO linearly increases with the CaO/Si02 ratio and with the hydration time, the effect is more remarkable.

59 Fig. 5-1-20. Hydrated MgO versus CaO/Si02 ratio for the hydration of the magnesia-

spinel and magnesia bricks in water at 60°C.

0.6 0.9 Ca0/Si02 Ratio

Fig. 5-1-21. Hydrated MgO versus CaO/Si02 ratio for the hydration of the magnesia-

chrome bricks in steam at 109°C.

60 5.2. Modulus of Elasticity (MOE) versus Hydration Degree

The MOE determination was conducted with the impulse excitation technique (IET).

Figures 5-2-1 to 5-2-8 show the MOE variation versus the hydration time for all the experimental samples, under different hydration conditions.

For MSI, during the first cycle of hydration (the first hour) in steam at 121°C, both MOE and the weight gain increased rapidly, followed by slower variations (Fig. 5-2-1 and Fig.

5-1-1). After the MOE reached a maximum, the weight continued to increase (Fig. 5-1-

1), but the MOE decreased to the initial value. MOE values below the initial value corresponded to the final rapid increase in weight.

-20 J Hydration Time, hrs

Fig. 5-2-1. MOE variation versus time for hydration in steam at 121°C for magnesia-

spinel and magnesia bricks (refer to Table 4-1 for the initial value of MOE).

61 In steam at 109 and 100°C and in humidity at 80°C, the changes in MOE have the same profiles as those at 121°C. During the first few cycles, the MOE quickly increased to the maximum values, and after that, the MOE slowly decreased until reduced below the initial values, and finally the samples disintegrated (Figs. 5-2-2 and 5-2-3). In humidity at

60°C, the MOE remained at the maximum level after hydration for 200 hours, and this could be related to the slow hydration.

For sample MS2, the MOE variation was different from that for MSI. The MOE decreased just after one cycle of hydration in steam at 121 and 109°C. Actually, the MOE maximum should have occurred during the first cycle, but could not be observed, owing to the high hydration rate. The maximum MOE was shown during the hydration in steam at a lower temperature (100°C) and in humidity at 80 and 60°C (Figures 5-2-3 to5-2-5).

Hydration Time, hrs

Fig. 5-2-2. MOE variation versus time for hydration in steam at 109°C for magnesia-

spinel and magnesia bricks (refer to Table 4-1 for the initial value of MOE).

62 The Ml brick had the lowest hydration rate among the chrome-free bricks. For all experimental conditions, the weight increase was slow and the MOE continued to increase slowly. The maximum value of MOE was not reached even after hydration for

200 hours.

The patterns of the MOE variation for the M2.brick were similar tb those for MSI. The

MOE reached its maximum after one cycle of hydration in steam at 121°C, and during the second cycle of hydration the MOE reduced to the initial value. Further hydration caused the disintegration of the samples. At low temperatures, the time for the MOE to reach the maximum and for the sample to disintegrate became longer than at high temperatures (Figs. 5-2-1 to 5-2-5).

Hydration Time, hrs

Fig. 5-2-3. MOE variation versus time for hydration in steam at 100°C for magnesia-

spinel and magnesia bricks (refer to Table 4-1 for the initial value of MOE).

63 -10 J Hydration Time, hrs

Fig. 5-2-4. MOE variation versus time for hydration in 98% relative humidity at 80°C for magnesia-spinel and magnesia bricks (refer to Table 4-1 for the initial value of MOE).

25 n

Hydration Time, hrs

Fig. 5-2-5. MOE variation versus time for hydration in 98% relative humidity at 60°C for magnesia-spinel and magnesia bricks (refer to Table 4-1 for the initial value of MOE).

64 0 14 28 42 56 70 84 Hydration Time, hrs

b)

Fig. 5-2-6. MOE variation versus time for hydration in steam at 109°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 (refer to Table 4-2 for the initial value of MOE).

65 a)

b)

Fig. 5-2-7. MOE variation versus time for hydration in steam at 121°C. (a) for brick MC2, MC3 and MC5 and (b) for MCI, MC4 and MC6 (refer to Table 4-2 for the initial value of MOE).

66 -20 J Hydration Time, hrs

Fig. 5-2-8. MOE variation versus time for hydration in steam at 130°C for the magnesia-

chrome bricks MCI, MC2, MC3, MC5 and MC6 (refer to Table 4-1 for the initial value

of MOE).

For all the magnesia-chrome brick samples, the profiles of the MOE variation were

similar. The MOE values first quickly reached their maximum, followed by slow

decreases below the initial levels, until the samples disintegrated (Figs.5-2-6 to 5-2-8).

The change in the MOE with hydration time for all brick samples obviously had three stages: i) an increase from the initial value to reach a maximum value, ii) a decrease from the maximum to the initial value, and iii) below the initial value until sample disintegration. This was a typical pattern for the MOE variation. The stages for the MOE were not exactly the same as the stages for the weight increase. Obviously, the weight

67 increase has a direct influence on the MOE variation, but the mechanisms affecting the

MOE value will be discussed in Chapter 5.3.

Under different temperatures, the changes in MOE versus hydration time were of the same profile. With increasing temperature, both the MOE maximum values and the time to reach the maximum tended to decrease. For example, for MSI sample, in steam at 100,

109 and 121°C, the maximum MOE variations were 37.6%, 34.3% and 33.4%, and the time to reach the maximum was 10, 6 and 4 hours, respectively (Figs. 5-2-1 to 5-2-5).

5. 3. Influence of Brucite on MOE Variation

In early stages of hydration, the modulus of elasticity generally increases with the degree of hydration for the magnesia-based refractory materials, owing to the development of brucite. The presence of brucite may affects MOE in the following ways: the produced brucite grows on the walls of the open pores and forms a film, which reinforces the brick and increase the MOE; due to the larger volume of brucite the formation of microcracks along the grain boundaries weakens them and have a negative effect on the MOE increase; the brucite fills in the small pores of the hydrated material, and reducing the porosity increase MOE.

The hydration of MgO-based bricks produces brucite, which grows on the walls of the open pores and on the surfaces and gradually forms a film. This film serves to reinforce the brick structure and it will maintain its elasticity as long as it remains continuous.

The hydration process of brick MC3 in steam at 109°C was chosen as an example to present this mechanism. The weight increase and MOE variation are shown in Fig. 5-3-1.

68 The samples hydrated for 0, 8, 24, and 38 hours were chosen to observe the

microstructures.

Hydration Time, hrs

Fig. 5-3-1. Weight increase and MOE variation versus the time for hydration in steam at 109°C for brick MC3.

When the SEM specimens were polished, the soft brucite was easily washed away. In order to be able to verify that the film of brucite formed inside the open pores, SEM/EDS analyses was performed on fresh fractures of sample. The SEM image of the fractured non-hydrated MC3 sample is shown in Fig. 5-3-2. The compositions of the grains in sample MC3 are shown in Table 5-3-1.

69 Fig. 5-3-2. SEM image for the fracture section of the non-hydrated MC3 sample.

Table 5-3-1. The compositions of different grains in MC3 shown in Fig. 5-3-2, wt%

Oxide Point A Point B Point C Point D MgO 89.1 14.9 58.0 22.7 _ A1203 - 13.9 7.63

Si02 0.84 - 1.00 37.2 CaO - - 0.35 38.4

Cr203 3.90 50.3 20.9 0.82 Fe 0 2 3 6.14 19.6 12.1 0.90

Ti02 - 1.36 - -

Fig.5-3-2 showed that the MC3 sample consisted of fused magnesia-chromite spinel grains, chrome ore grains, isolated impurities, and pores. The bond between the fused grains appeared strong, as the sample fractured intragranularly. Brucite deposited on the nore walls

WD15.2mm 20.OkV"x7.Ok " *5um

Fig. 5-3-3. The film of brucite on the wall of a pore after hydration for 8 hours.

As seen in Fig.5-3-1, after hydration for 8 hours at 109°C, the MOE value reached the maximum value. The SEM image of the MC3 sample displayed a film of brucite on the walls of pores (Fig. 5-3-3). The film consisted of brucite crystals of about 1 u.m, with a platelet-like shape. This crystal morphology indicated that the nucleation of the brucite was slow and the crystals developed very well [13]. This film can strengthen the structure of the brick and enhance modulus of elasticity and the strength. This result is in accordance with another work [8] on magnesia-chrome bricks, in which the authors found that the hydration of a small amount of magnesia in magnesia-chrome bricks could increase brick strength (modulus of rapture). In addition, the formation of brucite, during which the volume increases, induces a compressive stress on the neighborhood of a pore.

71 The compressive stress could affect the bonding of periclase grains, but for a low degree of hydration, when the stress was much less than the bonding strength, the modulus of elasticity would increase with the thickness of brucite film.

At the beginning of hydration, a large weight increase and a high increase in MOE takes place. A film of brucite forms on pore walls, which reduces the ulterior hydration rate.

During the early stage of hydration, an increasing thickness of the film is accompanied by the rising MOE. The maximum values of MOE would correspond to a maximum thickness of the film, called "critical thickness". When the brucite film exceeds the

"critical thickness", the stress accompanying its shrinkage during drying between hydration cycles, will exceed the maximum strength and cracking occurs, and then the

MOE begins to decrease. An empirical model [49] of the hydration process was used to calculate the thickness, h [um], of the brucite film. For a hydrated MgO brick of perfectly cylindrical open pores of median diameter do [p.rn], the thickness of the film of brucite, h was calculated with the following equation:

h=d0[l -(1 -AWk/Pf ] (5-1) with

k = Pbrick MMgo/pbrucite MH20 (5"2) where AW = the weight increase, %,

P = apparent porosity of the brick, %,

MMgo = molar mass of MgO, g/mol,

MH2o - molar mass of H2O, g/mol,

72 3 Pbrick = density for brick, g/cm ,

3 Pbrucite = density for brucite, g/cm ,

do = median pore diameter, jum.

For this hydration condition, Fig. 5-3-1 shows that when MC3 sample reaches the

maximum MOE, its weight increase, AW is 0.33% and the porosity, P is 13.0% (Table 5-

3-2), and the pore size, do is 15 -20 um (Fig. 5-3-3). From the equations (5-1) and (5-2),

the maximum thickness of the film of brucite is calculated at about 0.6-0.8 um, which

was in agreement with the SEM observation (Fig. 5-3-3).

After the reaction takes place at grain boundaries with the hydration process, the brucite

at periclase boundaries tends to separate periclase grains due to the larger volume of brucite than mangesia and gradually forms microcracks along the grain boundaries,

weakening the grain boundaries. The weakening of the bonding of periclase grains

further reduces the MOE. After hydration for 24 hours, the modulus of elasticity was only

about 70% of the maximum value. The fractograph of the MC3 sample is shown in Fig.

5-3-4. The fact that the weak grain boundaries make the sample most transgranularly

fracture and microcracks take place around magnesia particles indicates that the hydration proceeds along grain boundaries and the stress induced by the higher volume of brucite weakens the grain boundaries.

73 Fig. 5-3-4. The fractograph of the MC3 sample after hydration for 24 hours.

During further hydration to 38 hours, the MOE decreased to almost the initial value and the grain boundaries of magnesia particles were weaker. "Dusting" was observed on magnesia particles. Fig. 5-3-5a showed the fracture surface of the MC3 sample after hydration for 38 hours. In order to more clearly observe the "dusting", the SEM image of a polished sample is shown in Fig. 5-3-5b. After this point, "dusting" would occur through the whole brick, MOE continued to decrease with hydration, and the sample would lose its physical integrity. It is not clear why "dusting" takes place when the MOE reaches around its initial values yet.

74 Fig. 5-3-5. SEM images of the sample MC3 after hydration for 38 hours in steam at

109°C, (a) the fracture surface and (b) the polished sample.

75 The apparent porosity was tested at the different hydration time points mentioned above.

The results were presented in Table 5-3-2. The results indicated that when the apparent porosity decreased to the lowest value of 13.0%, the MOE reached the maximum value.

Table. 5-3-2. Apparent porosity of MC3 after hydration in steam at 109°C

Hydration Time, hrs 0 8 24 38 Apparent Porosity, % 13.6 13.0 13.9 14.9 MOE, GPa 35.3 40.6 38.0 35.5 MOE Variation, % 0.00 15.1 7.70 0.48

Although the film of brucite, cracking, and porosity affect the MOE, it is difficult to distinguish the effect of each of them on the MOE separately. The MOE variations are caused by their convolution results. The single mechanism of the influences will be studied in the future.

5.4. Influence of Hydration on Pore Size Distribution

The pore size distributions (PSD) versus hydration time presented in Figure 5-4-1 are curves of bi-modal type (at 0.42|im and at 2.5u.m for as-received sample), which maintain their shape during the hydration process, with changes only in the median pore diameter. Brick MC3, used here as an example, had 10% of 0.3um diameter pores before hydration, with the smaller pores mainly distributed at 0.42um and larger pores at 2.5um.

After 8 hours of hydration in steam at 109°C, the 0.3 urn pores disappeared and the number of pores of 0.35um, 0.4um, and 0.5um increased. After 16 hours of hydration there were no pores smaller than 0.35um and the number of those of 0.4um, 0.42um, and

2.5um slightly increased. The disappearance of the small pores and the relative decrease

76 of the large pores were considered the result of brucite deposition in the open pores. The

increase in MOE at the beginning of hydration was definitely related to the disappearance

of these small pores. Going back to the MOE variation with the hydration time (Fig. 5-2-

6a), brick MC3 started to show a decrease in MOE after 8 hours and reached the initial

value of the unused brick after 38 hours. A continuous increase in the number of pores of

0.4-0.5 urn after 24 hours and a jump to larger pores (0.7|j.m) after 38 hours were

observed in Figure 5-4-la. This behavior could be attributed to the occurrence of the

massive cracking at the grain boundaries. For the coarse pores (larger than lum), there was a continuous increase in the number of pores of ~ 2.5 (im, accompanied by the

appearance and the increase of the number of pores larger than 10 ju.m (Figure 5-4-1),

which did not exist in the brick before hydration.

77 8 Hydration Time, hrs _ -o- _ o 8

A- 16 24

2.5 5 7.5 10 25 100 Pore Size, um

b)

Fig. 5-4-1. Variation of PSD for different hydration periods in steam at 109°C for brick MC3. (a) smaller than lum pores and (b) larger than lum pores.

5.5. Influence of Hydration on Permeability

Permeability was measured after each hydration cycle and the changes during hydration are presented in Fig.5-5-1 for magnesia-spinel and magnesite bricks.

It can be seen that for all the bricks, the permeability decreased at first to a minimum value, and then showed a rapid increase. Comparing Fig. 5-5-1 and the changes in MOE in Fig. 5-2-6, permeability reached a minimum when the MOE was at its maximum. The increase in MOE corresponded to the decrease in the permeability, and vice versa. The decrease in permeability was related to filling with brucite inside the open pores,

78 resulting in reduction in porosity. After the maximum thickness of the brucite film is

reached and cracking.occurs along the grain boundaries, permeability gradually increases

until "dusting" occurs. Then macrocracking through the grain boundaries appears,

causing the sharp increase in permeability. The change in the permeability for magnesite- chrome bricks at 109°C was presented in Fig. 5-5-2.

For magnesia-chrome bricks, the changes in permeability were also well correlated with the opposite changes in MOE (Fig. 5-5-2 and Fig. 5-2-6). These results confirm the mechanism described above.

Hydration Time, hrs

Fig. 5-5-1. Permeability versus hydration time for brick MSI, MS2, Ml, and M2 in 98% relative humidity at 80°C.

79 1000 • MCI * MC4 - MC6 800 -+-MC3

600

i*TT,JTT1

400

a> CH 200

0

12 24 36 48 60 72 84 Hydration Time, hrs

Fig. 5-5-2. Permeability versus hydration time for brick MCI, MC3, MC4 and MC6 in steam at 109°C.

5.6. A Simplified Kinetic Analysis

During the kinetic analysis, the results are considered on the basis that the reaction takes place in the open pores, pores are prefect cylinders and distributed uniformly, and the reaction proceeds by propagation of a reaction front. The hydration model, which I have named "cylindrical pore model", is shown in Fig. 5-6-1, and the original radius and length are ro [cm] and Lo [cm], respectively. Other models have been considered, but so far have yielded results in less satisfactory agreement with the experimental data. If the reaction is limited by. the rate of chemical reaction at an interface, the basic kinetic equation could be of the following form:

80 dmldt = u-S (5-4) where m = reacted MgO in mass around a pore at time t [hr], g,

u = reaction rate by weight per unit area, g/hr/cm"2,

S = reaction area, cm2,

For a cylindrical pore of radius r at time t and the density of magnesia particle pMgo, Eq.

(5-4) becomes

2xpMgOL0r(dr I dt) = InL^ru (5-5)

Integration between the original ro and r, t from zero and t gives

r-r0 =(u/pMg0)t (5-6)

81 If a is the fraction of hydrated MgO, and w [g] is the average MgO mass around each pore, then,

2 2 a = n(r - r )L0pMgO I w (5-7)

According to the definition of apparent porosity, we obtain the following equations

P = NV0/Vsamp/eor

N = P- Vsample/V0 (5-8) where P = apparent porosity of a sample, %,

N = number of open pores in the sample,

3 Vmmp\e= volume of the sample, cm ,

Vo = nro Lo, the original volume of each open pore, cm .

Because only the open pores affect hydration, the average MgO mass around each pore w is obtained from Equation (5-8):

W = Vsample • Pbrick ' ^ = ™tUP brick ' P (5"9)

Substituting Eq.(5-9) into Eq.(5-7) gives

2 a = {(rlr0) -\)(pMgO-Plpbrick)

If let k,= Phrick ,then

PMSQP

82 r/r (1+kia) 1/2 0 = (5-10)

Combining Eq.(5-10) and Eq.(5-6) gives

,/2 R,(a) = (l+k,a) -1 = (ulp r0)*t = KH (5-11)

where K = u/pMOr0 is the rate constant, which is dependent on the hydration

temperature. Eq.(5-ll) is the rate equation controlled by the chemical reaction for the hydration with the "cylindrical pore model".

If the diffusion controls the hydration, under isothermal conditions, the growth rate of the interface layer is given with a good approximation by the parabolic relationship for the initial stage of the reaction, for which the equation can be written as follows [50]:

/l d = (DTt)' (5-12)

2 where 5 [cm] is the thickness of the reaction product layer (brucite), and DT [cm /sec] is the diffusion coefficient at temperature T, and t [hr] is time.

From this hydration model (Fig.5-6-1), the thickness of the reaction product, S, can be expressed by

l/2 /2 S = r0- [(l+k,a) -((l+kia)- (l-k2) + k2)' J (5-13) • with

_ P brick and k2 =

PMSQP

83 where a = fraction of hydrated MgO,

r0 = the original radius of pores, cm,

3 Pbrick - the density of the brick, g/cm ,

= 3 PMgo the density of magnesia, g/cm ,

3 pbrucite = the density of brucite, g/cm ,

P - the apparent porosity of the brick, %,

MMgo = the molar mass of MgO, g/mol,

MMg(OH)2 - the molar mass of Mg(OH)2, g/mol.

Substituting Eq.(5-13) into Eq(5-12) gives

1/2 l/2 2 R2(a) = [(l+k,a) - ((l+k,a)- (l-k2) + k2) J = (DT/r0) t = K- t (5-14) where K - Dj/ro is the rate constant, which is also dependent on the temperature. Eq.(5-

14) is the rate equation for the hydration controlled by the diffusion through the product layer within the "cylindrical pore model".

The activation energy and rate of a reaction are related by Arrhenius equation:

K = A*exp(-E/RT) (5-15) where K = the rate constant, 1/hr,

A = frequency factor,

E = the activation energy, J/mol,

R = gas constant, 8.314 J/K»mol,

7= absolute temperature, K.

84 Activation energies are determined by plotting the logarithm of K against l/T on a graph,

and determining the slope of the straight line that best fits the points.

In the kinetic analysis, brick Ml was chosen as an example, because it had the high

content of MgO and low CaO/Si02 ratio, meaning that calcium oxide has little influence

on the hydration of magnesia, and assume that the weight increase is obtained only by the

magnesia hydration.

The discussion in Chapter 5.1 indicated that there are three stages of hydration in terms of the weight increase. The first stage of relatively fast hydration is followed by the second stage with the slower hydration rate. The third stage is a fast hydration. The hydration duration for each stage depends on the hydration rate of a specific brick. The three stages are controlled by different hydration mechanisms. Studying the first two stages is more important to understand the hydration process, as the third stage corresponds to the

"dusting" process. For Ml there was no third stage because of a low hydration rate.

Therefore, the first two stages will be analyzed in this section.

In the mathematical analysis of the results of this work, Eq. (5-11) and Eq. (5-14) were the best adapted to the data of the first and the second stage, respectively, for brick Ml.

The curves of the Ml experimental data are shown in Fig.5-6-2.

The slopes of the straight lines of R/(a) and R2(a) versus time t are the rate constants of the Ml brick hydration under different conditions for the first two stages. The reaction rates and R-square are presented in Table 5-6-1.

85 Fig. 5-6-2. Hydration rate equations, Rl(a) and R2(a), versus time, t, for Ml samples. (a) the first stage of hydration in steam at 121, 109, and 100°C, (b) the first stage of hydration in humidity at 100, 80, and 60°C, (c) the second stage of hydration in steam at 121, 109, and 100°C, (d) the second stage of hydration in humidity at 100, 80, and 60°C.

86 Table 5-6-1. The reaction rate constants, K [1/hr], for Ml samples.

Stage I Conditions Stage II y^-lO5 R-Square ^•lO5 R-Square 121°C 180 0.990 4.47 0.984 Steam 109°C 113 0.972 2.35 0.996 100°C 96.2 0.975 1.55 0.993 100°C 96.2 0.975 1.55 0.993 Humidity 80°C 48.4 0.999 1.32 0.993 60°C 8.40 0.994 0.19 0.998

The calculation showed that the rate constant in the first stage was 30 ~ 60 times higher

than those in the second stage. During the first stage of hydration, the chemical reaction

is the governing process, since at the beginning of hydration the surfaces of the magnesia particles are relatively "clean". The water molecules can directly contact the magnesia surface and hydration occurs easily. With the film of brucite formed on the pore wall, for

Ml there was about 0.2% of MgO hydrated when the second stage started. During the second stage, the water molecules had to diffuse through the film of brucite to continue the reaction, making the hydration much slower. This stage is controlled by the diffusion process.

The MSI showed clear three stage curves of the weight increase versus hydration time.

After examining all equations it was found that Equations (5-11) and (5-14) were the best suited to analyze the first and second stage, respectively. The rate constants are calculated and are presented in Table 5-6-2. The experimental data for the samples MS2 and M2 are best conformed to Equation (5-11). It is indicated that the hydration of MS2 and M2 are processes controlled by chemical reaction. The calculated rate constants for MS2 and M2 are shown in Table 5-6-2.

87 Table 5-6-2. The reaction rate constants, K*103 [1/hr], of MSI, MS2 and M2 samples in

steam and humidity at different temperatures.

MSI Conditions MS2 M2 Stage I Stage II 121°C 17.1 2.12 36.3 27.3 Steam 109°C 14.1 1.46 17.2 14.6 100°C 9.41 0.61 14.6 9.33 100°C 9.41 0.61 14.6 9.33 Humidity 80°C 2.54 0.40 4.51 5.95 60°C 0.28 0.01 0.34 0.44

Comparing the rate constants at 121°C, the following order of hydration rate is obtained:

MS2>M2>MS1 >M1

Equation (5-14) is the best equation to fit the experimental data for all the magnesia- chrome bricks after the second cycle of hydration. This means that after the second cycle, hydration was controlled by diffusion, which is proved by the results of the weight increases and the MOE variations. During the first cycle, the largest weight increase and highest values for MOE were achieved. The calculated rate constants are shown in Table

5-6-3.

Table 5-6-3. The rate constants, K-104 [1/hr], for the magnesia-chrome bricks in steam.

Temperature MCI MC2 MC3 MC4 MC5 MC6 130°C 2.14 2.70 28.4 86.2 (146°C) 4.57 15.9 121°C 2.02 2.55 20.2 12.2 3.67 15.0 109°C 0.13 0.74 9.20 2.60 1.01 1.80

The hydration of magnesia-chrome brick was much slower than that of magnesia-spinel and magnesite bricks, and the reason is probably that the periclase (MgO) content in

88 magnesia-chrome brick is less than that in the other types of bricks. The effects of temperature on the hydration are different. For example, at 121°C, for MCI the hydration rate constant was 14 times higher than at 109°C, but for MC2 and MC3, the hydration rates increased only 2-3 times. The order in terms of the rate constants from high to low at 109°C is: MC3 > MC4 > MC6 > MC5 > MC2 > MCI. This order is consistent with that of the CaG7SiC>2 ratio from high to low for the same MgO content; also at low MgO contents there are low rate constants.

-4 -

-14 T— 1 1 1 1 1 ——I

2.5 2.6 2.7 2.8 2.9 3 3.1

3 in x io , K 1

Fig. 5-6-3. The plots of ln AT versus 1/Tfor sample Ml.

The activation energies can be calculated using the logarithm form of Arrhenius equation

(5-15) from the plots shown in Fig. 5-6-3 and are listed in Table 5-6-4. The difference is

'89 probably caused by the chemical compositions, microstructures, and the physical

properties, such as porosity and permeability. Informatively, other values for activation

energy for the hydration reaction of magnesia from other studies performed by other

researchers are shown in Table 5-6-5.

Table 5-6-4. Activation energies, E (kJ/mol), for the hydration of all the experimental

bricks in steam and humidity.

MgO and MgO-Spinel Bricks MgO-Cr203 Bricks Stage I 52.5 Ml MCI 178 Stage II 50.5 MC2 82.7 M2 68.5 MC3 69.5 Stage I 73.8 MC4 MSI 123 Stage II 88.9 MC5 94.6 MS2 80.6 MC6 140

Table 5-6-5. Activation energies from Literature for the hydration of magnesia

Activation Reference: Type of MgO Energy, kJ/mol Layden & Brindley (1963)1" 151 72.7 Sintered magnesia (at 1800°C) Bratton & Brindley (1965)r231 67.3 Sintered magnesia (at 1800°C) Smithson & Bakhshi (1969)[12] 58.9 Caustic magnesia Fruwirth et al (1985)r71 70.6 MgO single crystals (100-200um) Kitamura et al (1996)[91 65.5 Single crystalline magnesia

5.7. The Mechanism of Hydration of Magnesia-Based Refractory Bricks

From the discussion above, the mechanism of hydration for magnesia-based refractory bricks is suggested. The hydration occurs in three stages, and each stage is governed by different process.

90 Initially water or water vapor goes into open pores of bricks and a layer of water forms

on the walls of pores. The water reacts with magnesia and the brucite nucleates and

grows. A film of brucite consisting of intercrossed crystals forms on the walls of pores

and works as a reinforcement for the brick structure, resulting in an increase in modulus

of elasticity, which corresponds to an increase in the thickness of the film. During the

first stage, the chemical reaction is the governing process, since at the beginning of

hydration the surfaces of the magnesia particles are relatively "clean". The water

molecules can directly contact the magnesia surface and hydration occurs easily.

With the hydration progress the MOE increases accompanying the thickening of the

brucite film, which slows down the hydration rate. In the second stage, the hydration

reaction is controlled by the diffusion of water through the brucite film. When the film

exceeds the "critical thickness", cracks form in the film because the stress caused by its

shrinkage during drying between hydration cycles exceeds the maximum strength, and

MOE begins to decrease. When hydration takes place at periclase grain boundaries,

brucite weakens the bonding of periclase grains, and the MOE further decreases.

When the MOE decreases to the initial value, the third stage starts, and cracks gradually

accumulate and link, leading eventually to the total destruction of the bricks' integrity.

The phenomenon is known as "dusting".

91 6. CONCLUSIONS

In this research, the hydration of magnesite, magnesia-spinel, and magnesia-chrome refractory bricks in 98% relative humidity, water, and steam environment at temperatures between 60 and 130°C has been studied. The main conclusions from this research are the following:

1. The chemical composition of a magnesia-based brick, especially CaO/Si02 ratio, has a strong effect on its hydration, and increasing this ratio will cause an increase in the hydration rates. For example, 0.26% of MgO hydrated in the magnesite brick of

CaO/Si02 = 1.8 in water at 60°C for 12 hours, but 0.85% of MgO hydrated when

CaO/Si02 = 7.5. It was also confirmed that a uniform distribution of high-in-silica glass phases would protect against hydration. The hydration rate increases with periclase

(MgO) content.

2. The hydration rate increases with temperature. For example, only 1.18% of MgO hydrated in the M2 brick (magnesite brick of 96.5%MgO) in steam at 100°C for 4 hours, but 4.15% of MgO hydrated at 121°C during the same time. Kinetically, for the M2 brick, the hydration rate constant at 100°C is only one-third of that at 121°C, and they are

0.009 [1/hr] at 100°C and 0.027 [1/hr] at 121°C, respectively.

3. During the course of hydration, a film of brucite forms and modulus of elasticity

(MOE) of a magnesia-based refractory brick increases from the initial value MOEo, to a

maximum MOEmax, corresponding to the increase in the film thickness, followed by a relatively slow decrease. When modulus of elasticity decreases below the initial value

92 MOEo, the sample progressively disintegrates. The maximum value of modulus of elasticity corresponds to a "critical thickness" of the brucite (MgCOFfh) film. When the brucite film exceeds the "critical thickness" it cracks, the grain boundaries weaken, and consequently the MOE begins to decrease. The cracks gradually accumulate and link, leading eventually to the total destruction of the bricks' integrity. This phenomenon is known as "dusting".

For example, modulus of elasticity for brick MC3 (magnesia-chrome brick of 62.0 %

MgO and 18.0 % CV2O3) increases from 35.3 GPa (i.e. for initial, un-hydrated state; hydration film thickness is zero), to the maximum of 40.6 GPa when the "critical thickness" of the brucite film reaches about 1 um.

4. The study of the influence of hydration on pore size and pore size distribution indicates that the hydration starts with filling the small pores, below 0.42um with the product of hydration (i.e. brucite), consequently, relatively increasing the number of the larger pores, for example, pores of 0.5um increases from 7.5% (as-received sample) to

11.2% (hydration for 8 hours). After hydration takes place at periclase grain boundaries, microcracking occurs at grain boundaries and the microcracks density increases with hydration progress. Therefore the pore size distribution shifts to larger sizes. For example, the peak value of the pore size distribution for brick MC3 shifted from 0.42um

(hydrated for 8 hours) to 0.50um (24 hours), further to 0.75 um (38 hours).

5. The brick permeability to gas decreases with hydration in the early stage.

However, after "dusting" occurs, the permeability increases rapidly. During hydration, the permeability of brick M2 increases by factor of 14, i.e. from 72 to 1032 milliDarcys.

93 6. The hydration mechanism of magnesia-based refractory bricks is proposed. The

hydration occurs in three stages, and each stage is governed by different process:

I. During the first stage the brucite film forms and the hydration progress is

controlled by chemical reaction between MgO and water. For example, for brick Ml

(magnesite brick of 93.0% MgO), the rate constant is 1.80xl0"3 [1/hr] in steam at 121°C

in Stage I and activation energy is 52.5 kJ/mol.

II. In the second stage, hydration slows down. The hydration progress is now controlled by diffusion of water through the brucite film. For example, for the Ml brick, the second stage starts when there is about 0.2% MgO hydrated and the rate constant is

4.47x 10"5 [1/hr] and activation energy is 50.5 kJ/mol in steam at 121°C in Stage II.

III. The third stage starts when "dusting" takes place. Due to the formation of new surfaces accessible to water through cracks at grain boundaries, hydration accelerates again, and eventually leads to total disintegration of the refractory brick.

7. The modulus of elasticity measurement by the impulse excitation technique can be used as a relatively simple nondestructive method to assess the hydration degree of magnesia-based refractory bricks.

94 7. FUTURE WORK

This work is an initial stage of the larger study of the hydration of magnesia-based

refractory bricks. For further studies are recommended as follows:

1. A qualitative relationship between the hydration degree and modulus of elasticity determined by the impulse excitation technique has been established. However, for more

accurate evaluation of the degree of hydration by the non-destructive method of impulse

excitation technique, the quantitative correlation between the degree of hydration and the

modulus of elasticity must be developed.

2. The brucite film affects modulus of elasticity, so do porosity and cracking. It is important to understand the relationship between the characteristics of the brucite film,

porosity, cracking and modulus of elasticity.

3. Hydration of basic bricks depends on their products. Studying the effect of microstructure, such as pores size and pore distribution, on hydration is necessary to improve our understanding of this process, and thus to enhance the hydration resistance of basic bricks.

4. The ultimate goal of this research is to develop solutions to suppress the hydration of magnesia-based refractories. Future work must therefore include careful evaluation of all possible factors, including those related to installation and in-service process parameters.

95 8. REFERENCES

1. H. He and G. Oprea, "Test Results on Brick Smaple for Impala Platinum

Furnace," Technical Report, UBCeram, unpublished, (2001).

2. E. G. King, et al., "Thermodynamic Data for Mg(OH)2 (Brucite)," U.S. Bureau of

Miners, Report of investigations 8401 (1975).

3. University of Colorado, "Mineral Structure data, Periclase Group," University of

Colorado, http://ruby.colorado.edu/~smyth/min/periclase.html.

4. University of Colorado, "Mineral Structure data, Hydroxide Group," University

of Colorado, http://ruby.colorado.edu/~smyth/min/periclase.html.

5. F. Giauque and R. C. Archibald, "The Entropy of Water from the Third Law of

Thermodynamics. The Dissociation Pressure and Calorimetric Heat of the Reaction

Mg(OH)2 = MgO + H20. The Heat Capacities of Mg(OH)2 and MgO from 20 to 300°K,"

Journal of the American Chemical Society, 59, No. 3, (1937), pp. 561-569.

6. D. T. Livey, "The Properties of MgO Powders Prepared by the Decomposition of

Mg(OH)2," Transactions of The British Ceramic Society, 47, No. 5, (1957), pp. 217-236.

7. University of Cambridge, http://www.esc.cam.ac.uk/highpt/highpt.html.

8. J. A. Malarria and Tinivella, "Degradation of Magnesite—Chrome Refractory

Brick by Hydration," Journal of the American Ceramic Society, 80, No.9, (1997), pp.

2262-68.

9. A. Kitamura, et al, "Hydration Characteristics of Magnesia," Taikabutsu

Overseas, 16, No. 3, (1996), pp. 3-11.

10. 0. Fruhwirth, et al, "Dissolution and Hydration Kinetics of MgO," Surface

Technology, 24, No. 3, (1985), pp. 301-317.

96 11. M. T. Frost, et al, "Application of Caustic Calcined Magnesia to Effluent

Treatment," Transactions of the Institution of Mining Metallurgy, Section C, 99, (1990), pp. C117-C124.

12. G. L. Smithson and N. N. Bakhshi, "The Kinetics and Mechanism of The

Hydration of Magnesium Oxide in A Batch Reactor," The Canadian Journal of Chemical

Engineering, 47, No. 10, (1969), pp. 508-513.

13. C. Henris, et al, " Morphological Study of Magnesium Hydroxide Nanoparticles

Precipitated in Dilute Aqueous Solution," Journal of Crystal Growth, 249, (2003), pp.

321-330.

14. W. Feitknecht and H. Braun, "Der Mechanismus der hydratation von magnesiumoxide mit Wasserdampf," Helvetica Chimica Acta, 50, No. 211, (1967), pp.

2041-2053.

15. D. R. Glasson, "Reactivity of Lime and Related Oxides. IX. Hydration of

Magnesium Oxide," Journal of Applied Chemistry, 13, No. 3, (1963), pp. 119-123.

16. G. K. Layden and G. W. Brindley, " Kinetics of Vapor-phase Hydration of

Magnesium Oxide," Journal of the American Ceramic Society, 46, No. 11, (1963), pp.

518-522.

17. J. Chown and R. F. Deacon, "The Hydration of Magnesia by Water Vapor,"

Transactions /British Ceramic Society, 63, No. 2, (1964), pp. 19-102.

18. A. S. Bhatti and D. Dollimore, "The Rates of Hydration of Sea Water

Magnesias," Surface Technology, 22, No. 2, (1984), pp.181-188.

19. O. Levenspiel, Chemical Reaction Engineering, (New York: John Wiley and

Sons, Second Edition, 1972), pp. 367-368.

97 20. H. Y. Sohn and M. E. Wadsworth, Rate Processes of Extractive Metallurgy (New

York: Plenum Press, 1979), pp. 134-135 in Ref. [21].

21. A. M. Ranjifham and P. R. Khangaonkar, "Hydration of Calained Magnesite at

Elevated Temperatures under Turbulent Conditions," Minerals Engineering, 2, No. 2,

(1989), pp. 263-270.

22. R. J. Bratton and G. W. Brindley, "Kinetics of Vapor Phase Hydration of

Magnesium Oxide: Part 2 - Dependence on Temperature and Water Vapor Pressure,"

Transaction of the Faraday Society, 61, No.5, (1965), pp.1017-1025.

23. J. Blaha, "Kinetics of Hydration of Magnesium Oxide in Aqueous Suspension.

Part I. Method of Measurement and Evaluation of Experimental Data," Ceramics -

Silikaty, 39, No.2, (195), pp.41-51.

24. M. Frith, T. Buffrey, and I. Strawbridge, "Magnesia: A Refractories Manufacturer

Perspective," British Ceramic Transactions, Vol. 97, No. 1, 1998, pp29-34.

25. G. M. Biggar, "Phase Equilibria in Chrome-Bearing Basic Refractories," The

Refractories Journal, 48, No.l, (1972), pp. 6-9.

26. R. I. Razouk an R. Sh. Mikhail, "Surface Properties of Magnesium Oxide," The

Journal of Physical Chemistry, 61, No.7, (1957), pp. 886-891.

27. M. G. Kim, et al., "Shape and Size of Crystalline MgO Particles Formed by the

Decomposition of Mg(OH)2," J. Am. Ceram. Soc, 71, No. 8, (1988), pp. C372-C375.

28. V. S. S. Birchal, et al, "The Effect of Magnesite Calcination Conditions on

Magnesia Hydration," Materials Engineering, 13, No. 14-15,(2000), pp. 1629-1633.

29. J. Staron and S. Palco, "Production Technology of Magnesia Clinker," American

Ceramic Society Bulletin, 72, No. 8, (1993), pp. 83-87.

98 30. R. C. Doman, et al, "Phase Equilibriums in the System CaO-MgO," Journal of the

American Ceramic Society, 46, No. 7, (1963), pp. 313-316.

31. M. H. Leipold, "Impurity Distribution in MgO," Journal of the American

Ceramic Society, 49, (1966), pp. 498-502.

32. M. H. Leipold, "Addenda to Impurity Distribution in MgO," Journal of the

American Ceramic Society, 50, (1967), pp. 628-629.

33. T. V. Chusovitina, et al, "Effect of Calcium Oxide on the Hydration Activity of

the Powders of Electrical - Engineering Periclase," Refractories (Russian), 29, No. 7-8,

(1988), pp. 453-457.

34. A. Watanabe, et al., "Magnesia Containing Basic Castables," Proceedings of Unitec'95, Tokyo, Japan, 2, (1995), pp. 181-188.

35. N. L. Bowen and J. F. Schairer, "The System, Magnesia-Ferrous Oxide-Silica,"

American Journal of Science, 29, (1935), pp.153.

36. R. J. Bratton and G. W. Brindley, "Kinetics of Vapor Phase Hydration of

Magnesium Oxide: Part 3.— Effect of Iron Oxides in Solid Solution," Transaction of the

Faraday Society, 62, No. 10, (1966), pp. 2909-2915.

37. S. Chen, et al, "Effect of Additives on the Hydration Resistance of Materials

Systemized form the Magnesia-Calcia System," Journal of the American Ceramic

Society, 83, No. 7, (2000), pp. 1810-1812.

38. G. K. Layden and M. C. McQuarrie, "Effect of Minor Additions on Sintering

MgO," Journal of the American Ceramic Society, 42, (1959), pp. 89-92.

39. K. Matsui and K. Saihi, "Effects of Impurities on the Hydration of Magnesia

Clinkers," Proceedings ofUnitecr'95, Tokyo, Japan, 2, (1995), pp. 513-520.

99 40. J. Green, "Review Calcination of Precipitated Mg(OH)2 to Active MgO in the

Production of Refractory and Chemical Grade MgO," Journal of Material Science, 18,

(1983), pp. 637-651.

41. S. Spinner and W. E. Tefft, "A Method for Determining Mechanical Resonance

Frequencies and for Calculating Elastic Moduli from These Frequencies," Proceedings

ASTM, 61, (1961), pp. 1221-1238.

42. ASTM standard, C 1259-01, "Standard Test Method for Dynamic Young's

Modulus, Shear Modulus, and Poisson's Ratio for Advanced Ceramics by Impulse

Excitation of Vibration" (2001).

43. D. Munz and T. Fett, Ceramics: Mechanical Properties, Failure - Behaviour,

Materials Selection (Berlin; New York: Springer, 1999), pp. 17.

44. F. A. L. Dullien, Porous Media: Fluid Transport and Pore Structure (San Diego:

Acadamic Press Inc., Second Edition, 1992), pp. 9.

45. W. D. Kingery, et al, Introduction to Ceramics (New York: John Wiley & Sons,

Second Edition, 1975), pp. 185-186.

46. User's Manual for Capillary Flow Porometer (Version 6.0), by Porous Materials,

Inc.: D3-D5.

47. Porous Material Inc., "Capillary Flow Porometer," Porous Material Inc., http://www.pmiapp.com/publications/docs/capillary_flow_porometer.pdf.

48. K. Goto and W. E. Lee, "The 'Direct Bond' in Magnesia Chromite and Magnesia

Spinel Refractories," Journal of the American Ceramic Society, 78, No. 7, (1995), pp.

1753-1760.

100 49. P. Lauzon, et al., "Hydration Studies on Magnesia-Containing Refractories,"

Proceedings ofUNITECR '03, Osaka, Japan, (2003), pp. 54-57.

50. W. D. Kingery, et al, Introduction to Ceramics (New York: John Wiley & Sons,

Second Edition, 1975), pp. 386.

101