You are on page 1of 46

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/262578744

Particle Size Distribution of Limestone Fillers: Granulometry and Specific


Surface Area Investigations

Article  in  Particulate Science And Technology · April 2014


DOI: 10.1080/02726351.2013.873503

CITATIONS READS

19 928

2 authors:

Frédéric Michel Luc Courard


University of Liège University of Liège
38 PUBLICATIONS   484 CITATIONS    135 PUBLICATIONS   1,718 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

VALDEM - Valorization of demolition waste View project

VALDEM - Valorization of demolition waste View project

All content following this page was uploaded by Luc Courard on 02 September 2016.

The user has requested enhancement of the downloaded file.


Materials and Structures
Characterization of Particle Size, Surface Area, and Shape of Supplementary
Cementitious Materials
--Manuscript Draft--

Manuscript Number:

Full Title: Characterization of Particle Size, Surface Area, and Shape of Supplementary
Cementitious Materials

Article Type: Original Research

Keywords: Supplementary cementitious materials; particle size measurement; laser diffraction;


Blaine; BET; image analysis

Corresponding Author: Nele De Belie

BELGIUM

Corresponding Author Secondary


Information:

Corresponding Author's Institution:

Corresponding Author's Secondary


Institution:

First Author: Eleni C. Arvaniti

First Author Secondary Information:

Order of Authors: Eleni C. Arvaniti

Maria C.G. Juenger

Susan A. Bernal

Josée Duchesne

Luc Courard

Sophie Leroy

John L. Provis

Agnieszka Klemm

Nele De Belie

Order of Authors Secondary Information:

Abstract: The physical characterization of supplementary cementitious materials (SCMs) is of


great importance in optimizing the use of these materials in concrete in order to
minimize cost and maximize performance. Techniques that are currently used for the
determination of particle size, specific surface area and shape of cementitious
materials are described and critically evaluated in this paper. Recommendations for
testing using air permeability, sieving, laser diffraction, BET, image analysis and MIP
are also provided, representing an output from the work of the RILEM Technical
Committee on Hydration and Microstructure of Concrete with Supplementary
Cementitious Materials (TC-238-SCM).

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript
Click here to download Manuscript: paper_RILEM_TC_SCM_WG1_final.docx
Click here to view linked References

1 1 Characterization of Particle Size, Surface


2
3
4 2 Area, and Shape of Supplementary
5
6
7
3 Cementitious Materials
8
9 4 Eleni C. Arvaniti1, Maria C.G. Juenger2, Susan A. Bernal3, Josée Duchesne4, Luc
10
11 5 Courard5, Sophie Leroy5, John L. Provis3, Agnieszka Klemm6, Nele De Belie1
12
13 1
6 Magnel Laboratory for Concrete Research, Department of Structural
14
15
7 Engineering, Faculty of Engineering and Architecture, Ghent University,
16 8 Technologiepark Zwijnaarde 904, B-9052 Ghent, Belgium
17 2
18
9 University of Texas at Austin, Department of Civil, Architectural and
19 10 Environmental Engineering, 301 E. Dean Keeton St. C 1748, Austin, Texas
20 11 78712, USA
21 3
22 12 Department of Materials Science and Engineering, University of Sheffield, Sir
23 13 Robert Hadfield Building, Mappin St, Sheffield S1 3JD, United Kingdom
24 4
25 14 Département de géologie et de génie géologique, Université Laval, Pavillon
26 15 Adrien-Pouliot, local 4507, 1065, ave de la Médecine Québec, Qc, Canada, G1V
27 16 0A6
28 5
29 17 GeMMe research group, ArGEnCo Department,University of Liege, Liege,
30 18 Belgium
31 6
32 19 Department of Construction and Surveying / School of Engineering and Built
33 20 Environment, Glasgow Caledonian University, Cowcaddens Road, Glasgow, G4
34 21 0BA, UK
35
36 22
37
38 23 Abstract
39 24 The physical characterization of supplementary cementitious materials (SCMs) is
40
41 25 of great importance in optimizing the use of these materials in concrete in order to
42
43 26 minimize cost and maximize performance. Techniques that are currently used for
44
45 27 the determination of particle size, specific surface area and shape of cementitious
46
47 28 materials are described and critically evaluated in this paper. Recommendations
48
49
29 for testing using air permeability, sieving, laser diffraction, BET, image analysis
50 30 and MIP are also provided, representing an output from the work of the RILEM
51
52 31 Technical Committee on Hydration and Microstructure of Concrete with
53
54 32 Supplementary Cementitious Materials (TC-238-SCM).
55
56 33 Keywords: Supplementary cementitious materials, particle size measurement,
57
58 34 laser diffraction, Blaine, BET, image analysis
59
60
61
62
63 1
64
65
35 1 Introduction
1
2
3 36 The use of supplementary cementitious materials (SCMs) in the production of
4
5 37 concrete has increased worldwide over the past few decades [1,2]. These materials
6
7 38 can enhance the mechanical and durability properties of concrete, and contribute
8
9 39 to mitigation of the environmental impact associated with the construction
10 40 industry. SCMs are used as a partial replacement for Portland cement in concrete
11
12 41 reducing the fraction of Portland cement required to produce concrete with desired
13
14 42 performance. SCMs are mostly by-products of industrial processes such as fly
15
16 43 ashes derived from the coal-burning processes, blast furnace slags from the iron-
17
18 44 making industry, and silica fume from ferro-silicon industries [3]. However, in
19
20
45 recent years, greater attention has been given to natural materials with pozzolanic
21 46 activity such as calcined shales and clays along with metakaolin. Calcined clays
22
23 47 and shales are used as cement replacements, while metakaolin is more often used
24
25 48 as an additive to the cement.
26
27 49 The performance of SCMs in concrete is strongly dependent on their physical and
28
29 50 chemical characteristics, which vary depending on the nature and source of the
30
31 51 SCM. In general, the fineness is one of the most important physical properties
32
33 52 controlling the reactivity of SCMs and the subsequent strength development of
34
35 53 blended binders [4]. Reducing the average particle size increases the rate of
36
37
54 dissolution of the SCM, raising the pozzolanic activity and thus the development
38 55 of more strength-giving hydration products that enhance the long-term
39
40 56 performance of the concrete. Small particles can also facilitate nucleation and
41
42 57 growth of cement hydration products on the SCM surfaces, speeding up the early
43
44 58 cement hydration and therefore the strength development. However, reducing the
45
46 59 particle size of SCMs beyond an optimal value usually leads to an increased water
47
48
60 demand of the concrete mixtures to achieve a desired workability, which can
49 61 negatively affect both strength and its durability [5]. Further, if particle size is
50
51 62 decreased by grinding, this requires additional energy costs.
52
53
54
63 For most industrial control purposes, the primary characteristics measured in
55 64 powders are specific surface area, particle size distribution, particle shape, and
56
57 65 density. The specific surface area (defined on a mass basis) is the most common
58
59 66 property used to describe the fineness of Portland cement [6]. This surface area is
60
61 67 an integral parameter and gives no information about details of the actual particle
62
63 2
64
65
68 size distribution, which is probably of greater importance in defining concrete
1
2 69 performance. The description of particle shape encompasses information about the
3
4
70 sphericity and angularity, which affect workability and also the physical
5 71 phenomena utilized for particle-size measurement [7].
6
7
8 72 Even though SCMs are widely used by the construction industry, their physical
9
10
73 characterization is challenging. As the most used SCMs are industrial by-
11 74 products, in most cases the quality of these materials cannot be controlled during
12
13 75 their production, resulting in materials with varied characteristics [8]. There are
14
15 76 standard methods used to determine particle characteristics for cement that may
16
17 77 not be as accurate when applied to SCMs. For instance, the air permeability test
18
19 78 for specific surface area (Blaine), which is widely used for characterizing Portland
20
21
79 cements [9], is based on the principle of resistance to air flow through a partially
22 80 compacted sample of cement. This method relies on the assumptions that there is
23
24 81 a relatively limited range of particle sizes in the material, with consistent inter-
25
26 82 particle interactions, and that there are available and internationally accepted
27
28 83 reference powders with properties similar to the material of interest. These
29
30 84 conditions may not apply for all SCMs.
31
32 85 Deviations from expected results in physical characterization of powders are
33
34 86 associated with instrument limitations, improper sample preparation procedures
35
36 87 (e.g. inadequate dispersion), operator errors (e.g. improper instrument set-up or
37
38
88 poor calibration), or incorrect sampling [10]. Although numerous techniques for
39 89 the measurement of the physical characteristics of powders have been developed,
40
41 90 most of the techniques are unsatisfactory in some respect, and there is no general
42
43 91 method that may be applied with a reasonable confidence to a wide range of
44
45 92 materials spanning several orders of magnitude in particle size, and with diverse
46
47 93 particle shapes.
48
49 94 The key consideration for the proper and accurate determination of physical
50
51 95 properties of SCMs lies in the selection of the adequate instruments and methods.
52
53 96 This paper presents a critical overview of the techniques that are currently used
54
97 for the determination of the particle size, specific surface area and shape of
55
56 98 cementitious materials. The aim is to systematize the existing knowledge on this
57
58 99 subject and to identify the most suitable techniques and methods that can be
59
60 100 applied to characterize SCMs. The most important criteria that should be
61
62
63 3
64
65
101 considered as a starting point for method validation are also outlined. This paper
1
2 102 first discusses the methods of sample preparation prior to characterization, such as
3
4
103 sampling and dispersion, and then describes the standard methods for testing
5 104 specific surface area, particle size distribution, and particle shape, as well as tests
6
7 105 for properties such as density and refractive index that are used to provide
8
9 106 supporting information for the other tests. Finally, selected SCMs are
10
11 107 characterized with these methods in order to identify the factors that induce
12
13 108 variations in the results obtained from different techniques.
14
15
16 109 2 Methods for Sample Preparation
17
18
19 110 2.1 Sampling
20
21
22 111 Sampling is a very important step in the characterization of any material, as the
23
24
112 specimen selected has to be representative a large batch of material and provide
25 113 sufficient and accurate information to enable decisions on handling and use.
26
27 114 Sampling is one of the factors which can introduce the largest errors in particle
28
29 115 size, shape, and density measurements of a powder [11]. Whenever a powder is
30
31 116 analyzed, whether for physical or chemical assay, the quality of the measurement
32
33 117 depends on how representative the sample is of the bulk from which it is drawn.
34
35
118 Taking a sample of a few milligrams from a bulk of many tons, the chances of
36 119 measuring a non-representative sample are increased considerably. The
37
38 120 International Standard ISO 14488 [12], the American standards ASTM C 183-08
39
40 121 [13] and ASTM C311/C311M-13 [14] provide useful information on the
41
42 122 requirements for sampling of particulate materials.
43
44 123 Two types of sampling errors are possible [10]. First, statistical errors arising from
45
46 124 sample heterogeneity cannot be prevented, but can rather be estimated beforehand
47
48 125 and reduced by increasing the sample size. Even for an ideal random mixture, the
49
50 126 quantitative particle distribution in samples of a given magnitude is not constant
51
127 but is subject to random fluctuations. Second, there are errors that occur due to the
52
53 128 segregation of the bulk and depend on the previous history of the powder. Dry
54
55 129 powders tend to separate if they are stored for some time or they are vibrated
56
57 130 during storage. The larger particles tend to rise to the top and the smaller particles
58
59 131 collect at the bottom of the container. If the sample is taken from the top of the
60
61 132 container it will not contain the smaller particles, giving a biased measurement.
62
63 4
64
65
133 This error can be minimized by suitable mixing and building up a sample from a
1
2 134 large number of increments.
3
4 135 There are several different techniques of sampling that have been evaluated in
5
6 136 multiple studies (Table 1). The spinning riffle is the most reproducible method for
7
8 137 obtaining a representative sample for powdered materials when compared with
9
10
138 other techniques such as scoop sampling, table sampling, cone and quartering and
11 139 chute riffling [10].
12
13
14
140 Table 1 Methods of powder sampling and associated error (data from [10] and [15])
15 Relative Standard Estimated maximum
16 Method
17
Deviation (%) sample error (%)
18 Cone & Quartering 6.81 22.70
19 Scoop Sampling 5.14 17.10
20 Table Sampling 2.09 7.00
21
22
Chute Riffling 1.01 3.40
23 Spin Riffling 0.125 0.42
24
25 141
26
27
28 142 2.2 Dispersion
29
30
31 143 The term ‘dispersion’ has a variety of meanings, but in this context it makes
32
33
144 reference to the process of separating solid particles from each other to measure
34 145 the physical characteristics of a given powder. For the characterization of SCMs,
35
36 146 dispersion is particularly important in accurate determination of particle size
37
38 147 distribution; however, if the natural, agglomerated state is of interest, this should
39
40 148 be taken into account during the sample preparation to avoid the break-up of
41
42 149 agglomerated particles. In either case, the dispersion medium, whether air or
43 150 liquid, should not cause irreversible changes to the particle size through processes
44
45 151 such as dissolution, grinding or aggregation.
46
47
48 152 The overall process of dispersion of a powder in liquid media comprises three
49
153 main stages: (1) wetting of the powder, (2) breakdown of particle clusters and (3)
50
51 154 flocculation of the dispersed particles [16]. The wetting of the particles depends
52
53 155 on different factors including the nature of the liquid phase, the character of the
54
55 156 surface, and the dimensions of the interstices in the clusters. The wetting
56
57 157 effectiveness is influenced by the existing forces between individual particles in
58
59 158 the clusters. The major difficulty during the dispersion process is achieving the
60 159 breakup of agglomerates to finer particles after the powder has been wetted, as
61
62
63 5
64
65
160 particles tend to adhere together by weak forces unless the wetting is very strong
1
2 161 (i.e. contact angle is very low). In such cases, the penetration of liquid into the
3
4
162 particles can generate sufficient pressure to promote the dispersion.
5
6 163 Dispersibility is then defined as the ease with which a dry powder can be
7
8 164 dispersed in a particular liquid medium, which is essentially a measure of the
9
10
165 effectiveness of the first two stages of the dispersion process. The dispersability is
11 166 dependent on lyophilicity, particle size, specific gravity and ionic charges on the
12
13 167 surface (zeta-potential) of the material [17]. Dispersability is especially important
14
15 168 when characterizing SCMs because there are situations (e.g. silica fume) in which
16
17 169 the particles are highly agglomerated in the dry state, and therefore must be
18
19 170 properly dispersed in order to determine the “true” particle size distribution
20
21
171 (PSD). There are no accepted standard methods for dispersing SCM particles prior
22 172 to analysis, and therefore the degree of dispersion will vary depending on the
23
24 173 method that is used. This could potentially introduce a large source of variation at
25
26 174 the sample preparation stage [18].
27
28
29 175 2.3 Outgassing
30
31
32
176 Outgassing involves the heating of the powder at a given temperature in helium or
33 177 nitrogen flow. Outgassing strongly influences the measurement of specific surface
34
35 178 area; testing partially moist particles covered with molecules of previously
36
37 179 adsorbed gases or vapor, can lead to reduced or variable specific surface values
38
39 180 [19]. There are no established outgassing conditions to assure accurate
40
41 181 measurements of specific surface of powders. Most laboratories use their own
42
182 standard protocol at an established temperature, gas pressure and time of
43
44 183 outgassing, independent of the chemistry and structure of the material to test.
45
46
47 184 The ideal practice for determining the outgassing conditions should involve the
48
49
185 study of the potential impact of different outgassing conditions on the physical
50 186 and chemical properties of the tested material, to make sure that the original
51
52 187 surface of the particles evaluated is preserved after this treatment. For most
53
54 188 purposes, the outgassing temperature can be selected within the range where the
55
56 189 thermogravimetric trace of the tested powder exhibits a minimum slope [20].
57
58 190 Outgassing can be conducted at room temperature (20 - 25C) when the powder is
59
60 191 treated with a combined purge of a non-reactive, dry gas flow under vacuum, or
61
62
63 6
64
65
192 when the specimens are subjected to desorption-adsorption cycles. These methods
1
2 193 of outgassing are strongly recommended when analyzing materials that can suffer
3
4
194 structural changes when exposed to elevated temperatures.
5
6 195 The effect of the outgassing conditions on the specific surface measurements of
7
8 196 the minerals goethite, quartz, calcite and kaolinite has been evaluated by Clausen
9
10
197 and Fabricius [19], where it was observed in goethite that increased outgassing
11 198 temperatures promoted a rise in the specific surface area, reaching a maximum
12
13 199 value at temperatures between 150C and 250C. This is associated with the phase
14
15 200 changes from a ferrihydrite to disordered hematite. Conversely for quartz, calcite
16
17 201 and kaolinite after dry heating above 100C the BET values are close to stable.
18
19 202 BET values calculated in these mineral by heating the specimens between 20C
20
21 203 and 250C do not promote significant changes in the values obtained. This
22
23 204 indicates that outgassing at room temperature of materials containing oxide
24
25 205 minerals, such as SCMs, can enable measurement of reliable specific surface
26
27
206 areas, as long as the outgassing time is prolonged (2h).
28
29
30 207 3 Methods for Particle Size Characterization of
31
32 208 SCMs
33
34
35 209 The classical techniques for determining fineness and particle size distribution in
36
37 210 cementitious materials include sieving, air permeability testing (Blaine), gas
38
39 211 adsorption (BET), laser light scattering, and image analysis. Mercury intrusion
40
41 212 porosimetry (MIP) for particle size analysis is also considered in this study, even
42
213 though it is not a widely used technique.
43
44
45 214 3.1 Sieving analysis
46
47
48 215 The simplest means of assessing the fineness of SCMs is through a sieve analysis
49
50 216 since it does not require the specialized instrumentation used in other methods.
51
52 217 However, the information obtained from sieve analysis of a fine powder is more
53
218 limited than that obtained through more sophisticated methods, as the most
54
55 219 commonly used sieve analysis procedures for powders utilize only one sieve size.
56
57 220 For example, ASTM C618 [21] specifies a maximum of 34% by mass retained on
58
59 221 the no. 325 sieve (45 μm opening size) for fly ash and natural pozzolans when
60
61 222 wet-sieved. Like density testing, the standard method used is one for portland
62
63 7
64
65
223 cement, ASTM C430 “Standard Test Method for Fineness of Hydraulic Cement
1
2 224 by the 45-μm (No. 325) Sieve” [22]. The test standardizes the material mass to be
3
4
225 tested, the water pressure and nozzle type, and sieve calibration procedures. As
5 226 long as the material is in contact with water does not react. Sources of error in this
6
7 227 test can arise if the pozzolans are not adequately dispersed, as in densified silica
8
9 228 fume for example; agglomerated particles will not pass the sieve opening under
10
11 229 the low water-pressure specified.
12
13 230 An alternative to the wet sieving process is to use a dry, forced-air process to
14
15 231 sieve SCMs as described in EN196-6. This process can be more rapid than a wet
16
17 232 process since the sample does not need to be dried after testing. For example,
18
19 233 Hooton and Buckingham [23] demonstrated that an air-jet sieve using forced air
20
21
234 from a pressure-controlled vacuum measured the percent fly ash retained on the
22 235 no. 325 sieve in approximately 2 minutes. The values obtained are necessarily
23
24 236 slightly different than those obtained using the wet-sieve process, tending to be
25
26 237 slightly lower [33], but can be corrected using empirically-determined calibration
27
28 238 factors.
29
30
31 239 3.2 Air permeability test (Blaine)
32
33 240 Air permeability methods measure the resistance of flow of air through a packed
34
35 241 bed of cement of known dimensions and porosity. The time taken for a fixed
36
37 242 quantity of air to flow through a compacted material bed of specified dimension
38
39 243 and porosity is measured. Under standardized conditions, the specific surface of
40
41 244 the material, commonly referred to as the Blaine fineness, is proportional to t,
42
43 245 where t is the time for a given quantity of air to flow through the compacted bed.
44 246 The number and size range of individual pores in the specified bed are determined
45
46 247 by the particle size distribution, which also influences the time for the specified
47
48 248 air flow.
49
50 249 In 1939, Lea and Nurse introduced the constant flow-rate method that forms the
51
52 250 basis of British Standard BS 4550 [24]. A simpler, constant volume method that is
53
54 251 widely used in USA, UK and many other countries was developed by Niesel [25].
55
56 252 The apparatus (Fig. 1) consists of a U-tube manometer, a plunger, a permeability
57
58 253 cell and a perforated disc. It is calibrated according to the Lea and Nurse method
59
60 254 and the results are analyzed using the Carman-Kozeny equation [26] for viscous
61 255 flow through a bed, which involves knowledge of the density of the cement (or of
62
63 8
64
65
256 the SCM). However, this ensures that the bed is uniform (which is very difficult to
1
2 257 achieve for platy-shaped particles such as those in most metakaolins), and that
3
4
258 none of the particles are very highly irregular in shape (which is violated for fly
5 259 ashes with punctured cenospheres and/or unburnt coal residues, or rice husk ashes
6
7 260 retaining some of the geometry of the original plant material).
8
9 261
10
11 Standard taper –
12 female coupling to
13 fit bottom of cell
14
15 Valve or clamp
16
17
18
19
20
21
22
23 Glass tube
24
25
26
27
28
29
30 262
31
32 263 Fig. 1 Blaine air permeability apparatus
33 264
34
35 265 Today, two updated standards are in widespread use for the analysis of cement by
36
37 266 this technique, the American standard ASTM C204-07 [27] and the European
38
39 267 Standard EN 196-6 [28]. The bed of the material is prepared in a special
40
41 268 permeability cell to porosity e=0.500±0.005. The weight of the sample (m1) is
42
43
269 calculated from Equation 1:
44
45 270 [1]
46
47
48 271 where ρ is the density of the cement [g/cm3], and V is the volume of the bed [cm3].
49
50 272 The specific surface area, S, is expressed as:
51
52
53 273 [2]
54
55
56 274 where K is the apparatus constant, e the porosity of the bed, t the measured time
57
58 275 [s], ρ the density of the sample [g/cm3], and η is the viscosity of air at the test
59
60 276 temperature.
61
62
63 9
64
65
277 The apparatus constant is determined by measuring the permeability of the
1
2 278 reference material of a known specific surface area. Fine materials other than
3
4
279 cement may prove difficult to form into a compacted bed of porosity e = 0.500 as
5 280 described in this method. The reason for this may lie in the fact that thumb
6
7 281 pressure on the plunger cap may fail to bring it in contact with the top of the cell
8
9 282 or, after making contact and removing the pressure, the plunger may move
10
11 283 upwards as the bed restores semi-elastically to a larger volume. The porosity of e
12
13 284 = 0.500 is therefore likely to be unattainable for some materials, and this issue
14
15
285 will be revisited in section 4.2.
16
17 286 For such cases, the porosity required for a well-compacted bed needs to be
18
19 287 determined experimentally. The mass of the material to make the bed (m) then
20
21
288 becomes, in grams:
22
23 289 [3]
24
25
26 290 The specific surface area, S, is determined by Equation 4.
27
28
29
30 291 [4]
31
32
33
34 292 where e0 is the porosity of the bed of the reference material, S0 is the specific
35
36 293 surface of the reference material [cm2/g], t0 is the mean of the three flow times
37
38 294 measured on the reference material [s], ρ0 is the density of the reference material
39
40 295 [g/cm3], and η0 is the air viscosity at the mean of the three temperatures at which
41
42
296 the three triplicate measurements for the reference material were collected [Pa·s].
43
44 297 The air permeability test is a simple and rapid method that is used in the cement
45
46 298 industry [29]. However, air permeability is an indirect method and suffers from a
47
48
299 number of weaknesses, including an inability to account for variable particle
49 300 shape and bed tortuosity. The bed porosity is assumed to be close to 0.500, which
50
51 301 can lead to serious errors with some non-cement materials [30]. A significant
52
53 302 amount of the surface area of pores and cracks do not contribute to the flow
54
55 303 resistance, and so a lower result than expected may be obtained.
56
57 304 The method is comparative rather than absolute and, therefore, a reference sample
58
59 305 of known specific surface is required for calibration of the apparatus. The
60
61 306 reference material must have similar shape, particle size distribution, and surface
62
63 10
64
65
307 properties to the material of interest or it cannot be a valid comparison. In
1
2 308 addition, the test is designed for cement, so it becomes extremely unreliable at
3
4
309 surface areas greater than 500 m2/kg [29]. Its application to fly ash was suggested
5 310 to be of debatable value because of the unknown extent of influence of the
6
7 311 internal surface of unburned carbon particles present. Kiattikomol et al. [31] came
8
9 312 to the conclusion that Blaine fineness may not be sufficient to indicate the
10
11 313 fineness of fly ash, especially for fly ash with spongy phases. The full form of the
12
13 314 Carman-Kozeny equation includes an explicit sphericity term, which is
14
15
315 incorporated into the apparatus constant K in the Blaine method, and so it is
16 316 essential that the reference material is of a similar particle shape to the sample to
17
18 317 be analyzed.
19
20
21 318 3.3 Brunauer, Emmett and Teller (BET) Surface Area Analysis
22
23
24 319 In contrast to air permeability, the BET technique is a fundamental measurement
25
26
320 of specific surface area because it makes no assumption about the shape of the
27 321 particles. The BET method is based on the adsorption of a gas on the surface of
28
29 322 the solid, including any surface pores and cracks that the gas molecules can
30
31 323 access, and calculating the amount of adsorbed gas corresponding to a
32
33 324 monomolecular layer on the surface. Nitrogen is the most commonly used gas, but
34
35 325 any other inert gas can in principle be used. The physical adsorption of the gas
36
326 results from Van der Waals forces between the gas molecules and the adsorbent
37
38 327 surface area of the powder. The measurements are conducted at low temperature
39
40 328 (often the boiling point (-196 °C) of liquid nitrogen at atmospheric pressure, when
41
42 329 N2 is the probe molecule) and the amount of gas adsorbed can be measured by a
43
44 330 volumetric or continuous flow procedure.
45
46 331 Prior to BET analysis it is necessary to remove the gases and vapors that can be
47
48 332 physically adsorbed on the surface of the particles. This procedure is known as
49
50 333 outgassing (Section 2.3). It is important to bear in mind that BET analysis has
51
52 334 some limitations, as the possibilities of micropore filling or penetration into
53 335 cavities of molecular size are not considered in the measurement, which can
54
55 336 generate false results. When characterizing a material it is recommended to
56
57 337 measure at least three, but preferably five or more points, in the adequate pressure
58
59 338 range on the N2 sorption isotherm, to obtain reliable results. The conditions of
60
61
62
63 11
64
65
339 outgassing, the temperature of the measurements and the range of linearity of the
1
2 340 BET plot should be reported along with the BET values [20].
3
4
5 341 3.4 Laser diffraction
6
7 342 Laser diffraction (LD) is rapidly becoming a more popular method for particle
8
9 343 size determination [18]. Technology and instrument characteristics have rapidly
10
11 344 developed in the last decades, and now LD is considered to be one of the quicker,
12
13 345 easier and more reproducible methods of characterizing particle size, because it
14
15
346 provides a complete picture of the full size distribution [32]. However, in laser
16 347 diffraction the mathematical models used assume that the material is isotropic and
17
18 348 consists of particles which can be approximated as spheres, meaning that the size
19
20 349 of the particle is determined as the diameter of a spherical particle with an
21
22 350 equivalent volume. These assumptions do not always hold for cements and SCMs,
23
24 351 as will be discussed in more detail below.
25
26 352 The International Standard ISO 13320 [33] on Particle Size Analysis for Laser
27
28 353 Diffraction Measurements is an introduction to laser diffraction particle sizing
29
30 354 systems giving information on theory, guidance on both dispersion and sampling,
31
32
355 and a methodology for proper quality control. However, the process by which a
33 356 method can be validated is not clear from this document.
34
35
36 357 It should be made clear that laser diffraction instruments do not measure particle
37
38
358 size distributions (PSD). What is measured is the light scattered by the particles.
39 359 To relate this to the particle size distribution, critical assumptions are made about
40
41 360 the optical properties of the material under analysis. A mathematical model is
42
43 361 needed to convert light scattering data to particle size distribution. Two optical
44
45 362 models are commonly used to calculate PSD, the Fraunhofer diffraction model
46
47 363 and the Mie theory.
48
49 364 The Fraunhofer approximation assumes that: (1) the particle being measured is
50
51 365 much larger than the wavelength of the light employed (ISO13320 defines this as
52
53 366 being greater than 40λ, i.e. 25μm when a He-Ne laser is used), (2) all sizes of
54
55
367 particles scatter with equal efficiencies, and (3) the particles are opaque,
56 368 transmitting no light. The Fraunhofer model does not make use of any knowledge
57
58 369 of the optical properties of the sample, and only scattering at the contour of the
59
60 370 particles (i.e. diffracted light) is considered to calculate the projected area of a
61
62
63 12
64
65
371 sample. It is important to note that diffraction is independent of the composition
1
2 372 of the particles [34], unlike the reflection and refraction that are not considered in
3
4
373 this model.
5
6 374 Mie theory [35], on the other hand, is a more accepted theory used in LD
7
8 375 measurements. The latest laser diffraction instruments use the full Mie theory,
9
10
376 which completely solves the equations for interaction of light with matter
11 377 (including diffraction, reflection and refraction of light). This can provide accurate
12
13 378 results over a large size range (typically 0.02 – 2000μm), as long as the optical
14
15 379 properties (refractive and absorption indices) of both the material and medium are
16
17 380 known. The Mie theory determines the volume of the particle, as opposed to
18
19 381 Fraunhofer model which predicts size based on a projected area. Whereas the
20
21
382 Fraunhofer approximation is not suitable for samples that are transparent or
22 383 semitransparent, and for small particles (i.e. less than 50 μm), Mie theory can
23
24 384 apply under these conditions [33].
25
26
27
385 Optical parameters
28 386 In order to apply the Mie theory, the optical parameters of the tested particles are
29
30 387 required. The optical properties determine how light interacts with a material, and
31
32 388 are defined by the complex index of refraction, ñ:
33
34
35
389 ñ=n - ik [5]
36
37 390 where i = and n and k are the real and the imaginary parts of the complex
38
39
40
391 index of refraction. The real part, n, is called the refractive index while k is the
41 392 absorption (or extinction) coefficient. Both coefficients are dependent on the
42
43 393 frequency of light and standard refractive index measurements (n) are often
44
45 394 tabulated for wavelengths emitted by a sodium flame or a sodium vapor lamp
46
47 395 (589.3 nm) designated “D” [36]. The absorption coefficient ‘k’ is 0 or very close
48
49 396 to 0 for transparent or translucent material, and becomes important for opaque
50
397 media. In thin sections under the microscope a material will look opaque if its k-
51
52 398 value (the absorption coefficient), is greater than 0.01 [37].
53
54
55 399 The real part of the refractive index of a liquid or a gas is easily measured using a
56
57
400 suitable refractometer. However, the refractive index of a solid material in the
58 401 form of a fine powder is more difficult to measure, due mainly to the fact that the
59
60 402 material is often composed of more than one phase, which differ in their optical
61
62
63 13
64
65
403 properties. In fact, only a limited number of materials have a single, isotopic
1
2 404 refractive index: crystals belonging to the cubic crystal system, glasses and
3
4
405 amorphous substances. Minerals in other crystallographic systems are anisotropic,
5 406 with different refractive indices depending on crystal alignment with 2 or 3 main
6
7 407 values described as α, β and γ in tabulations of reference data [36,38]. The real
8
9 408 part of the complex index of refraction can be determined under an optical
10
11 409 microscope using the immersion method, in which the index of a solid is
12
13 410 compared with that of a liquid of known index [39]. A material grain which is
14
15
411 immersed in a liquid of matching refractive index as itself disappears from view.
16 412 However, if the liquid is of a different refractive index the grain stands out,
17
18 413 surrounded at the interface between the grain and the liquid by a thin band of light
19
20 414 known as the Becke line [39]. The greater the difference in the refractive indices
21
22 415 between the fragment and immersion, the greater is the intensity of the interface.
23
24 416 By using a series of liquids of varying refractive indices, the refractive index of
25
417 the material grain can be determined [39].
26
27
28 418 Direct measurement of the fundamental optical properties (refractive index n and
29
30 419 extinction coefficient k) can also be constructed using spectroscopic ellipsometry,
31
420 which is an optical measurement technique that characterizes light polarization
32
33 421 after reflection (or transmission) from a sample over a wide spectral range [40].
34
35
36 422 The use of correct input values for the optical parameters is very important for
37
38
423 particle size determination of smaller particles. When particles are large and have
39 424 a high refractive index difference compared to the suspension medium (e.g. air),
40
41 425 and if the absorption is low [41], the errors resulting from incorrect input of these
42
43 426 parameters are much smaller. Zhang and Xu [42] reported results on the effect of
44
45 427 particle refractive index on size measurement and noted that ‘it is well known
46
47 428 from Mie theory that the scattered light differs for particles with different
48 429 refractive indices, although the size or the distribution of the particles may be the
49
50 430 same’. Their conclusion was that if an incorrect refractive index is assumed when
51
52 431 the particle size distribution is computed from a measured scattered energy
53
54 432 distribution, a 10% error will be involved under most circumstances, but greater
55
56 433 errors may occur if the assumed refractive index is much less or greater than the
57
58 434 actual one.
59
60
61
62
63 14
64
65
435 In the case of Portland cement, the refractive index is not a single value since
1
2 436 cement is a multiphase powder. Mean values are often calculated based on the
3
4
437 known optical properties for each constituent pure phase [43]. According to
5 438 Hackley et al. [41] for cementitious powders, absorption becomes important for
6
7 439 the fine fraction, below about 1 μm in diameter, where it can have a large impact
8
9 440 on the particle size distribution. For n ≥ 1.6 (fairly refractive materials), the model
10
11 441 is not very sensitive to the choice of n for weakly absorbing or transparent
12
13 442 materials (i.e., k < 0.1). It is only moderately sensitive at k = 0.1. The magnitude
14
15
443 of the calculated submicron fraction depends on the choice of k, with the
16 444 dependence being stronger as n becomes smaller [41].
17
18
19 445 Table 2 presents refractive indices n values for phases often present in Portland
20
21
446 cement and SCMs.
22
23 447 Table 2 Typical refractive index values (n) for phases often present in cements and SCMs
24 Phase n ref Phase n ref
25
26 Pure C3S 1.7139-1.07238 [44] β-C2S 1.717-1.735 [44]
27
28 C4AF 1.96-2.04 [44] γ-C2S 1.642-1.654 [44]
29 Arcanite 1.4935-1.4973 [44] Ca(OH)2 1.545-1.573 [44]
30
31 Gypsum 1.5205-1.5296 [44] Hemihydrate 1.559-1.5836 [44]
32 γ-CaSO4 1.505-1.548 [44] Syngénite 1.5011-1.5176 [44]
33
34 MgO 1.736 [38] Gehlenite 1.658-1.655 [38]
35 Akermanite 1.632-1.64 [38] C-S-H 1.49-1.530 [38]
36
37 Hydrogarnet 1.6041.734 [38] Merwinite 1.708-1.724 [38]
38 Mirabilite 1.394-1.398 [38] Thenardite 1.464-1.485 [38]
39
40 Ettringite 1.462-1.466 [38] Monosulfate 1.488-1.504 [38]
41 Thaumasite 1.468-1.504 [38] Quartz 1.544-1.553 [38]
42
43 Calcite 1.486-1.658 [38] Mullite 1.642-1.654 [38]
44 Magnetite 2.42 [45] Maghemite 2.54 [45]
45
46 Hematite 2.87-3.22 [45] Cristobalite 1.485-1.487 [45]
47 Rutile 2.605-2.908 [45] Anatase 2.488-2.561 [45]
48
49 448 Note – range of values for anisotropic substances
50
51 449 Values for the absorption coefficient k are less frequently reported. For cements
52
53 450 the imaginary parts k of the complex refractive index are reported within the very
54
55 451 wide interval of k = 0.003 to 1.0 [46]. A mean value of 0.1 was used by Hooton
56
57
452 and Buckingham [24] for ordinary Portland cement and fly ash and 0.001 for
58 453 silica fume. Gupta and Wall [47] presented a range for k from 0.005 to 0.01 for
59
60 454 char-free ash from subbituminous coal, while values between 0.005 and 0.05 have
61
62
63 15
64
65
455 also been reported [48,49]. However, those authors (working on the radiative
1
2 456 properties of fly ashes) concluded that it is not acceptable to ignore the
3
4
457 wavelength-dependence of fly ash refractive index and that previous studies
5 458 employing n = 1.5 and k ranging from 0.005 to 0.05 seem to overestimate the
6
7 459 Plank mean absorption coefficient of fly ash particles. Liu and Swithenbank [49]
8
9 460 mentioned that the average value of the imaginary part of the fly ash complex
10
11 461 refractive index k = 0.012 estimated by Gupta and Wall [50] is definitely too high.
12
13 462 Reported values in the literature for the refractive index and the absorption
14
15
463 coefficient of ordinary Portland cement and SCMs are shown in Table 3.
16
17 464 Table 3 Typical refractive index values (real) and absorption index values (imaginary) of cement
18 465 and cementitious materials. OPC – Ordinary Portland Cement; BFS – Blast furnace slag; GFS –
19
20 466 Gasification slag
21
22
Material Refractive Absorption Reference
23 Index, n coefficient, k
24 OPC 1.73 0.1 [51]
25
26
1.73 0.1 [51]
Fly ash
27 1.50 0.005 – 0.05 [49]
28
29
Fly ash, Class F 1.56 1.0 [52]
30 Fly ash, Class C 1.65 0.1 [52]
31
32
BFS, GFS 1.62 1.0 [52]
33 Silica fume 1.53 0.001 [51]
34
35
467
36
37 468 Obscuration level and specific surface area
38
39 469 Another factor that affects the results in the laser light scattering method is the
40
41 470 obscuration, which is related to the sample concentration. Practically, it is the
42 471 fraction of light that is lost from the main beam when the sample is introduced. It
43
44 472 gives a visual indication of how much sample is added. Further, the measurement
45
46 473 is affected by the dispersant that is used for the de-agglomeration of the sample
47
48 474 and the measurement time. The optimum measurement time depends on the size
49
50 475 of the sample and particle size distribution. Normally a measurement time of
51
52
476 10sec is used, with 3 replicates.
53
54 477 The specific surface area (SSA) is then calculated from the particle size
55
56 478 distribution by Eq. 6, assuming that the particles are spherical and non-porous:
57
58
59
60 479 [6]
61
62
63 16
64
65
480 where Vi is the relative volume in class i with mean class diameter of di, ρ is the
1
2 481 particle density, g/cm3 and D[3,2] is the surface weighted mean diameter (μm).
3
4
5 482 3.5 Particle size analysis with microscopy image analysis
6
7 483 Image-based, particle size analysis relies on the principle developed by Medalia
8
9 484 (1970) [53]. Modern image analysis uses scientific cameras to provide low
10
11 485 distortion digital images that are instantaneously processed to extract particle size
12
13 486 and shape.
14
15 487 The imaging setups have to be calibrated in terms of illumination and spatial
16
17 488 resolution. Illumination is usually adjusted through trial and error procedures to
18
19 489 optimize the contrast between the background and the objects to be measured. A
20
21
490 blank image consisting in the imaging of the single background is usually
22 491 acquired to allow spatial correction of the illumination of the field of view.
23
24
25 492 The illumination intensity also needs to fit to the camera sensitivity to guarantee
26
27
493 the shorter exposure time possible. Camera saturation should however always be
28 494 avoided, as it can cause some blurring effect and compromise accurate
29
30 495 measurement. Various illumination geometries can be envisaged. However, the
31
32 496 highest resolution is achieved with an axial back-lighting of the particles [54,55].
33
34 497 Once the illumination setup is fixed, the spatial resolution in the field of view has
35
36 498 to be calibrated. This is usually performed thanks to the imaging of a reference
37
38 499 grate which size is well-known.
39
40 500 Overall imaging setups are commonly classified into two main categories:
41
42 501 dynamic or static type. In dynamic image analysis setups, particles are moving or
43
44 502 free-falling in the field of view of the camera which induces some uncertainties
45
46 503 related to the particle position and orientation. The impact on the measurement is
47
48 504 particularly sensible in the case of elongated particles. In the static setups,
49
50
505 particles are at rest in a plan thus the probability to measure the particle longest
51 506 dimensions is statistically high. The more controlled static setups should be
52
53 507 preferred for accurate measurements [56].
54
55
56
57
58
59
60
61
62
63 17
64
65
508 3.6 Mercury Intrusion Porosimetry (MIP) for Particle Size Distribution
1
2
3
509 Mercury Intrusion Porosimetry (MIP) has formed the basis for a number of
4 510 internationally recognised standard analyses; however it seems, as observed by
5
6 511 León [57], that the full extent of its capabilities remains underexploited.
7
8 512 Generally MIP is used to study the volume, distribution and interconnectivity of
9
10 513 the voids (pores) within porous solid and fine-grained samples, relying on the
11
12 514 Washburn equation to build a picture of the microstructure. On the premise that
13
515 for a given pore radius, a certain pressure is required to intrude a non-wetting
14
15 516 fluid (mercury) and by recording the volume intruded for each pressure increment,
16
17 517 an intrusion curve may be derived. It can be represented by the Washburn
18
19 518 formula:
20
21 519 P = -2γcos(θ)/r [7]
22
23
24 520 Where P is the pressure, γ is the surface tension of the fluid, θ is the contact angle
25
26
521 between the fluid and the material’s surface, and r is the pore radius [57,58].
27
28 522 The Washburn theorem assumes a model of cylindrical pores; the MIP technique
29
30 523 further assumes that these cylinders get progressively smaller towards the inside
31
32
524 of the sample; as pressures increase for further intrusion, it is assumed that this is
33 525 due to progressively smaller pore radii [59]. Naturally, these assumptions are
34
35 526 seldom representative of a porous material used in practice, however the data
36
37 527 produced when compared to data from other materials also assessed under MIP,
38
39 528 can be valuable for depicting trends and interrelationships.
40
41 529 There is little information in the literature about particle size distribution
42
43 530 determined via MIP [57,60,61]. Particle size distribution (PSD) by MIP is derived
44
45 531 from Mayer and Stowe’s [62] relationship established between particle size and
46
47 532 breakthrough pressure required to fill the interstitial voids between a packed bed
48
49
533 of spheres. Following the development of PSD by MIP, Mayer and Stowe
50 534 presented the benefits of the spherical model for characterising certain types of
51
52 535 porous solids, over that of the cylindrical model [63].
53
54
55
536 Employing the same principle of interfacial tensions giving resistant force to an
56 537 intruding, non-wetting liquid, a curve of intrusion vs. applied pressure is
57
58 538 produced. PSD by MIP relies on the premise that this curve contains structural
59
60 539 information about the studied material.
61
62
63 18
64
65
540 MIP uses a model of cylinders for ‘typical’ pore analyses and a model of spheres
1
2 541 for particle size analyses. The particle size distribution as analysed by the MIP
3
4
542 curve is perhaps better termed the “Equivalent Spherical Size Distribution”, as the
5 543 calculated PSD derives from the modelled set of spheres which best represents the
6
7 544 logged experimental data [64].
8
9
10
545 The question of how representative these results are is therefore dependent on
11 546 how similar the particle geometry is to that of a set of spheres. Plate-like or very
12
13 547 angular particles conform less well to the mathematical model than mono-sized
14
15 548 well-rounded particles.
16
17 549 This application of the MIP technique is subject to some criticism arising from,
18
19 550 amongst other aspects, the set of inherent assumptions it relies on. Particle size
20
21 551 distribution is an extension of many of these assumptions, including its own
22
23 552 additional ones, and as such must be undertaken using a considered approach. The
24
25 553 results of the technique display an approximation of the particle size distribution,
26
27
554 rather than a measurement; it provides a ‘feel’ for the characteristics of the
28 555 particles, used along with complementary techniques and viewed in an objective
29
30 556 context.
31
32
33
557 Whilst Huggett et al. [60] considered the assumptions of Mayer-Stowe PSD by
34 558 MIP to be “gross”, seeking to refine the method and presenting their modified
35
36 559 alternative approach, they did indeed find that the Mayer-Stowe technique
37
38 560 presented “a good approximation” of PSD for certain types of samples.
39
40 561 Practically speaking, however crude the mathematical model may be when
41
42 562 compared to the true sample, careful management of certain variables can hold
43
44 563 significant benefit to the overall representivity of the analysis. One such
45
46 564 consideration is that of fine pressure increments as a pragmatic way of increasing
47
48 565 accuracy. Another is the contact angle assumed for analysis; setting a control
49
50
566 contact angle (e.g. [65]), if perhaps a little crude, remains useful for comparing
51 567 data from physically and chemically similar materials analyzed under the same
52
53 568 technique [57].
54
55
56
57
58
59
60
61
62
63 19
64
65
569 4 Materials and Methods for Demonstration Testing
1
2
3 570 4.1 Materials
4
5
6 571 Two batches of fly ash (FA1, FA2), two ground granulated blast furnace slags
7
8 572 (BFS1, BFS2), and a densified silica fume (SF), were used to demonstrate the
9
10 573 methods and techniques discussed in this paper. The chemical compositions of the
11
12
574 materials are presented in Table 4. Scanning electron micrographs providing a
13 575 qualitative indication of the shape and morphology of the materials under
14
15 576 investigation are shown in Figure 2. The fly ash and silica fume particles are
16
17 577 generally spherical, whereas the BFS particles are angular and irregularly shaped.
18
19 578 Table 4 Chemical composition of the materials used in this study, determined via X-ray
20
21 579 fluorescence. Loss on ignition (LOI) was determined at 1000C
22
23
wt.% FA1 FA2 BFS1 BFS2 SF
24 SiO2 54.19 51.37 36.37 33.86 95.60
25 Al2O3 23.5 28.71 9.83 8.91 0.34
26 Fe2O3 7.92 5.10 0.26 0.69 0.11
27
28
CaO 3.02 3.56 41.24 42.64 0.23
29 MgO 1.92 1.01 7.41 7.39 0.37
30 Na2O 1.08 0.29 0.28 0.28 0.28
31 K2O 3.38 1.77 0.41 0.52 0.92
32
33 P2O5 0.27 0.64 / / 0.08
34 SO3 0.94 1.11 1.62 1.62 0.28
35 Cl- 0.003 0.001 0.02 0.01 /
36 Reactive SiO2 41.86 37.48 / / /
37
38 Free CaO 0.1 <0,1 / / /
39 Na-equivalent 3.31 1.46 0.28 0.28 /
40 LOI 1.84 3.6 1.3 ⁄ 1.66
41
42 580
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 20
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 581
27
28 582 Fig. 2 Scanning electron microscope micrograph of (a) fly ash particles at 4500X. Spherical
29
30 583 particles; (b) slag particles at 1100X. Irregular shape. (c) densified silica fume agglomerates at
31 584 230X and d) 22000X. Spherical porous particles, used in this study
32
33
34 585 4.2 Methods
35
36
37
586 Sieving test
38 587 Wet-sieve analyses of the fly ashes, slags, and silica fume were performed
39
40 588 following the procedures described in ASTM C430 [22]. A 1.000 g sample of the
41
42 589 SCM tested was placed on a clean, dry 45 μm (No. 325) sieve, and then a gentle
43
44 590 stream of water was passed through to wet the sample. The water spray nozzle
45
46 591 was then adjusted to 69 ± 4 kPa with the bottom of the nozzle about 12 mm from
47
48
592 the top of the sieve frame. The sample was washed for 1 min, moving the sieve
49 593 with a circular motion in a horizontal plane at the rate of one motion per second in
50
51 594 the spray. Immediately upon removing the sieve from the spray, the sample in the
52
53 595 sieve was rinsed once with about 50 cm3 of deionized water, then the lower
54
55 596 surface of the sieve was blotted gently using a damp cloth. The sieve and sample
56
57 597 residue were dried on a hot plate, where the sieve was supported in a manner that
58
598 air could pass freely beneath it, and then cooled. The residue was weighted after
59
60 599 being brushed from the sieve.
61
62
63 21
64
65
600 Blaine test
1
2 601 The air permeability test was performed according to the method described in the
3
4
602 European Standard EN 196-6:2010 [66]. In the same standard the pycnometer
5 603 method for the determination of the density of the SCMs is also described.
6
7
8 604 BET testing
9
10
605 Nitrogen adsorption–desorption measurements were carried out at 77 K on a
11 606 Micromeretics Tristar 3000 apparatus to determine the Brunauer–Emmett–Teller
12
13 607 (BET) surface area. Before measurement samples were dried and kept in a
14
15 608 dessicator. Samples were cooled with liquid nitrogen and analyzed by measuring
16
17 609 the volume of gas (N2) adsorbed at specific pressures. The pore volume was taken
18
19 610 from the adsorption branch of the isotherm at Pi/P0 = 0.95 assuming complete
20
21
611 pore saturation.
22
23 612 Laser diffraction
24
25 613 LD analysis was performed on a Mastersizer 2000E (Malvern Instruments,
26
27
614 Malvern, UK; 633 nm red laser) with a Hydro 2000SM wet unit (50 - 120 ml).
28 615 The manufacturer specifications state that this instrument is capable of measuring
29
30 616 powders with a size distribution ranging from 0.1 μm to 1000 μm. The chosen
31
32 617 dispersant was isopropanol (IPA). The sample is insoluble in IPA and the
33
34 618 dispersant does not react with the sample itself. An external ultrasonic bath (35
35
36 619 kHz, 320 W) was used for the de-agglomeration of the particles increasing the
37
38
620 dispersion efficiency. The refractive index of IPA was set at 1.39 [67]. A suitable
39 621 method was developed for the determination of particle size distribution of each
40
41 622 material using the light scattering technique. The development of the method is
42
43 623 described in the Results and Discussion section.
44
45 624 Image analysis testing
46
47 625 Static image analysis was performed by means of the Occhio 500 Nano image
48
49 626 analyzer. This instrument includes an integrated vacuum dispersion system and a
50
51 627 high-quality optical component which allows assessment of size and shape of a set
52
53 628 of dispersed particles. A few milligrams of particles were dispersed onto a circular
54
629 glass slide which is moved above a collimated blue (490 nm) LED backlighting.
55
56 630 Pictures of individual particles were captured with a 13921040 pixels video
57
58 631 camera fitted with a telecentric lens. The system routinely determined the
59
60 632 magnification through the imaging of a calibrated grid and the working resolution
61
62
63 22
64
65
633 was determined as 0.563 µm/pixel. The inscribed disk diameter (DIN) of each
1
2 634 particle is calculated in real time to build size distribution curves weighted by
3
4
635 apparent volume (PSD-V’) [68], making the assumption that particles have
5 636 identical densities and flatness ratios, whatever their size.
6
7
8 637 The particle image acquisition method proceeds by scanning the first 50000
9
10
638 particles, ensuring that particles are scanned at least once on the whole diameter
11 639 of the glass slide. The accuracy associated with the estimation of PSD-V,
12
13 640 expressed as the two-sided 95% confidence interval, is computed by the bootstrap
14
15 641 method [68,69].
16
17 642 MIP
18
19 643 The porosimeter used was an AutoPore IV 9500 by Micromeritics, with in-built
20
21 644 Mayer-Stowe data reduction pack. A base filling pressure of 0.5psia is employed
22
23 645 to fill the interstitial voids surrounding the collective sample (undulations across
24
25 646 the ‘surface’ of the grouped specimen). This pressure will fill voids of 360µm –
26
27
647 interstitial or interparticle [57]. This then forms the ‘zero’ value from which the
28 648 analysis proceeds. Samples were first dried in an oven at 75°C, for around two
29
30 649 hours, subsequently acclimatised in a desiccator to minimise moisture uptake
31
32 650 from the environment during cooling. The penetrometer type employed was “3
33
34 651 Bulb, 1.190 Stem, Powder”.
35
36
37 652 5 Results and discussion
38
39
40 653 5.1 Particle size by sieving
41
42
43 654 The results of the wet-sieve analysis are shown in Table 5. All of the materials
44
45 655 pass the ASTM C618 [21] criterion for fly ash and natural pozzolans with less
46
47
656 than 34% of the material retained. Silica fume was not tested using this method
48 657 because the criteria established are only for fly ash and natural pozzolans. Sieve
49
50 658 analysis is not an appropriate test method for assessing fineness of silica fume.
51
52
53
659 Interestingly, one slag sample (BFS1) has a significant higher mass percent
54 660 retained than the other (BFS2). The BFS1 sample had clumps of particles retained
55
56 661 on the sieve after testing, as shown in Figure 3. It can be concluded that clumping
57
58 662 of particles causes potentially erroneous results.It is clear that materials should be
59
60 663 adequately dispersed before testing, a step which is not specified in the ASTM
61
62
63 23
64
65
664 C430 [22] standard. Methods of dispersion discussed in section 2.2 of this paper
1
2 665 may be employed.
3
4 666 Table 5 Percent of material passing the no. 325 (45 μm) sieve on wet-sieving following ASTM
5 667 C430
6
7 Sample ID % retained on sieve
8
9 FA1 20.82
10 FA2 14.67
11
12 BFS1 11.31
13 BFS2 1.52
14 668
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 669
42
43 670 Fig. 3 BFS1 retained on no. 325 (45 μm) sieve after wet-sieving following ASTM C 430
44
45 671 5.2 Surface area by air-permeability
46
47
48 672 In order to measure the surface area by the air-permeability test the density of the
49
50 673 material is needed. The density (Table 6) for the five materials tested was
51
52 674 measured by the pycnometer method. When applying the air-permeability method
53
54
675 to the SCMs, problems occur during the procedure. In the case of blast furnace
55 676 slag, by using the mass as determined in Eq. 1, the adjusted thumb pressure on the
56
57 677 plunger (Fig. 4) is not sufficient to form a bed of porosity 0.500 as required by the
58
59 678 standards. In this case, samples of different mass values were prepared to produce
60
61 679 a test bed with porosity of 0.530±0.005, as calculated using Eq. 3.
62
63 24
64
65
680 Table 6 Density of the SCMs under investigation measured by the pycnometer method
1 Sample ID FA1 FA2 BFS1 BFS2 SF
2
3 Density [g/cm³] 2.136 2.221 2.784 2.924 2.217
4 681
5
6
7
8
9
10
11
12
13
14
15 a b
16 682
17
18 683 Fig. 4 (a) Plunger not in contact with the top of the container (cylinder inside diameter 12.70 ±
19 684 0.10 mm) when blast furnace slag is measured, (b) perfect contact plunger with – top of container
20
21 685
22
23 686 For the densified silica fume it was again impossible to prepare a compacted bed
24
25
687 of porosity 0.500 (Fig. 5 b & c) following the typical procedure as it is described
26 688 in the European Standard EN196-6 [28]. While the pressure applied by the finger
27
28 689 was released, the bed initially formed responded semi-elastically and expanded
29
30 690 again to become unpacked. The time measured in the Blaine apparatus was in the
31
32 691 range of a few seconds (2 - 3 sec) which could not be considered as accurate.
33
34 692
35
36
37
38
39
40
41
42
43 a b c
44 693
45
46 694 Fig. 5 (a) Compacted bed of blast furnace slag. Compacted bed of silica fume, (b) in the container
47 695 after removing the plunger and (c) out of the container
48
49 696
50
51 697 A modified plunger was prepared which was perforated in the bottom surface, as
52
53 698 shown in Fig. 6. This modification was introduced by Teipel and Winter [70] in
54
55
699 order to avoid the dust that is produced when the plunger is removed from the
56 700 permeability cell (Fig. 5b) prior to measurement. The modified plunger does not
57
58 701 need to be removed ensuring the compacted bed formation.
59
60 702
61
62
63 25
64
65
1
2
3
4
5
6
7
8
9 703
10
11 704 Fig. 6 Traditional (right) and modified (left) perforated plunger according to Teipel and Winter
12 705 [70]
13
14 706
15
16 707 The apparatus was calibrated for both plungers in order to determine the K value
17
18 708 (Eq. 2). Since for the case of SCMs an internationally accepted reference material
19
709 is not yet available, three standard reference materials, Ref. 1, Ref. 2 and Ref. 3,
20
21 710 of quartz were used, which each had a density of 2.65 g/cm³ and porosity e=0.50,
22
23 711 only differing in specific surface areas (Table 7). The determination of the volume
24
25 712 of the permeability cell was conducted using mercury as described in EN196-6
26
27 713 [28]. For the calibration of the apparatus it was difficult to obtain two values of
28
29 714 the volume differing by less than 0.005 cm3 as mentioned in the standard
30 715 procedure.
31
32
33 716 Table 7 Specific surface areas of the standard reference materials. Calculated volume, and K value
34 717 for each of the reference materials using the traditional plunger
35
36 Quartz Ref. 1 Ref. 2 Ref. 3
37
38
Specific surface area [cm²/g] 2750 4030 2950
39 Calibration
40
1.865 1.878 1.890
41 Volume [cm³]
42 (±0.007) (±0.007) (±0.005)
43 K 22.90 22.51 23.81
44
45
718
46 719 The second standard reference material, Ref. 2 ((Materialprüfanstalt für Steine
47
48 720 und Erden) was used for the calibration of the apparatus by both traditional and
49
50 721 modified plungers. This reference material had a specific surface area closer to the
51
52 722 SCMs under investigation. In Table 8, the results from the calibration of the
53
54 723 apparatus are shown.
55
56 724 Table 8 Calibration of the apparatus by both, traditional and modified, plungers using Ref. 2
57
58 Original Modified
Parameters
59 plunger plunger
60 K 22.51 25.02
61 V [cm³] 1.878±0.007 1.972±0.017
62
63 26
64
65
725 After calibration, Eq. 3 was used to calculate the mass in order to form a bed of
1
2 726 specific porosity. By using the calculated mass it was impossible to obtain a good
3
4
727 contact of the plunger with the top surface of the permeability cell. Therefore this
5 728 mass was modified by trial and error using Eq. 3. In some cases the porosity of the
6
7 729 bed was increased, as shown in Table 9.
8
9 730 Table 9 Results of the mass used for the formation of a bed with desired porosity
10
11 Original plunger Modified plunger
12
13 Calculated Measured Calculated Measured
14
Sample e e
mass [g] mass [g] mass [g] mass [g]
15 ID
16 FA1 2.0056 2.0056 0.50 2.1060 2.1063 0.50
17 FA2 2.0859 2.0861 0.50 2.1903 2.1902 0.50
18
19 BFS1 2.6142 2.5506 0.51 2.7451 2.6502 0.52
20 SF 2.0822 / / 2.1864 / /
21 731
22
23 732 The results of the specific surface area measured for each material are shown in
24
25 733 Table 10; the measured SSA differs when the measurement is conducted using the
26
27
734 modified plunger. For the silica fume no results were obtained since it was
28 735 impossible to form a well compacted bed, as noted above.
29
30
31 736 Table 10 Specific surface area of SCMs measured using the traditional and the modified plunger
32 2
Specific surface area SSA [m /kg]
33 Sample ID
34 Traditional plunger Modified plunger
35 FA 1 253 240
36
FA 2 367 387
37
38 BFS1 373 369
39 SF / /
40
41 737
42
43
44
738 5.3 Surface area by BET
45
46 739 The surface area of the powders determined by BET method are reported in Table
47
48 740 11. The difference between the surface areas measured by Blaine (Table 10) and
49
50 741 BET (Table 11) could be attributed to the porosity and surface roughness of the
51
52 742 particles. BET surface area is determined by the monolayer coverage of the
53 743 exposed (pores and cracks included) surface of the particles by nitrogen
54
55 744 molecules. If the particles are porous, or have a rough surface structure, the BET
56
57 745 surface area will be greater than the Blaine surface area.
58
59 746
60
61 747
62
63 27
64
65
748 Table 11 Specific surface area measured by BET method
1
BET
2 Sample ID
3 Specific surface area (SSA) [m2/kg]
4 FA1 452
5 FA2 560
6
7
BFS1 1025
8 BFS2 721
9 SF 17102
10
11 749
12
13 750 5.4 Particle size distribution by laser diffraction
14
15
16 751 5.4.1 Method development
17
18 752 Before using a measurement technique for routine analysis, it is necessary to
19
20 753 develop a reliable methodology that maximizes accuracy and precision. For laser
21
22 754 diffraction, there are several parameters that should be optimized to ensure good
23
24 755 results. These include: stirrer rate /air pressure (wet / dry method), ultrasonication
25
26 756 frequency and duration for particle dispersion, measurement time, obscuration
27
28
757 levels and optical parameters. It is essential that the user examines the effects of
29 758 these parameters on the measured values for the instrument used, particularly
30
31 759 when studying materials such as SCMs which are diverse in physicochemical
32
33 760 properties.
34
35 761 For the laser diffraction unit used in this study, the values of the analytical
36
37 762 parameters were varied in order to determine the most appropriate value of each
38
39 763 parameter based on the changes to dv10, dv50, and dv90 for the materials tested.
40
41 764 These percentile diameters, dv10, dv50 and dv90 represent the size (in
42
43 765 micrometers) below which 10%, 50% or 90%, respectively, of the sample falls.
44
45
766 The procedure of the determination of the instrumentation parameters for wet
46 767 measurements is demonstrated for the blast furnace slag only (sample BFS1).
47
48 768 Blast furnace slag was chosen since it has irregularly shaped but chemically
49
50 769 homogeneous particles (Fig. 2B), and thus provides a relatively well-characterized
51
52 770 deviation from the assumptions inherent in the technique. A similar procedure was
53
54 771 followed for evaluation of the fly ash as presented in ref. [71]. According to
55
772 Arvaniti et al [71] the measurement parameters used for fly ash were: 2 min
56
57 773 sonication time, 20 sec measurement time, 1700 rpm stirrer rate, 15% - 25%
58
59 774 obscuration limits and two pairs of optical parameters n=1.65, k=0 or 0.001 and
60
61 775 n=1.73, k=0 or 0.001.
62
63 28
64
65
776 Wet dispersion
1
2 777 Isopropanol was used as a dispersant since it does not react with slag. Different
3
4 778 sonication times were used and the dv10, dv50 and dv90 before and after
5
6 779 sonication are shown in Figure 7. De-agglomeration of the particles is observed
7
8 780 after 2 min in the ultrasonic bath, and after longer durations of sonication (4-6
9
10
781 min) the dv10 and dv50 essentially stabilize, while fluctuations occur in dv90
11 782 because of re-agglomeration. An ultrasonication time of 2 min was chosen as the
12
13 783 ideal condition to disperse the slag (BFS1) used in this study.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 784
30 785 Fig. 7 Variation of the particle size fractions of anhydrous blast furnace slag dispersed in
31 786 isopropanol as function of the time of ultrasonication. Stirring speed: 1700 rpm; obscuration: 10%-
32
33 787 15%; measurement time: 20 sec; n=1.62, k=1
34
35 788 Stirring speed
36
37 789 Figure 8 shows the particle size distribution changes when using different stirrer
38
39 790 rates in the laser diffraction instrument. Each point shown on the plot corresponds
40
41 791 to a single measurement, and six consecutive measurements were taken for each
42
43
792 stirrer rate. At low rates (500 rpm) the particle sizes are lower than observed when
44 793 increasing the stirring rate. This is consistent with the fact that a slow stirring rate
45
46 794 does not provide the force required to suspend the larger particles within the
47
48 795 sample. The observed increase in dv90 at higher stirrer rates (3000 rpm) could be
49
50 796 attributed to the formation of bubbles in the dispersant. From 1000 rpm to 2500
51
52 797 rpm, the particle sizes are relatively stable, indicating that the particles are
53
798 correctly presented into the measurement cell. The stirrer rate is therefore
54
55 799 recommended to be set at a value at the center of this range (i.e. at around 1700
56
57 800 rpm for this example), which also agrees with the suggestions in [72].
58
59
60
61
62
63 29
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 801
17
18 802 Fig. 8 Variation of the particle size fractions of anhydrous blast furnace slag dispersed in
19 803 isopropanol as function of the stirrer rate. Sonication time: 2 min; obscuration: 10%-15%;
20
21 804 measurement time: 20 sec; n=1.62, k=1
22
23 805 Measurement time
24
25 806 To verify the effect of the duration of measurements, particles were analyzed
26
27
807 using measurement times of 10, 20, 30 and 40 sec. Figure 9 shows that the
28 808 measurement time does not have an effect on the values of dv10 and dv50.
29
30 809 However, a slight fluctuation is noticeable in dv90 across the whole range of
31
32 810 measurement times. Based on these data, at standard measurement time of 20 sec
33
34 811 was chosen.
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 812
52 813 Fig. 9 Influence of the measurement time on dv10, dv50 and dv90 for blast furnace slag. Sonication
53
54 814 time: 2 min; stirrer rate: 1700 rpm; obscuration: 10%-15%; n=1.62, k=1
55
56 815 Sample concentration
57
58 816 The obscuration rate defines the quantity of the sample that is added in the
59
817 measurement cell. If the amount of sample added to the dispersion unit is too low
60
61 818 then the results will not be reproducible due to a low signal to noise ratio. If the
62
63 30
64
65
819 obscuration is too high then the measurement becomes affected by multiple
1
2 820 scattering, causing a reduction in the measured particle size at higher
3
4
821 obscurations. An obscuration titration was carried out, and the measured particle
5 822 size parameters as a function of obscuration are shown in Figure 10. The limits
6
7 823 were set as indicated in Figure 10 (e.g. limits of 5%-6% mean that a fraction of
8
9 824 5%-6% of the light is lost from the main beam when the sample is introduced),
10
11 825 and the sample was added until a value in between these limits is reached. Then
12
13 826 the particle size measurements started automatically. Increased obscuration limits
14
15
827 seem to increase the particle size of dv90 and dv50, but it does not seem to have a
16 828 significant effect in the dv10 values.
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 829
34 830 Fig. 10 Variation of the particle size fractions of blast furnace slag as function of the obscuration
35
36 831 titrations values. Sonication time: 2 min; stirrer rate: 1700 rpm; measurement time: 20 sec; n=1.62,
37 832 k=1
38
39 833 Based on the fact that the particle size should be independent of the obscuration
40
41 834 within a given concentration range, the limits 9% - 15% showing the minimum
42
43 835 fluctuations of the results were chosen.
44
45
46
836 Influence of optical parameters on particle size analysis
47 837 The light scattering data obtained from the measurements were analyzed using the
48
49 838 Mie model. Since there are not many data available in the literature concerning
50
51 839 the optical properties of slags, the refractive indices published for cement and
52
53 840 other cementitious materials were used. The raw data from laser diffraction were
54
55 841 analyzed using the real refractive indices, n: 1.50, 1.56, 1.65 and 1.73 and the
56 842 imaginary absorption coefficient, k: 0, 0.001, 0.005, 0.01, 0.05, 0.1 and 1. Each n
57
58 843 was combined with each k giving a total of 28 optical models which were used for
59
60
61
62
63 31
64
65
844 analysis. The corresponding results are shown in Fig. 11. Each point corresponds
1
2 845 to one optical model.
3
4
5
6
7
8
9
10
11
12
13 kk
14
15
16 nn
17 846
18
19
20
21
22
23
24
25
26
27
28
29
30
31 847
32
33
34
35
36
37
38
39
40
41
k
42
43
44
n
848
45
46 849 Fig. 11 Influence of optical parameters on calculated dv10, dv50 and dv90 for BFS1 sample.
47 850 Sonication time: 2 min; stirrer rate: 1700 rpm; measurement time: 20 sec; obscuration limits: 9% -
48
49 851 15%
50
51 852 Using different n and k values modifies the size of the smaller particles registered
52
53 853 (dv10 and dv50) in the analyzed slag. Using a small refractive index (n=1.50),
54
55 854 higher particle sizes are registered with increased k values. Conversely, when
56 855 increasing the assumed refractive index from 1.56 to 1.73, higher k values tend to
57
58 856 reduce the particle size. This effect is more evident in the dv10 fraction. The
59
60 857 significant variation in the particle size recorded using different optical parameters
61
62
63 32
64
65
858 highlights the importance of detailed optical characterization of the SCMs prior to
1
2 859 laser diffraction analysis to ensure that the particle size distribution results are
3
4
860 representative.
5
6 861 The volume weighted percentiles dv10, dv50, dv90 and the specific surface area
7
8 862 (Table 12 ) were measured using the LD method with the appropriate parameters
9
10
863 for the different materials. For FA1 sample specific surface area is higher than the
11 864 Blaine surface area (Table 10) and lower than BET surface area (Table 11). FA2
12
13 865 sample gave lower SSA when measured with LD method (Table 12) compared to
14
15 866 the Blaine (Table 10) and BET (Table 11) surface area. The d v10, dv50 and dv90
16
17 867 that are measured for each fly ash sample using different optical parameters show
18
19 868 slightly different results which are more profound for FA1 sample.
20
21 869 Blast furnace slag samples give much lower SSA measured by LD (Table 12),
22
23 870 than the ones measured by BET (Table 11) and Blaine (Table 10). By changing
24
25 871 the absorption index from k=0 to k=0.1 the calculated SSA is double and the
26
27
872 particle size is lower. The optical parameters were chosen based on the results
28 873 illustrated in figure 11. The n=1.56 and k=0 or k=0.1 give the most stable results
29
30 874 for dv10, dv50 and dv90.
31
32
33
875 In the case of SF the determination of PSD by LD method gave a bimodal
34 876 distribution (Figure 12). The calculated SSA (Table 12) was much lower than the
35
36 877 BET (Table 11) surface area. The large particles (dv10, dv50, dv90) of SF sample
37
38 878 could be the result of insufficient dispersion of the sample before the
39
40 879 measurement. Although the sonication time was extended to 55 min it was
41
42 880 impossible with isopropanol to get separated single particles. For particles smaller
43
44
881 than 1 µm, it is difficult to overcome the Van Der Waals forces and get particles
45 882 separated.
46
47
48 883 The R45, which represents the percentage of material retained on a no. 325 (45
49
50
884 μm) sieve, was also calculated by the LD method. Comparing the results in Table
51 885 12 with the ones that were obtained using the wet sieving method (Table 5), only
52
53 886 FA2 sample gives calculated (LD method) values close to the measured (wet
54
55 887 sieving method) ones. For the other samples, except silica fume which was not
56
57 888 measured by wet sieving method, the calculated values are lower than the
58
59 889 measured ones.
60
61
62
63 33
64
65
890 Table 12 Volume weighted percentiles, specific surface area and R45 for fly ash, blast furnace
1 891 slag and silica fume samples calculated by using different optical parameters. Samples were
2
3 892 measured with the LD method.
4
5 SSA
dv10 dv50 dv90 R45 (%)
6 [m2/kg]
7
8
FA1 RI=1.65, AI=0.001 3.068 11.215 62.699 15.19 384
9 RI=1.73, AI=0.001 2.800 10.869 61.768 14.87 415
10 FA2 RI=1.65, AI=0.001 4.296 14.944 53.907 14.05 279
11
12 RI=1.73, AI=0.001 4.110 14.719 53.633 13.90 293
13 BFS1 RI=1.56, AI=0.1 1.233 5.625 16.618 0.26 682
14 RI=1.56, AI=0 2.873 7.316 20.090 0.30 370
15
16 BFS2 RI=1.56, AI=0.1 1.173 6.402 19.717 0.45 640
17 RI=1.73, AI=0.001 2.262 7.544 20.950 0.42 402
18 4.07
19
SF RI=1.53, AI=0.001 0.187 0.418 19.090 6180
20 893
21
22
9
23
24 8
25 7
26
6
Volume [%]

27
28 5
29 4
30
31 3
32 2
33
34 1
35 0
36 1 10 100
37 894 Particle size [μm]
38
39 895 Fig. 12 Particle size distribution of SF as determined by laser diffraction. Sonication time: 55 min;
40 896 stirrer rate: 2000 rpm; measurement time: 20 sec; obscuration limits: 5% - 10%; n=1.53, k=0.001
41
42
43
897
44
45 898 5.5 Particle size and shape by image analysis
46
47
48 899 The volume weighted percentiles dv10, dv50, dv90 of two samples of blast furnace
49
50 900 slag (BFS1) performed via static image analysis (SIA) are reported in Table 13;
51
52 901 dvMAX corresponds to the largest particle identified. The corresponding
53
54
902 cumulative curve of this material is shown in Figure 13.
55 903
56
57 904
58
59 905
60
61 906
62
63 34
64
65
907 Table 13 Volume weighted Xth percentiles of the particle diameter of BFS1 obtained by SIA.
1
Percentile Series 1 Series 2
2
3 dv10 [µm] 10.66 9.00
4
5 dv50 [µm] 35.00 35.06
6
7 dv90 [µm] 65.30 64.65
8
9 dvMAX [µm] 99.44 95.75
10
11 908
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 909
31 910
32
33 911 Fig. 13 Particle size distribution of BFS1 as determined by image analysis
34
35 912 From image analysis, particle shape parameters such as elongation and bluntness
36
37 913 can easily be calculated. The particle elongation appears to be quite low with 90
38
39 914 percent of the particles getting an elongation ratio lower than 0.40. Regarding
40
41 915 particle bluntness, it is necessary to consider narrow size fractions because the
42
43 916 numbers of pixels involved in the calculation of this parameter will strongly
44 917 impact the results. For the same reason, it is not relevant to calculate bluntness
45
46 918 from a particle with less than 100 pixels. Figure 13 shows that some particles look
47
48 919 rounded and blunt while others are rougher. However, apparent roughness may
49
50 920 result from the blind analysis of two touching particles as shown in figure 13 (left
51
52 921 hand bottom corner). In such case, an additional filtering may be necessary. The
53
54
922 diameter of the red disk is proportional to the area of the particle. The “O.Blu
55 923 ntness” refers to Occhio’s definition of Bluntness [54].
56
57 924
58
59 925
60
61 926
62
63 35
64
65
927
1
2 928
3
4
929
5 930
6
7 931
8
9 932
10
11 933
12
13 934
14
15
935
16 936
17
18 937
19
20 938 Fig. 13 Shape analysis of BFS1 particles, size fraction >25µm
21
22 939 Differences between PSD and SIA results are usually attributed to insufficient
23
24 940 dispersion of the fine particles in the case of SIA, or inaccuracies in the optical
25
26 941 model in the case of laser diffraction [73]. In this study, different optical models
27
28 942 were used when analyzing the particle size of the slag via laser diffraction, and the
29 943 results derived from this technique are three times smaller than the ones obtained
30
31 944 applying SIA.
32
33
34 945 5.6 Particle Size Distribution determined using MIP
35
36
37 946 Limitations of the MIP technique are evident where considering a substance
38
39 947 where inter- and intra-particle voids are of a similar size, and distinction between
40 948 the volume of mercury intruded into the pores of the particle, and into the spaces
41
42 949 between the particles cannot practically be made. However, for the majority of
43
44 950 powder samples, it may be expected that while their inter-particle voids may be
45
46 951 high (spaces between particles), their intra-particle voids (spaces within particles)
47
48 952 are typically small; and hence, the distinction tends to be simple.
49
50 953 Mercury intrusion does not measure particle sizes in the way that techniques such
51
52 954 as laser diffraction and image analysis do. In MIP technique it is assumed that
53
54 955 particles have the same shape, spherical, and that they are evenly packed. The
55
56 956 mean particle size is estimated by the pressure it takes to fill the interstitial
57 957 volume with mercury. In other words, the particle size distribution derived from
58
59 958 this method is the size distribution of spheres that, when applied to the
60
61
62
63 36
64
65
959 mathematical model, most closely reproduce the experimental penetration data.
1
2 960 The size unit is ‘equivalent spherical size’.
3
4 961 Results obtained from MIP analysis of the fly ash, blast furnace slag and silica
5
6 962 fume are shown in Figure 14. The PSD graphs indicate a predominant particle size
7
8 963 around the 5-10µm range for FA1(Figure 14a), FA2 (Figure 14b), BFS1 (Figure
9
10
964 14c) and BFS2 (Figure 14d) samples. The silica fume PSD indicates a
11 965 predominant particle size centring around the 100-200 nm (Fig. 14e). Comparing
12
13 966 the results for SF from LD (Figure 12) and MIP (Figure 14) can be seen that in the
14
15 967 first method a bimodal distribution was found and measured particles have a
16
17 968 predominant size of 10 μm, whereas in the second method smaller particles of 0.1
18
19 969 μm are measured. In all the samples analyzed, the PSD peaks were sharply
20
21
970 defined, suggestive of a rather monosized (relative to the size spectrum) particle
22 971 arrangement.
23
24 972
25
26 1.20
Intrusion Volume (mL/g)

27 a
1.00
28
29 0.80
30
31 0.60
32 0.40
33
34 0.20
35
0.00
36
1 10 100 1000 10000 100000 1000000
37 Particle Equivalent Spherical Size Dia.(nm)
38 973
39 1.20
Intrusion Volume (mL/g)

40 b
41 1.00
42 0.80
43
44 0.60
45
0.40
46
47 0.20
48
49 0.00
50 1 10 100 1000 10000 100000 1000000
974 Particle Equivalent Spherical Size Dia.(nm)
51
52
53
54
55
56
57
58
59
60
61
62
63 37
64
65
0.90

Intrusion Volume (mL/g)


1 0.80 c
2 0.70
3 0.60
4 0.50
5
0.40
6
7 0.30
8 0.20
9 0.10
10 0.00
11 1 10 100 1000 10000 100000 1000000
12 975 Particle Equivalent Spherical Size Dia.(nm)
13
1.80
14
Intrusion Volume (mL/g)

15 1.60 d
16 1.40
17 1.20
18 1.00
19 0.80
20
0.60
21
22 0.40
23 0.20
24 0.00
25 1 10 100 1000 10000 100000 1000000
26 Particle Equivalent Spherical Size Dia.(nm)
27
976
28 2.50
Intrusion Volume (mL/g)

29 e
30 2.00
31
32 1.50
33
34 1.00
35
36 0.50
37
38 0.00
39 1 10 100 1000 10000 100000 1000000
40 977 Particle Equivalent Spherical Size Dia.(nm)
41
42 978 Fig. 14 Particle size distribution of (a) BFS1, (b) BFS2 and (c) FA1, (d) FA2 and (e) SF samples
43
44 979 using MIP.
45
46
47 980 6 Conclusions
48
49
50 981 The use of SCMs has become widespread in the concrete and cement industry.
51 982 The shape and size peculiarities of SCMs differentiate them from cement, and
52
53 983 therefore the techniques that are currently used for the physical characterization of
54
55 984 cement may not be directly applicable to SCMs. Considering particle size
56
57 985 distribution as one of the most important parameters for the optimization of SCM
58
59 986 utilization, several techniques that are used for the determination of PSD have
60
61
987 been investigated in this study. Some of them are more widely used in cement (i.e.
62
63 38
64
65
988 air permeability test, sieving, laser diffraction, BET, image analysis) while MIP is
1
2 989 not generally applied to measure particle size distributions, but seems to be a
3
4
990 promising technique for the PSD determination of SCMs. However, it must be
5 991 realized that particle size analysis is not an objective in itself but is a means to
6
7 992 correlate powder properties with some process of manufacture, usage or
8
9 993 preparation.
10
11 994 Samples of fly ash, blast furnace slag and silica fume were measured using the
12
13 995 techniques presented in this study and the following conclusions are drawn:
14
15
16 996  Materials should be adequately dispersed before testing.
17
18 997  When applying an air permeability method to the SCMs, problems occur
19
20 998 during the procedure related to the difficulty to form a good compacted bed of
21
22 999 specific porosity.
23
24 1000  For particles that are porous, or have a rough surface structure, the BET surface
25
26 1001 area is found to be greater than the Blaine surface area.
27
28 1002  Wet sieving overestimated the results because agglomerated particles remained
29
30 1003 on the sieve. The wet sieving method could not be used for silica fume since
31
32 1004 the criteria established for this method are only valid for fly ash and natural
33
34 1005 pozzolans.
35
36 1006  For the wet LD technique a method was developed for the measurement of
37
38 1007 SCMs. The optimized parameters for the sample of blast furnace slag are
39
40 1008 demonstrated in this study. The analysis presented here indicated that results
41
42 1009 obtained by LD method are strongly influenced by the optical parameters of the
43
44 1010 material that is measured.
45
46 1011  Analysis of microscope images of blast furnace slag gave particle sizes three
47
48 1012 times larger than the ones obtained by laser diffraction.
49
50 1013  MIP is also presented as a method for the determination of particle size
51
52 1014 distribution of SCMs giving quite satisfactory results. The particle size
53
54 1015 distribution derived from this method is the size distribution of spheres that,
55
56 1016 when applied to the mathematical model, most closely reproduce the
57
58 1017 experimental penetration data.
59
60
61
62
63 39
64
65
1018  Accuracy can be difficult to define for size analysis of non-spherical particles;
1
2 1019 all sizing techniques give different answers. For irregularly shaped particles,
3
4 1020 characterization of particle size must include information on particle shape.
5
6 1021
7
8 1022
9
10 1023 References
11
12 1024 [1] Lothenbach B, Scrivener K, Hooton RD (2011) Supplementary cementitious
13 1025 materials. Cement and Concrete Research 41 (3):217-229.
14
15 1026 [2] Snellings R, Mertens G, Elsen J (2012) Supplementary Cementitious
16 1027 Materials. Reviews in Mineralogy & Geochemistry 74:211-278.
17
18 1028 [3] Siddique R, Khan MI (2011) Supplementary Cementing Materials. Springer.
19
20 1029 [4] Celik IB (2009) The effects of particle size distribution and surface area upon
21 1030 cement strength development. Powder Technol 188 (3):272-276.
22
23 1031 [5] Scrivener KL, Kirkpatrick RJ (2008) Innovation in use and research on
24 1032 cementitious material. Cement and Concrete Research 38 (2):128-136.
25
26 1033 [6] Ursula Stark AM Particle size distribution of cements and mineral admixtures
27 1034 - Standard and sophisticated measurements. In: Owens DGGaG (ed) 11th
28 1035 International Congress on the Chemistry of Cement (ICCC), Durban, South
29 1036 Africa, 11-16 May 2003 2003.
30
31 1037 [7] Naito M, Hayakawa O, Nakahira K, Mori H, Tsubaki J (1998) Effect of
32 1038 particle shape on the particle size distribution measured with commercial
33 1039 equipment. Powder Technology 100:52-60.
34
35 1040 [8] Juenger M, Provis JL, Elsen J, Matthes W, Hooton RD, Duchesne J, Courard
36 1041 L, He H, Michel F, Snellings R, Belie ND (2012) Supplementary Cementitious
37
38 1042 Materials for Concrete: Characterization Needs. MRS Proceedings 1488.
39 1043 [9] Hewlett PC (1998) Lea's Chemistry of Cement and Concrete, 4th Ed. Elsevier,
40
41 1044 Oxford, UK.
42 1045 [10] Allen T (1997) Particle size measurement, vol 1. 5th edn. Chapman and Hall,
43
44
1046 London.
45 1047 [11] Allen T (2003) Powder Sampling and Particle Size Determination.
46
47 1048 [12] ISO 14488:2007 (2007) Particulate materials -- Sampling and sample
48 1049 splitting for the determination of particulate properties.
49
50 1050 [13] ASTM C 183 – 08 (2008) Standard Practice for Sampling and the Amount of
51 1051 Testing of Hydraulic Cement.
52
53 1052 [14] ASTM C311/C311M − 13 (2013) Standard Test Methods for Sampling and
54 1053 Testing Fly Ash or Natural Pozzolans for Use in Portland-Cement Concrete.
55
56 1054 [15] Green DW (1997) Perry's Chemical Engineers' Handbook. 7th edn. McGraw-
57 1055 Hill, London.
58
59 1056 [16] Parfitt GD (1977) The dispersion of powders in liquids—an introduction.
60 1057 Powder Technology 17 (2):157-162.
61
62
63 40
64
65
1058 [17] Takeo Mitsui BS, Susumu Takada BS (1969) On Factors influencing
1 1059 dispersibility and wettability of powder in water. J. Soc. Cosmetic Chemists
2 1060 20:335-351.
3
4 1061 [18] Ferraris CF, Hackley VA, Aviles AI, Charles E. Buchanan J (2002) Analysis
5 1062 of the ASTM Round-Robin Test on Particle Size Distribution of Portland Cement:
6
7
1063 Phase I. NISTIR 6883.
8 1064 [19] Clausen L, Fabricius I (2000) BET Measurements: Outgassing of Minerals.
9
1065 Journal of Colloid and Interface Science 227 (1):7-15.
10
11 1066 [20] Pierotti R, Rouquerol J (1985) Reporting physisorption data for gas/solid
12 1067 systems with special reference to the determination of surface area and porosity.
13
14 1068 Pure Appl Chem 57 (4):603-619.
15 1069 [21] ASTM C618-12a (2012) Standard Specification for Coal Fly Ash and Raw or
16
17 1070 Calcined Natural Pozzolan for Use in Concrete.
18 1071 [22] ASTM C430-08 (2008) Fineness of Hydraulic Cement by the 45-μm (No.
19
20 1072 325) Sieve.
21 1073 [23] Hooton RD, Buckingham JHP The Use and Standardization of Rapid Test
22
23
1074 Methods for Measuring the Carbon Content and Fineness of Fly Ash Pozzolan. In:
24 1075 7th International Ash Utilization Symposium, Orlando, Florida, March 4-7, 1985.
25 1076 Proceedings of the 7th International Ash Utilization Symposium.
26
27 1077 [24] BS4550:1978 4550:1978 (1978) Methods of testing cements.
28
29
1078 [25] Niesel K ( 1973) Determination of the specific surface by measurement of
30 1079 permeability. 6 (3):227-231.
31 1080 [26] Schulz NF (1974) Measurement of surface areas by permeametry.
32
33 1081 International Journal of Mineral Processing 1 (1):65-79.
34 1082 [27] ASTM C204-07 (2007) Standard test methods for fineness of hydraulic
35
36 1083 cement by air-permeability apparatus.
37 1084 [28] EN 196-6 (2010) Methods of testing cement: Determination of fineness
38
39 1085 [29] Potgieter JH, Strydom CA (1996) An investigation into the correlation
40 1086 between different surface area determination techniques applied to various
41
42
1087 limestone-related compounds. Cement and Concrete Research 26 (11):1613-1617.
43 1088 [30] Iyer RS, Stanmore BR (1995) Surface area of flyashes. Cement and Concrete
44
45
1089 Research 25 (7):1403-1405.
46 1090 [31] Kiattikomol K, Jaturapitakkul C, Songpiriyakij S, Chutubtim S (2001) A
47
48
1091 study of ground coarse fly ashes with different finenesses from various sources as
49 1092 pozzolanic materials. Cement and Concrete Composites 23 (4–5):335-343.
50
1093 [32] Frías M, Sánchez de Rojas MI, Luxán Md, García N (1991) Determination of
51
52 1094 specific surface area by the laser diffraction technique. Comparison with the
53 1095 Blaine permeability method. Cement and Concrete Research 21 (5):709-717.
54
55 1096 [33] ISO 13320:2009 (2009) Particle Size Analysis - Laser Diffraction Methods.
56 1097 Part I: General Principles.
57
58 1098 [34] Ferraris CF, Hackley VA, Aviles AI (2004) Measurement of particle size
59 1099 distribution in Portland cement powder: Analysis of ASTM round robin studies.
60 1100 Journal of Cement, Concrete and Aggregates 26 (2).
61
62
63 41
64
65
1101 [35] Wriedt T (1998) A Review of Elastic Light Scattering Theories. Particle &
1 1102 Particle Systems Characterization 15:67–74.
2
3 1103 [36] Wahlstrom EE (1979) Optical Crystallography 5th edn. Wiley, New York
4
5 1104 [37] Mitra S (1996) Fundamentals of optical, spectroscopic and X-Ray
6 1105 mineralogy. New Age International:336.
7
8
1106 [38] Glossary of Minerals.
9 1107 http://freeit.free.fr/Knovel/Concrete%20Petrography%20-
10 1108 %20A%20Handbook%20of%20Investigative%20Techniques/92669_gl.
11
12 1109 [39] Hewlett P (2003) Lea’s Chemistry of Cement and Concrete. 4th edn.
13 1110 Butterworth Heinemann.
14
15 1111 [40] Fujiwara H (2007) Spectroscopic ellipsometry principles and applications.
16 1112 John Wiley and Sons, England.
17
18 1113 [41] Hackley VA, Lum LS, Gintautas V, Ferraris CF (2004) Particle size analysis
19 1114 by laser diffraction spectrometry: application to cementitious powders.
20
21 1115 [42] Zhang HJ, Xu GD (1992) The effect of particle refractive index on size
22 1116 measurement. Powder Technology 70:189-192.
23
24 1117 [43] Ferraris CF, Bullard JW, Hackley V Particle size distribution by laser
25 1118 diffraction spectrometry: application to cementitious powders. In: World congress
26 1119 on particle technology, Orlando, 2006. Proceedings of the 5th world congress on
27 1120 particle technology.
28
29 1121 [44] Taylor HFW (ed) (1997) Cement chemistry. 2nd edition edn.
30
31 1122 [45] www.webmineral.com.
32
33
1123 [46] Mykhaylo P. Gorsky PPM, Andrew P. Maksimyak (2010) Optical correlation
34 1124 technique for cement particle size measurements. Optica Applicata XL (2).
35
1125 [47] Gupta RP, Wall TF (1985) The Optical Properties of Fly Ash in Coal Fired
36
37 1126 Furnaces. Combustion and Flame 61:145-151.
38 1127 [48] Goodwin DG, Mitchner M (1989) Flyash radiative properties and effects on
39
40 1128 radiative heat transfer in coal-fired systems. Inl. J. Heat Mass Transfer. 32
41 1129 (4):627-638.
42
43 1130 [49] Liu F, Swithenbank J (1993) The effects of particle size distribution and
44 1131 refractive index on fly-ash radiative properties using a simplified approach.
45 1132 International Journal of Heat and Mass Transfer 36 (7):1905-1912.
46
47 1133 [50] Gupta RP, Wall TF (1981) The complex refractive index of particles. J. Phys.
48 1134 D: Appl. Phys. 14:95-98.
49
50 1135 [51] Cyr M, Tagnit-Hamou A (2001) Particle size distribution of fine powders by
51 1136 LASER diffraction spectrometry. Case of cementitious materials. Mater Struct 34
52 1137 (6):342-350.
53
54 1138 [52] Jewell RB, Rathbone RF (2009) Optical properties of coal combustion
55 1139 byproducts for particle-size analysis by laser diffraction. Combustion and
56 1140 Gasification Products 1:1-6.
57
58 1141 [53] Medalia A (1970 ) Dynamic shape factors of particles. Powder Technology
59 1142 4:117–138.
60
61
62
63 42
64
65
1143 [54] Leroy S, Dislaire G, Bastin D, Pirard E (2011) Optical analysis of particle
1 1144 size and chromite liberation from pulp samples of a UG2 ore regrinding circuit.
2 1145 Minerals Engineering 24:1340–1347.
3
4 1146 [55] Gregoire M, Dislaire G, Pirard E (2007) Accuracy of Size Distributions
5 1147 Obtained from Single Particle Static Digital Image Analysis. Partec, Nurnberg,
6
7 1148 [56] Hart JR, Zhu Y, Pirard E (2009) Advances in the Characterization of
8 1149 Industrial Minerals. Ch 4: Particle size and shape characterization: current
9
1150 technology and practice.
10
11 1151 [57] León CA (1998) New perspectives in mercury porosimetry. Advances in
12 1152 Colloid and Interface Science 76-77:341-372.
13
14 1153 [58] Glass HJ, With G (1996) Reliability and reproducability of mercury intrusion
15 1154 porosimetry. Journal of the European Ceramic Society 17: 753-757.
16
17 1155 [59] Giesche H (2006) Mercury porosimetry: A general (practical) overview. Part.
18 1156 Part. Syst. Charact. 23:9-19.
19
20 1157 [60] Huggett S, Mathews P, Matthews T (1999) Estimating particle size
21 1158 distributions from a network model of porous media. Powder Technology
22
23
1159 104:169-179.
24 1160 [61] Lucarelli L (2013) The use of mercury intrusion porosimetry for the
25
26
1161 determination of particle size distribution on nano-particles carbon black.
27 1162 http://s3.ceelantech.com/docs/MercuryIntrusionPorosimetry.pdf [Accessed
28 1163 October, 21st 2013].
29
30 1164 [62] Mayer RP, Stowe RA (1965). Journal of Colloid Science 20:893-911.
31 1165 [63] Mayer RP, Stowe RA (2005) Packed uniform sphere model for solids:
32
33 1166 Interstitial access opening sizes and pressure deficiencies for wetting liquids with
34 1167 comparison to reported experimental results. Journal of Colloid and Interface
35 1168 Science 294:139-150.
36
37 1169 [64] Webb PA ( 2001) An introduction to the physical characterization of
38 1170 materials by mercury intrusion porosimetry with emphasis on reduction and
39 1171 presentation of experimental data. Micromeritics Instrument Corporation,
40
41
1172 Georgia.
42 1173 [65] Drake LC, Ritter HL (1945) Pressure porosimeter and determination of
43
1174 complete macropore-size distributions. Industrial & Engineering Chemistry
44
45 1175 Analytical Edition 17:782-786.
46 1176 [66] EU EN 196-6 Methods of testing cement - Part 6: determination of fineness.
47
48 1177 [67] Sample Dispersion and Refractive Index Guide. (1.0)
49
50 1178 [68] Gregoire MP, Dislaire G, Pirard E Accuracy of size distributions obtained
51 1179 from single particle static digital image analysis. In: Proc. Partec Conf. Nürenberg
52 1180 2007. p 4.
53
54 1181 [69] Michel F, Gregoire M, Pirard E Size distribution of powders in range of 1-
55 1182 100 µm: a comparison of static digital image analysis and laser diffraction. In:
56
57
1183 Proc. Partec Conf. Nürenberg 2007. p 4.
58 1184 [70] Teipel U, Winter H (2011) Characterization of the specific surface area with
59
60
1185 the permeation method. At Mineral Processing 52.
61
62
63 43
64
65
1186 [71] Arvaniti EC, Belie ND Method development for the particle size analysis of
1 1187 supplementary cementitious materials. In: XIII DBMC, Sao Paulo, Brazil, 2-5
2 1188 September 2014 (to be published).
3
4 1189 [72] Quercia G, Hüsken G, Brouwers HJH (2012) Water demand of amorphous
5 1190 nano silica and its impact on the workability of cement paste. Cement and
6
7
1191 Concrete Research 42 (2):344-357.
8 1192 [73] Tinke AP, Carnicer A, Govoreanu R, Scheltjens G, Lauwerysen L, Mertens
9
1193 N, Vanhoutte K, Brewster ME (2008) Particle shape and orientation in laser
10
11 1194 diffraction and static image analysis size distribution analysis of micrometer sized
12 1195 rectangular particles. Powder Technology 2 (186):154-167.
13
14 1196
15 1197
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 44
64
65
View publication stats

You might also like